Stirling numbers of the second kind

Last updated

The 15 partitions of a 4-element set ordered in a Hasse diagram
There are S(4,1), ..., S(4, 4) = 1, 7, 6, 1 partitions containing 1, 2, 3, 4 sets. Set partitions 4; Hasse; circles.svg
The 15 partitions of a 4-element set ordered in a Hasse diagram
There are S(4,1), ..., S(4, 4) = 1, 7, 6, 1 partitions containing 1, 2, 3, 4 sets.

In mathematics, particularly in combinatorics, a Stirling number of the second kind (or Stirling partition number) is the number of ways to partition a set of n objects into k non-empty subsets and is denoted by or . [1] Stirling numbers of the second kind occur in the field of mathematics called combinatorics and the study of partitions. They are named after James Stirling.

Contents

The Stirling numbers of the first and second kind can be understood as inverses of one another when viewed as triangular matrices. This article is devoted to specifics of Stirling numbers of the second kind. Identities linking the two kinds appear in the article on Stirling numbers.

Definition

The Stirling numbers of the second kind, written or or with other notations, count the number of ways to partition a set of labelled objects into nonempty unlabelled subsets. Equivalently, they count the number of different equivalence relations with precisely equivalence classes that can be defined on an element set. In fact, there is a bijection between the set of partitions and the set of equivalence relations on a given set. Obviously,

for n ≥ 0, and for n ≥ 1,

as the only way to partition an n-element set into n parts is to put each element of the set into its own part, and the only way to partition a nonempty set into one part is to put all of the elements in the same part. Unlike Stirling numbers of the first kind, they can be calculated using a one-sum formula: [2]

The Stirling numbers of the second kind may also be characterized as the numbers that arise when one expresses powers of an indeterminate x in terms of the falling factorials [3]

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

In other words

Notation

Various notations have been used for Stirling numbers of the second kind. The brace notation was used by Imanuel Marx and Antonio Salmeri in 1962 for variants of these numbers. [4] [5] This led Knuth to use it, as shown here, in the first volume of The Art of Computer Programming (1968). [6] [7] According to the third edition of The Art of Computer Programming, this notation was also used earlier by Jovan Karamata in 1935. [8] [9] The notation S(n, k) was used by Richard Stanley in his book Enumerative Combinatorics and also, much earlier, by many other writers. [6]

The notations used on this page for Stirling numbers are not universal, and may conflict with notations in other sources.

Relation to Bell numbers

Since the Stirling number counts set partitions of an n-element set into k parts, the sum

over all values of k is the total number of partitions of a set with n members. This number is known as the nth Bell number.

Analogously, the ordered Bell numbers can be computed from the Stirling numbers of the second kind via

[10]

Table of values

Below is a triangular array of values for the Stirling numbers of the second kind (sequence A008277 in the OEIS ):

k
n
012345678910
01
101
2011
30131
401761
5011525101
601319065151
70163301350140211
80112796617011050266281
9012553025777069512646462361
100151193303410542525228275880750451

As with the binomial coefficients, this table could be extended to k > n, but the entries would all be 0.

Properties

Recurrence relation

Stirling numbers of the second kind obey the recurrence relation

with initial conditions

For instance, the number 25 in column k = 3 and row n = 5 is given by 25 = 7 + (3×6), where 7 is the number above and to the left of 25, 6 is the number above 25 and 3 is the column containing the 6.

To prove this recurrence, observe that a partition of the objects into k nonempty subsets either contains the -th object as a singleton or it does not. The number of ways that the singleton is one of the subsets is given by

since we must partition the remaining n objects into the available subsets. In the other case the -th object belongs to a subset containing other objects. The number of ways is given by

since we partition all objects other than the -th into k subsets, and then we are left with k choices for inserting object . Summing these two values gives the desired result.

Another recurrence relation is given by

[ citation needed ]

Simple identities

Some simple identities include

This is because dividing n elements 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 pick those two elements;

and

To see this, first note that there are 2nordered pairs of complementary subsets A and B. In one case, A is empty, and in another B is empty, so 2n  2 ordered pairs of subsets remain. Finally, since we want unordered pairs rather than ordered pairs we divide this last number by 2, giving the result above.

Another explicit expansion of the recurrence-relation gives identities in the spirit of the above example.

Identities

The table in section 6.1 of Concrete Mathematics provides a plethora of generalized forms of finite sums involving the Stirling numbers. Several particular finite sums relevant to this article include.

Explicit formula

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

This can be derived by using inclusion-exclusion to count the number of surjections from n to k and using the fact that the number of such surjections is .

Additionally, this formula is a special case of the kth forward difference of the monomial evaluated at x = 0:

Because the Bernoulli polynomials may be written in terms of these forward differences, one immediately obtains a relation in the Bernoulli numbers:

The evaluation of incomplete exponential Bell polynomial Bn,k(x1,x2,...) on the sequence of ones equals a Stirling number of the second kind:

Another explicit formula given in the NIST Handbook of Mathematical Functions is

Parity

Parity of Stirling numbers of the second kind. Stirling numbers of the second kind - Parity.svg
Parity of Stirling numbers of the second kind.

The parity of a Stirling number of the second kind is equal to the parity of a related binomial coefficient:

where

This relation is specified by mapping n and k coordinates onto the Sierpiński triangle.

More directly, let two sets contain positions of 1's in binary representations of results of respective expressions:

One can mimic a bitwise AND operation by intersecting these two sets:

to obtain the parity of a Stirling number of the second kind in O(1) time. In pseudocode:

where is the Iverson bracket.

The parity of a central Stirling number of the second kind is odd if and only if is a fibbinary number, a number whose binary representation has no two consecutive 1s. [11]

Generating functions

For a fixed integer n, the ordinary generating function for Stirling numbers of the second kind is given by

where are Touchard polynomials. If one sums the Stirling numbers against the falling factorial instead, one can show the following identities, among others:

and

Which has special case

.

For a fixed integer k, the Stirling numbers of the second kind have rational ordinary generating function

and have an exponential generating function given by

A mixed bivariate generating function for the Stirling numbers of the second kind is

Lower and upper bounds

If and , then

[12]

Asymptotic approximation

For fixed value of the asymptotic value of the Stirling numbers of the second kind as is given by

If (where o denotes the little o notation) then

[13]

A uniformly valid approximation also exists: for all k such that 1 < k < n, one has

where , and is the unique solution to . [14] Relative error is bounded by about .

Maximum for fixed n

For fixed , has a single maximum, which is attained for at most two consecutive values of k. That is, there is an integer such that

Looking at the table of values above, the first few values for are

When is large

and the maximum value of the Stirling number can be approximated with

[12]

Applications

Moments of the Poisson distribution

If X is a random variable with a Poisson distribution with expected value λ, then its n-th 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 Dobiński's formula).

Moments of fixed points of random permutations

Let the random variable X be the number of fixed points of a uniformly distributed random permutation of a finite set of size m. Then the nth moment of X is

Note: The upper bound of summation is m, not n.

In other words, the nth moment of this probability distribution is the number of partitions of a set of size n into no more than m parts. This is proved in the article on random permutation statistics, although the notation is a bit different.

Rhyming schemes

The Stirling numbers of the second kind can represent the total number of rhyme schemes for a poem of n lines. gives the number of possible rhyming schemes for n lines using k unique rhyming syllables. As an example, for a poem of 3 lines, there is 1 rhyme scheme using just one rhyme (aaa), 3 rhyme schemes using two rhymes (aab, aba, abb), and 1 rhyme scheme using three rhymes (abc).

Variants

r-Stirling numbers of the second kind

The r-Stirling number of the second kind counts the number of partitions of a set of n objects into k non-empty disjoint subsets, such that the first r elements are in distinct subsets. [15] These numbers satisfy the recurrence relation

Some combinatorial identities and a connection between these numbers and context-free grammars can be found in [16]

Associated Stirling numbers of the second kind

An r-associated Stirling number of the second kind is the number of ways to partition a set of n objects into k subsets, with each subset containing at least r elements. [17] It is denoted by and obeys the recurrence relation

The 2-associated numbers (sequence A008299 in the OEIS ) appear elsewhere as "Ward numbers" and as the magnitudes of the coefficients of Mahler polynomials.

Reduced Stirling numbers of the second kind

Denote the n objects to partition by the integers 1, 2, ..., n. Define the reduced Stirling numbers of the second kind, denoted , to be the number of ways to partition the integers 1, 2, ..., n into k nonempty subsets such that all elements in each subset have pairwise distance at least d. That is, for any integers i and j in a given subset, it is required that . It has been shown that these numbers satisfy

(hence the name "reduced"). [18] Observe (both by definition and by the reduction formula), that , the familiar Stirling numbers of the second kind.

See also

Related Research Articles

<span class="mw-page-title-main">Binomial coefficient</span> Number of subsets of a given size

In mathematics, the binomial coefficients are the positive integers that occur as coefficients in the binomial theorem. Commonly, a binomial coefficient is indexed by a pair of integers nk ≥ 0 and is written It is the coefficient of the xk term in the polynomial expansion of the binomial power (1 + x)n; this coefficient can be computed by the multiplicative formula

In mathematics, the Bernoulli numbersBn are a sequence of rational numbers which occur frequently in analysis. The Bernoulli numbers appear in the Taylor series expansions of the tangent and hyperbolic tangent functions, in Faulhaber's formula for the sum of m-th powers of the first n positive integers, in the Euler–Maclaurin formula, and in expressions for certain values of the Riemann zeta function.

In mathematics, the Euler numbers are a sequence En of integers defined by the Taylor series expansion

In mathematics, a power series is an infinite series of the form

<span class="mw-page-title-main">Lah number</span> Mathematical sequence

In mathematics, the (signed and unsigned) Lah numbers are coefficients expressing rising factorials in terms of falling factorials and vice versa. They were discovered by Ivo Lah in 1954. Explicitly, the unsigned Lah numbers are given by the formula involving the binomial coefficient

In mathematics, Stirling numbers arise in a variety of analytic and combinatorial problems. They are named after James Stirling, who introduced them in a purely algebraic setting in his book Methodus differentialis (1730). They were rediscovered and given a combinatorial meaning by Masanobu Saka in 1782.

In combinatorial mathematics, the Bell numbers count the possible partitions of a set. These numbers have been studied by mathematicians since the 19th century, and their roots go back to medieval Japan. In an example of Stigler's law of eponymy, they are named after Eric Temple Bell, who wrote about them in the 1930s.

In mathematics, the falling factorial is defined as the polynomial

<span class="mw-page-title-main">Partition of a set</span> Mathematical ways to group elements of a set

In mathematics, a partition of a set is a grouping of its elements into non-empty subsets, in such a way that every element is included in exactly one subset.

The Touchard polynomials, studied by Jacques Touchard, also called the exponential polynomials or Bell polynomials, comprise a polynomial sequence of binomial type defined by

<span class="mw-page-title-main">Inclusion–exclusion principle</span> Counting technique in combinatorics

In combinatorics, a branch of mathematics, the inclusion–exclusion principle is a counting technique which generalizes the familiar method of obtaining the number of elements in the union of two finite sets; symbolically expressed as

In mathematics, especially in combinatorics, Stirling numbers of the first kind arise in the study of permutations. In particular, the Stirling numbers of the first kind count permutations according to their number of cycles.

Euclid's theorem is a fundamental statement in number theory that asserts that there are infinitely many prime numbers. It was first proven by Euclid in his work Elements. There are several proofs of the theorem.

In mathematics, poly-Bernoulli numbers, denoted as , were defined by M. Kaneko as

In mathematics, the Stirling polynomials are a family of polynomials that generalize important sequences of numbers appearing in combinatorics and analysis, which are closely related to the Stirling numbers, the Bernoulli numbers, and the generalized Bernoulli polynomials. There are multiple variants of the Stirling polynomial sequence considered below most notably including the Sheffer sequence form of the sequence, , defined characteristically through the special form of its exponential generating function, and the Stirling (convolution) polynomials, , which also satisfy a characteristic ordinary generating function and that are of use in generalizing the Stirling numbers to arbitrary complex-valued inputs. We consider the "convolution polynomial" variant of this sequence and its properties second in the last subsection of the article. Still other variants of the Stirling polynomials are studied in the supplementary links to the articles given in the references.

<span class="mw-page-title-main">Rand index</span> Measure of similarity between two data clusterings

The Rand index or Rand measure in statistics, and in particular in data clustering, is a measure of the similarity between two data clusterings. A form of the Rand index may be defined that is adjusted for the chance grouping of elements, this is the adjusted Rand index. The Rand index is the accuracy of determining if a link belongs within a cluster or not.

<span class="mw-page-title-main">Eulerian number</span> Polynomial sequence

In combinatorics, the Eulerian number is the number of permutations of the numbers 1 to in which exactly elements are greater than the previous element.

<span class="mw-page-title-main">Ordered Bell number</span> Number of weak orderings

In number theory and enumerative combinatorics, the ordered Bell numbers or Fubini numbers count the number of weak orderings on a set of elements. Weak orderings arrange their elements into a sequence allowing ties, such as might arise as the outcome of a horse race). Starting from , these numbers are

The use of exponential generating functions (EGFs) to study the properties of Stirling numbers is a classical exercise in combinatorial mathematics and possibly the canonical example of how symbolic combinatorics is used. It also illustrates the parallels in the construction of these two types of numbers, lending support to the binomial-style notation that is used for them.

The partition function or configuration integral, as used in probability theory, information theory and dynamical systems, is a generalization of the definition of a partition function in statistical mechanics. It is a special case of a normalizing constant in probability theory, for the Boltzmann distribution. The partition function occurs in many problems of probability theory because, in situations where there is a natural symmetry, its associated probability measure, the Gibbs measure, has the Markov property. This means that the partition function occurs not only in physical systems with translation symmetry, but also in such varied settings as neural networks, and applications such as genomics, corpus linguistics and artificial intelligence, which employ Markov networks, and Markov logic networks. The Gibbs measure is also the unique measure that has the property of maximizing the entropy for a fixed expectation value of the energy; this underlies the appearance of the partition function in maximum entropy methods and the algorithms derived therefrom.

References

  1. Ronald L. Graham, Donald E. Knuth, Oren Patashnik (1988) Concrete Mathematics , Addison–Wesley, Reading MA. ISBN   0-201-14236-8, p. 244.
  2. "Stirling Numbers of the Second Kind, Theorem 3.4.1".
  3. Confusingly, the notation that combinatorialists use for falling factorials coincides with the notation used in special functions for rising factorials; see Pochhammer symbol.
  4. Transformation of Series by a Variant of Stirling's Numbers, Imanuel Marx, The American Mathematical Monthly69, #6 (June–July 1962), pp. 530–532, JSTOR   2311194.
  5. Antonio Salmeri, Introduzione alla teoria dei coefficienti fattoriali, Giornale di Matematiche di Battaglini 90 (1962), pp. 44–54.
  6. 1 2 Knuth, D.E. (1992), "Two notes on notation", Amer. Math. Monthly, 99 (5): 403–422, arXiv: math/9205211 , Bibcode:1992math......5211K, doi:10.2307/2325085, JSTOR   2325085, S2CID   119584305
  7. Donald E. Knuth, Fundamental Algorithms, Reading, Mass.: Addison–Wesley, 1968.
  8. p. 66, Donald E. Knuth, Fundamental Algorithms, 3rd ed., Reading, Mass.: Addison–Wesley, 1997.
  9. Jovan Karamata, Théorèmes sur la sommabilité exponentielle et d'autres sommabilités s'y rattachant, Mathematica (Cluj) 9 (1935), pp, 164–178.
  10. Sprugnoli, Renzo (1994), "Riordan arrays and combinatorial sums" (PDF), Discrete Mathematics, 132 (1–3): 267–290, doi: 10.1016/0012-365X(92)00570-H , MR   1297386
  11. Chan, O-Yeat; Manna, Dante (2010), "Congruences for Stirling numbers of the second kind" (PDF), Gems in Experimental Mathematics, Contemporary Mathematics, vol. 517, Providence, Rhode Island: American Mathematical Society, pp. 97–111, doi:10.1090/conm/517/10135, MR   2731094
  12. 1 2 Rennie, B.C.; Dobson, A.J. (1969). "On stirling numbers of the second kind". Journal of Combinatorial Theory . 7 (2): 116–121. doi: 10.1016/S0021-9800(69)80045-1 . ISSN   0021-9800.
  13. L. C. Hsu, Note on an Asymptotic Expansion of the nth Difference of Zero, AMS Vol.19 NO.2 1948, pp. 273--277
  14. N. M. Temme, Asymptotic Estimates of Stirling Numbers, STUDIES IN APPLIED MATHEMATICS 89:233-243 (1993), Elsevier Science Publishing.
  15. Broder, A. (1984). The r-Stirling numbers. Discrete Mathematics 49, 241-259
  16. Triana, J. (2022). r-Stirling numbers of the second kind through context-free grammars. Journal of automata, languages and combinatorics 27(4), 323-333
  17. L. Comtet, Advanced Combinatorics, Reidel, 1974, p. 222.
  18. A. Mohr and T.D. Porter, Applications of Chromatic Polynomials Involving Stirling Numbers , Journal of Combinatorial Mathematics and Combinatorial Computing 70 (2009), 57–64.