Stirling numbers of the first kind

Last updated

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 (counting fixed points as cycles of length one).

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 first kind. Identities linking the two kinds appear in the article on Stirling numbers.

Definitions

Stirling numbers of the first kind are the coefficients in the expansion of the falling factorial

into powers of the variable :

For example, , leading to the values , , and .

Subsequently, it was discovered that the absolute values of these numbers are equal to the number of permutations of certain kinds. These absolute values, which are known as unsigned Stirling numbers of the first kind, are often denoted or . They may be defined directly to be the number of permutations of elements with disjoint cycles. For example, of the permutations of three elements, there is one permutation with three cycles (the identity permutation, given in one-line notation by or in cycle notation by ), three permutations with two cycles (, , and ) and two permutations with one cycle ( and ). Thus, , and . These can be seen to agree with the previous calculation of for . It was observed by Alfréd Rényi that the unsigned Stirling number also count the number of permutations of size with left-to-right maxima. [1]

The unsigned Stirling numbers may also be defined algebraically, as the coefficients of the rising factorial:

.

The notations used on this page for Stirling numbers are not universal, and may conflict with notations in other sources. (The square bracket notation is also common notation for the Gaussian coefficients.)

Definition by permutation

can be defined as the number of permutations on elements with cycles.

s(4,2)=11 Stirling number of the first kind s(4,2).svg
s(4,2)=11

The image at right shows that : the symmetric group on 4 objects has 3 permutations of the form

(having 2 orbits, each of size 2),

and 8 permutations of the form

(having 1 orbit of size 3 and 1 orbit of size 1).

These numbers can be calculated by considering the orbits as conjugancy classes (last bullet point).

Signs

The signs of the (signed) Stirling numbers of the first kind are predictable and depend on the parity of nk. In particular,

Recurrence relation

The unsigned Stirling numbers of the first kind can be calculated by the recurrence relation

for , with the initial conditions

for .

It follows immediately that the (signed) Stirling numbers of the first kind satisfy the recurrence

.
Algebraic proof

We prove the recurrence relation using the definition of Stirling numbers in terms of rising factorials. Distributing the last term of the product, we have

The coefficient of on the left-hand side of this equation is . The coefficient of in is , while the coefficient of in is . Since the two sides are equal as polynomials, the coefficients of on both sides must be equal, and the result follows.

Combinatorial proof

We prove the recurrence relation using the definition of Stirling numbers in terms of permutations with a given number of cycles (or equivalently, orbits).

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 increases the number of cycles by 1 and so 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 objects 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 ways to perform this insertion, inserting the new object immediately following any of the already present. This explains the term of the recurrence relation. These two cases include all possibilities, so the recurrence relation follows.

Table of values

Below is a triangular array of unsigned values for the Stirling numbers of the first kind, similar in form to Pascal's triangle. These values are easy to generate using the recurrence relation in the previous section.

k
n
012345678910
01
101
2011
30231
4061161
50245035101
6012027422585151
7072017641624735175211
805040130681313267691960322281
904032010958411812467284224494536546361
10036288010265761172700723680269325632739450870451

Properties

Simple identities

Using the Kronecker delta one have,

and

if , or more generally if k > n.

Also

and

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. Generalizations of the Stirling numbers of both kinds to arbitrary complex-valued inputs may be extended through the relations of these triangles to the Stirling convolution polynomials. [2]

Combinatorial proofs

These identities may be derived by enumerating permutations directly. For example, a permutation of n elements with n  3 cycles must have one of the following forms:

  • n  6 fixed points and three two-cycles
  • n  5 fixed points, a three-cycle and a two-cycle, or
  • n  4 fixed points and a four-cycle.

The three types may be enumerated as follows:

  • choose the six elements that go into the two-cycles, decompose them into two-cycles and take into account that the order of the cycles is not important:
  • choose the five elements that go into the three-cycle and the two-cycle, choose the elements of the three-cycle and take into account that three elements generate two three-cycles:
  • choose the four elements of the four-cycle and take into account that four elements generate six four-cycles:

Sum the three contributions to obtain

Note that all the combinatorial proofs above use either binomials or multinomials of .

Therefore if is prime, then:

for .

Expansions for fixed k

Since the Stirling numbers are the coefficients of a polynomial with roots 0, 1, ..., n − 1, one has by Vieta's formulas that

In other words, the Stirling numbers of the first kind are given by elementary symmetric polynomials evaluated at 0, 1, ..., n − 1. [3] In this form, the simple identities given above take the form

and so on.

One may produce alternative forms for the Stirling numbers of the first kind with a similar approach preceded by some algebraic manipulation: since

it follows from Newton's formulas that one can expand the Stirling numbers of the first kind in terms of generalized harmonic numbers. This yields identities like

where Hn is the harmonic number and Hn(m) is the generalized harmonic number

These relations can be generalized to give

where w(n, m) is defined recursively in terms of the generalized harmonic numbers by

(Here δ is the Kronecker delta function and is the Pochhammer symbol.) [4]

For fixed these weighted harmonic number expansions are generated by the generating function

where the notation means extraction of the coefficient of from the following formal power series (see the non-exponential Bell polynomials and section 3 of [5] ).

More generally, sums related to these weighted harmonic number expansions of the Stirling numbers of the first kind can be defined through generalized zeta series transforms of generating functions. [6] [7]

One can also "invert" the relations for these Stirling numbers given in terms of the -order harmonic numbers to write the integer-order generalized harmonic numbers in terms of weighted sums of terms involving the Stirling numbers of the first kind. For example, when the second-order and third-order harmonic numbers are given by

More generally, one can invert the Bell polynomial generating function for the Stirling numbers expanded in terms of the -order harmonic numbers to obtain that for integers

Finite sums

Since permutations are partitioned by number of cycles, one has

The identity

can be proved by the techniques on the page Stirling numbers and exponential generating functions.

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

Additionally, if we define the second-order Eulerian numbers by the triangular recurrence relation [8]

we arrive at the following identity related to the form of the Stirling convolution polynomials which can be employed to generalize both Stirling number triangles to arbitrary real, or complex-valued, values of the input :

Particular expansions of the previous identity lead to the following identities expanding the Stirling numbers of the first kind for the first few small values of :

Software tools for working with finite sums involving Stirling numbers and Eulerian numbers are provided by the RISC Stirling.m package utilities in Mathematica. Other software packages for guessing formulas for sequences (and polynomial sequence sums) involving Stirling numbers and other special triangles is available for both Mathematica and Sage here and here, respectively. [9]


Congruences

The following congruence identity may be proved via a generating function-based approach: [10]

More recent results providing Jacobi-type J-fractions that generate the single factorial function and generalized factorial-related products lead to other new congruence results for the Stirling numbers of the first kind. [11] For example, working modulo we can prove that

Where is the Iverson bracket.

and working modulo we can similarly prove that

More generally, for fixed integers if we define the ordered roots

then we may expand congruences for these Stirling numbers defined as the coefficients

in the following form where the functions, , denote fixed polynomials of degree in for each , , and :

Section 6.2 of the reference cited above provides more explicit expansions related to these congruences for the -order harmonic numbers and for the generalized factorial products, .

Generating functions

A variety of identities may be derived by manipulating the generating function (see change of basis):

Using the equality

it follows that

and

(This identity is valid for formal power series, and the sum converges in the complex plane for |z| < 1.) Other identities arise by exchanging the order of summation, taking derivatives, making substitutions for z or u, etc. For example, we may derive: [12]

or

and

where and are the Riemann zeta function and the Hurwitz zeta function respectively, and even evaluate this integral

where is the gamma function. There also exist more complicated expressions for the zeta-functions involving the Stirling numbers. One, for example, has

This series generalizes Hasse's series for the Hurwitz zeta-function (we obtain Hasse's series by setting k=1). [13] [14]

Asymptotics

The next estimate given in terms of the Euler gamma constant applies: [15]

For fixed we have the following estimate :

Explicit formula

It is well-known that we don't know any one-sum formula for Stirling numbers of the first kind. A two-sum formula can be obtained using one of the symmetric formulae for Stirling numbers in conjunction with the explicit formula for Stirling numbers of the second kind.

As discussed earlier, by Vieta's formulas, one get

The Stirling number s(n,n-p) can be found from the formula [16]

where The sum is a sum over all partitions of p.

Another exact nested sum expansion for these Stirling numbers is computed by elementary symmetric polynomials corresponding to the coefficients in of a product of the form . In particular, we see that

Newton's identities combined with the above expansions may be used to give an alternate proof of the weighted expansions involving the generalized harmonic numbers already noted above.

Relations to natural logarithm function

The nth derivative of the μth power of the natural logarithm involves the signed Stirling numbers of the first kind:

where is the falling factorial, and is the signed Stirling number.

It can be proved by using mathematical induction.

Other formulas

Stirling numbers of the first kind appear in the formula for Gregory coefficients and in a finite sum identity involving Bell number [17]

Infinite series involving the finite sums with the Stirling numbers often lead to the special functions. For example [12] [18]

and

or even

where γm are the Stieltjes constants and δm,0 represents the Kronecker delta function.

Notice that this last identity immediately implies relations between the polylogarithm functions, the Stirling number exponential generating functions given above, and the Stirling-number-based power series for the generalized Nielsen polylogarithm functions.

Generalizations

There are many notions of generalized Stirling numbers that may be defined (depending on application) in a number of differing combinatorial contexts. In so much as the Stirling numbers of the first kind correspond to the coefficients of the distinct polynomial expansions of the single factorial function, , we may extend this notion to define triangular recurrence relations for more general classes of products.

In particular, for any fixed arithmetic function and symbolic parameters , related generalized factorial products of the form

may be studied from the point of view of the classes of generalized Stirling numbers of the first kind defined by the following coefficients of the powers of in the expansions of and then by the next corresponding triangular recurrence relation:

These coefficients satisfy a number of analogous properties to those for the Stirling numbers of the first kind as well as recurrence relations and functional equations related to the f-harmonic numbers, . [19]

One special case of these bracketed coefficients corresponding to allows us to expand the multiple factorial, or multifactorial functions as polynomials in (see generalizations of the double factorial). [20]

The Stirling numbers of both kinds, the binomial coefficients, and the first and second-order Eulerian numbers are all defined by special cases of a triangular super-recurrence of the form

for integers and where whenever or . In this sense, the form of the Stirling numbers of the first kind may also be generalized by this parameterized super-recurrence for fixed scalars (not all zero).

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, 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 mathematics, a generating function is a representation of an infinite sequence of numbers as the coefficients of a formal power series. Unlike an ordinary series, the formal power series is not required to converge: in fact, the generating function is not actually regarded as a function, and the "variable" remains an indeterminate. Generating functions were first introduced by Abraham de Moivre in 1730, in order to solve the general linear recurrence problem. One can generalize to formal power series in more than one indeterminate, to encode information about infinite multi-dimensional arrays of numbers.

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

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

<span class="mw-page-title-main">Double factorial</span> Mathematical function

In mathematics, the double factorial of a number n, denoted by n, is the product of all the positive integers up to n that have the same parity as n. That is,

In combinatorial mathematics, the Bell polynomials, named in honor of Eric Temple Bell, are used in the study of set partitions. They are related to Stirling and Bell numbers. They also occur in many applications, such as in Faà di Bruno's formula.

<span class="mw-page-title-main">Lambert series</span> Mathematical term

In mathematics, a Lambert series, named for Johann Heinrich Lambert, is a series taking the form

<span class="mw-page-title-main">Stirling numbers of the second kind</span> Numbers parameterizing ways to partition a set

In mathematics, particularly in combinatorics, a Stirling number of the second kind is the number of ways to partition a set of n objects into k non-empty subsets and is denoted by or . 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.

In combinatorics, the binomial transform is a sequence transformation that computes its forward differences. It is closely related to the Euler transform, which is the result of applying the binomial transform to the sequence associated with its ordinary generating function.

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.

The statistics of random permutations, such as the cycle structure of a random permutation are of fundamental importance in the analysis of algorithms, especially of sorting algorithms, which operate on random permutations. Suppose, for example, that we are using quickselect to select a random element of a random permutation. Quickselect will perform a partial sort on the array, as it partitions the array according to the pivot. Hence a permutation will be less disordered after quickselect has been performed. The amount of disorder that remains may be analysed with generating functions. These generating functions depend in a fundamental way on the generating functions of random permutation statistics. Hence it is of vital importance to compute these generating functions.

<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.

In mathematics, Faulhaber's formula, named after the early 17th century mathematician Johann Faulhaber, expresses the sum of the p-th powers of the first n positive integers

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.

In mathematics, a transformation of a sequence's generating function provides a method of converting the generating function for one sequence into a generating function enumerating another. These transformations typically involve integral formulas applied to a sequence generating function or weighted sums over the higher-order derivatives of these functions.

References

  1. Rényi, Alfred (1962). "Théorie des éléments saillants d'une suite d'observations". Annales scientifiques de l'Université de Clermont. Mathématiques. Tome 8 (2): 7–13.
  2. See section 6.2 and 6.5 of Concrete Mathematics.
  3. Richard P. Stanley, Enumerative Combinatorics, volume 1 (2nd ed.). Page 34 of the online version.
  4. Adamchik, V. (1996). "On Stirling numbers and Euler sums" (PDF).{{cite journal}}: Cite journal requires |journal= (help)
  5. Flajolet and Sedgewick (1995). "Mellin transforms and asymptotics: Finite differences and Rice's integrals" (PDF). Theoretical Computer Science. 144 (1–2): 101–124. doi:10.1016/0304-3975(94)00281-m.
  6. Schmidt, M. D. (30 October 2016). "Zeta Series Generating Function Transformations Related to Polylogarithm Functions and the k-Order Harmonic Numbers". arXiv: 1610.09666 [math.CO].
  7. Schmidt, M. D. (3 November 2016). "Zeta Series Generating Function Transformations Related to Generalized Stirling Numbers and Partial Sums of the Hurwitz Zeta Function". arXiv: 1611.00957 [math.CO].
  8. A table of the second-order Eulerian numbers and a synopsis of their properties is found in section 6.2 of Concrete Mathematics. For example, we have that . These numbers also have the following combinatorial interpretation: If we form all permutations of the multiset with the property that all numbers between the two occurrences of are greater than for , then is the number of such permutations that have ascents.
  9. Schmidt, M. D. (2016). "A Computer Algebra Package for Polynomial Sequence Recognition". arXiv: 1609.07301 [math.CO].
  10. Herbert Wilf, Generatingfunctionology, Section 4.6.
  11. Schmidt, M. D. (2017). "Jacobi-Type Continued Fractions for the Ordinary Generating Functions of Generalized Factorial Functions". J. Integer Seq. 20 (3). arXiv: 1610.09691 .
  12. 1 2 Ia. V. Blagouchine (2016). "Two series expansions for the logarithm of the gamma function involving Stirling numbers and containing only rational coefficients for certain arguments related to π1". Journal of Mathematical Analysis and Applications. 442 (2): 404–434. arXiv: 1408.3902 . doi:10.1016/j.jmaa.2016.04.032. S2CID   119661147. arXiv
  13. Blagouchine, Iaroslav V. (2018). "Three Notes on Ser's and Hasse's Representations for the Zeta-functions". INTEGERS: The Electronic Journal of Combinatorial Number Theory. 18A: 1–45. arXiv: 1606.02044 . Bibcode:2016arXiv160602044B.
  14. See also some more interesting series representations and expansions mentioned in Connon's article: Connon, D. F. (2007). "Some series and integrals involving the Riemann zeta function, binomial coefficients and the harmonic numbers (Volume I)". arXiv: 0710.4022 [math.HO]..
  15. These estimates are found in Section 26.8 of the NIST Handbook of Mathematical Functions.
  16. J. Malenfant, "Finite, Closed-form Expressions for the Partition Function and for Euler, Bernoulli, and Stirling Numbers"
  17. Komatsu, Takao; Pita-Ruiz, Claudio (2018). "Some formulas for Bell numbers". Filomat. 32 (11): 3881–3889. doi: 10.2298/FIL1811881K . ISSN   0354-5180.
  18. Ia. V. Blagouchine (2016). "Expansions of generalized Euler's constants into the series of polynomials in π2 and into the formal enveloping series with rational coefficients only". Journal of Number Theory. 158 (2): 365–396. doi:10.1016/j.jnt.2015.06.012. arXiv
  19. Combinatorial Identities for Generalized Stirling Numbers Expanding f-Factorial Functions and the f-Harmonic Numbers (2016).
  20. Schmidt, Maxie D. (2010). "Generalized j-Factorial Functions, Polynomials, and Applications". J. Integer Seq. 13.