Jump to content

Stirling number: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Kkuhts (talk | contribs)
 
(33 intermediate revisions by 11 users not shown)
Line 1: Line 1:
{{Short description|Important sequences in combinatorics}}
{{Use American English|date = March 2019}}
{{Use American English|date = March 2019}}
{{Short description|Important sequences in combinatorics}}
In [[mathematics]], '''Stirling numbers''' arise in a variety of [[Analysis (mathematics)|analytic]] and [[combinatorics|combinatorial]] problems. They are named after [[James Stirling (mathematician)|James Stirling]], who introduced them in a purely algebraic setting in his book ''Methodus differentialis'' (1730).{{sfn|Mansour|Schork|2015|p=5}} They were rediscovered and given a combinatorial meaning by Masanobu Saka in 1782.{{sfn|Mansour|Schork|2015|p=4}}
In [[mathematics]], '''Stirling numbers''' arise in a variety of [[Analysis (mathematics)|analytic]] and [[combinatorics|combinatorial]] problems. They are named after [[James Stirling (mathematician)|James Stirling]], who introduced them in a purely algebraic setting in his book ''Methodus differentialis'' (1730).{{sfn|Mansour|Schork|2015|p=5}} They were rediscovered and given a combinatorial meaning by Masanobu Saka in 1782.{{sfn|Mansour|Schork|2015|p=4}}


Two different sets of numbers bear this name: the [[Stirling numbers of the first kind]] and the [[Stirling numbers of the second kind]]. Additionally, [[Lah numbers]] are sometimes referred to as Stirling numbers of the third kind. Each kind is detailed in its respective article, this one serving as a description of relations between them.
Two different sets of numbers bear this name: the [[Stirling numbers of the first kind]] and the [[Stirling numbers of the second kind]]. Additionally, [[Lah numbers]] are sometimes referred to as Stirling numbers of the third kind. Each kind is detailed in its respective article, this one serving as a description of relations between them.


A common property of all three kinds is that they describe coefficients relating three different sequences of polynomials that frequently arise in combinatorics. Moreover, all three can be defined as the number of partitions of ''n'' elements into ''k'' non-empty subsets, with different ways of counting orderings within each subset.
A common property of all three kinds is that they describe coefficients relating three different sequences of polynomials that frequently arise in combinatorics. Moreover, all three can be defined as the number of partitions of ''n'' elements into ''k'' non-empty subsets, where each subset is endowed with a certain kind of order (no order, cyclical, or linear).


==Notation==
==Notation==
{{main|Stirling numbers of the first kind|Stirling numbers of the second kind}}
{{main|Stirling numbers of the first kind|Stirling numbers of the second kind}}
Several different notations for Stirling numbers are in use. Common notation for ordinary (signed) '''Stirling numbers of the first kind''' is:
Several different notations for Stirling numbers are in use. Ordinary (signed) '''Stirling numbers of the first kind''' are commonly denoted:


: <math> s(n,k)\,</math>
: <math> s(n,k)\,.</math>
For '''unsigned Stirling numbers of the first kind''', which count the number of [[permutation]]s of ''n'' elements with ''k'' disjoint [[cyclic permutation|cycle]]s, is:
'''Unsigned Stirling numbers of the first kind''', which count the number of [[permutation]]s of ''n'' elements with ''k'' disjoint [[cyclic permutation|cycle]]s, are denoted:


: <math> \biggl[{n \atop k}\biggr] =c(n,k)=|s(n,k)|=(-1)^{n-k} s(n,k)\,</math>
: <math> \biggl[{n \atop k}\biggr] =c(n,k)=|s(n,k)|=(-1)^{n-k} s(n,k)\,</math>
And for '''Stirling numbers of the second kind''', which count the number of ways to partition a set of ''n'' elements into ''k'' nonempty subsets:<ref>Ronald L. Graham, Donald E. Knuth, Oren Patashnik (1988) ''[[Concrete Mathematics]]'', Addison-Wesley, Reading MA. {{isbn|0-201-14236-8}}, p.&nbsp;244.</ref>
'''Stirling numbers of the second kind''', which count the number of ways to partition a set of ''n'' elements into ''k'' nonempty subsets:<ref>Ronald L. Graham, Donald E. Knuth, Oren Patashnik (1988) ''[[Concrete Mathematics]]'', Addison-Wesley, Reading MA. {{isbn|0-201-14236-8}}, p.&nbsp;244.</ref>


: <math> \biggl\{{\!n\! \atop \!k\!}\biggr\} = S(n,k) = S_n^{(k)} \,</math>
: <math> \biggl\{{\!n\! \atop \!k\!}\biggr\} = S(n,k) = S_n^{(k)} \,</math>

For example, the sum <math display=inline>\displaystyle\sum_{k=0}^n \left[{n \atop k}\right] = n!</math> is the number of all permutations, while the sum <math display=inline>\displaystyle\sum_{k=0}^n \left\{{\!n\! \atop \!k\!}\right\} = B_n</math> is the ''n''th [[Bell numbers|Bell number]].


[[Abramowitz and Stegun]] use an uppercase <math>S</math> and a [[blackletter]] <math>\mathfrak S</math>, respectively, for the first and second kinds of Stirling number. The notation of brackets and braces, in analogy to [[binomial coefficients]], was introduced in 1935 by [[Jovan Karamata]] and promoted later by [[Donald Knuth]]. (The bracket notation conflicts with a common notation for [[Gaussian coefficient]]s.<ref>[[Donald Knuth]]</ref>) The mathematical motivation for this type of notation, as well as additional Stirling number formulae, may be found on the page for [[Stirling numbers and exponential generating functions]].
[[Abramowitz and Stegun]] use an uppercase <math>S</math> and a [[blackletter]] <math>\mathfrak S</math>, respectively, for the first and second kinds of Stirling number. The notation of brackets and braces, in analogy to [[binomial coefficients]], was introduced in 1935 by [[Jovan Karamata]] and promoted later by [[Donald Knuth]]. (The bracket notation conflicts with a common notation for [[Gaussian coefficient]]s.<ref>[[Donald Knuth]]</ref>) The mathematical motivation for this type of notation, as well as additional Stirling number formulae, may be found on the page for [[Stirling numbers and exponential generating functions]].


In later years, the more reasonable <math>s_1(n,k)</math> and <math>s_2(n,k)</math> were used.
Another infrequent notation is <math>s_1(n,k)</math> and <math>s_2(n,k)</math>.


==Expansions of falling and rising factorials==
==Expansions of falling and rising factorials<span class="anchor" id="falling_and_rising_anchor"></span>==
Stirling numbers express coefficients in expansions of [[falling and rising factorials]] (also known as the Pochhammer symbol) as polynomials.
Stirling numbers express coefficients in expansions of [[falling and rising factorials]] (also known as the Pochhammer symbol) as polynomials.


That is, the '''falling factorial''', defined as <math>(x)_{n} = x(x-1)\cdots(x-n+1)</math>, is a polynomial in ''x'' of degree ''n'' whose expansion is
That is, the '''falling factorial''', defined as <math>\ (x)_{n} = x(x-1)\ \cdots(x-n+1)\ ,</math> is a polynomial in {{mvar|x}} of degree {{mvar|n}} whose expansion is


:<math>(x)_{n} = \sum_{k=0}^n s(n,k) x^k</math>
:<math>(x)_{n}\ =\ \sum_{k=0}^n\ s(n,k)\ x^k\ </math>


with (signed) Stirling numbers of the first kind as coefficients.
with (signed) Stirling numbers of the first kind as coefficients.


Note that (''x'')<sub>0</sub> = 1 because it is an [[empty product]]. [[Combinatorics|Combinatorialists]] also sometimes use the notation <math style="vertical-align:baseline;">x^{\underline{n}}</math> for the falling factorial, and <math style="vertical-align:baseline;">x^{\overline{n}}</math> for the rising factorial.<ref>{{cite book|last=Aigner|first=Martin|title=A Course In Enumeration|url=https://archive.org/details/courseenumeratio00aign_007|url-access=limited|publisher=Springer|year=2007|pages=[https://archive.org/details/courseenumeratio00aign_007/page/n563 561]|chapter=Section 1.2 - Subsets and Binomial Coefficients|isbn=978-3-540-39032-9}}</ref> (Confusingly, the Pochhammer symbol that many use for ''falling'' factorials is used in [[special function]]s for ''rising'' factorials.)
Note that <math>\ (x)_0 \equiv 1\ ,</math> by convention, because it is an [[empty product]]. The notations <math style="vertical-align:baseline;">\ x^{\underline{n}}\ </math> for the falling factorial and <math style="vertical-align:baseline;">\ x^{\overline{n}}\ </math> for the rising factorial are also often used.<ref>{{cite book |last=Aigner |first=Martin |title=A Course in Enumeration |url=https://archive.org/details/courseenumeratio00aign_007 |url-access=limited |publisher=Springer |year=2007 |pages=[https://archive.org/details/courseenumeratio00aign_007/page/n563 561] |chapter=Section 1.2 - Subsets and binomial coefficients |isbn=978-3-540-39032-9}}</ref> (Confusingly, the Pochhammer symbol that many use for ''falling'' factorials is used in [[special function]]s for ''rising'' factorials.)


Similarly, the '''rising factorial''', defined as <math>x^{(n)} = x(x+1)\cdots(x+n-1)</math>, is a polynomial in ''x'' of degree ''n'' whose expansion is
Similarly, the '''rising factorial''', defined as <math>\ x^{(n)}\ =\ x(x+1)\ \cdots(x+n-1)\ ,</math> is a polynomial in {{mvar|x}} of degree {{mvar|n}} whose expansion is


:<math>x^{(n)} = \sum_{k=0}^n \biggl[{n \atop k}\biggr] x^k = \sum_{k=0}^n (-1)^{n-k} s(n,k) x^k</math>
:<math> x^{(n)}\ =\ \sum_{k=0}^n\ \biggl[{n \atop k}\biggr]\ x^k\ =\ \sum_{k=0}^n\ (-1)^{n-k}\ s(n,k)\ x^k\ ,</math>


with unsigned Stirling numbers of the first kind as coefficients.
with unsigned Stirling numbers of the first kind as coefficients.
One of these expansions can be derived from the other by observing that <math>x^{(n)} = (-1)^n (-x)_{n}</math>.
One of these expansions can be derived from the other by observing that <math>\ x^{(n)} = (-1)^n (-x)_{n} ~.</math>


Stirling numbers of the second kind express the reverse relations:
Stirling numbers of the second kind express the reverse relations:


:<math>x^n = \sum_{k=0}^n S(n,k) (x)_k</math>
:<math>\ x^n\ =\ \sum_{k=0}^n\ S(n,k)\ (x)_k\ </math>


and
and


:<math>x^n = \sum_{k=0}^n (-1)^{n-k} S(n,k) x^{(k)}.</math>
:<math>\ x^n\ =\ \sum_{k=0}^n\ (-1)^{n-k}\ S(n,k)\ x^{(k)} ~.</math>


==As change of basis coefficients==
==As change of basis coefficients==
Line 58: Line 56:
That is, every polynomial in ''x'' can be written as a sum <math>a_0 x^{(0)} + a_1 x^{(1)} + \dots + a_n x^{(n)}</math> for some unique coefficients <math>a_i</math> (similarly for the other two bases).
That is, every polynomial in ''x'' can be written as a sum <math>a_0 x^{(0)} + a_1 x^{(1)} + \dots + a_n x^{(n)}</math> for some unique coefficients <math>a_i</math> (similarly for the other two bases).
The above relations then express the [[change of basis]] between them, as summarized in the following [[commutative diagram]]:
The above relations then express the [[change of basis]] between them, as summarized in the following [[commutative diagram]]:
: [[File:Stirling numbers as polynomial basis change.svg|400px]]
: [[File:Stirling numbers as polynomial basis changes.png|center|500x500px|A diagram of how different [[Stirling number|Stirling numbers]] give coefficients for changing one basis of polynomials to another ]]
The coefficients for the two bottom changes are described by the [[#Lah numbers|Lah numbers]] below.
The coefficients for the two bottom changes are described by the [[#Lah numbers|Lah numbers]] below.
Since coefficients in any basis are unique, one can define Stirling numbers this way, as the coefficients expressing polynomials of one basis in terms of another, that is, the unique numbers relating <math>x^n</math> with falling and rising factorials as above.
Since coefficients in any basis are unique, one can define Stirling numbers this way, as the coefficients expressing polynomials of one basis in terms of another, that is, the unique numbers relating <math>x^n</math> with falling and rising factorials as above.
Line 67: Line 65:
===Example===
===Example===
Expressing a polynomial in the basis of falling factorials is useful for calculating sums of the polynomial evaluated at consecutive integers.
Expressing a polynomial in the basis of falling factorials is useful for calculating sums of the polynomial evaluated at consecutive integers.
Indeed, the sum of a falling factorial is simply expressed as another falling factorial (for ''k''≠−1)
Indeed, the sum of falling factorials with fixed ''k'' can expressed as another falling factorial (for <math>k\ne-1</math>)


:<math>\sum_{0\leq i < n} (i)_k = \frac{(n)_{k+1}}{k+1} </math>
:<math>\sum_{0\leq i < n} (i)_k = \frac{(n)_{k+1}}{k+1} </math>


This can be proved by [[mathematical induction|induction]].
This is analogous to the integral <math display="inline">\int_0^n x^kdx = n^{k+1} / (k+1)</math>, though the sum should be over integers ''i'' strictly less than ''n''.


For example, the sum of fourth powers of integers up to ''n'' (this time with ''n'' included), is:
For example, the sum of fourth powers of integers up to ''n'' (this time with ''n'' included), is:


:<math>\begin{align}
:<math>\begin{align}
&\sum_{i=0}^{n} i^4 = \sum_{i=0}^{n} \sum_{k=0}^4 \biggl\{{\!4\! \atop \!k\!}\biggr\} (i)_k = \sum_{k=0}^4 \biggl\{{\!n\! \atop \!k\!}\biggr\} \frac{(n+1)_{k+1}}{k+1} \\[8mu]
\sum_{i=0}^{n} i^4 & = \sum_{i=0}^n \sum_{k=0}^4 \biggl\{{\!4\! \atop \!k\!}\biggr\} (i)_k = \sum_{k=0}^4 \biggl\{{\!4\! \atop \!k\!}\biggr\} \sum_{i=0}^n (i)_k = \sum_{k=0}^4 \biggl\{{\!4\! \atop \!k\!}\biggr\} \frac{(n{+}1)_{k+1}}{k{+}1} \\[10mu]
&\quad = \biggl\{{\!4\! \atop \!1\!}\biggr\} \frac{(n+1)_{2}}2
& = \biggl\{{\!4\! \atop \!1\!}\biggr\} \frac{(n{+}1)_{2}}2
+ \biggl\{{\!4\! \atop \!2\!}\biggr\} \frac{(n+1)_{3}}3
+ \biggl\{{\!4\! \atop \!2\!}\biggr\} \frac{(n{+}1)_{3}}3
+ \biggl\{{\!4\! \atop \!3\!}\biggr\} \frac{(n+1)_{4}}4
+ \biggl\{{\!4\! \atop \!3\!}\biggr\} \frac{(n{+}1)_{4}}4
+ \biggl\{{\!4\! \atop \!4\!}\biggr\} \frac{(n+1)_{5}}5 \\[8mu]
+ \biggl\{{\!4\! \atop \!4\!}\biggr\} \frac{(n{+}1)_{5}}5 \\[8mu]
&\quad = \frac12 (n+1)_{2} + \frac73 (n+1)_{3} + \frac64 (n+1)_{4} + \frac15 (n+1)_{5}
& = \frac12 (n{+}1)_{2} + \frac73 (n{+}1)_{3} + \frac64 (n{+}1)_{4} + \frac15 (n{+}1)_{5}\,.
\end{align}</math>
\end{align}</math>


Here the Stirling numbers can be computed from their definition as the number of partitions of 4 elements into ''k'' non-empty unlabeled subsets.
Here the Stirling numbers can be computed from their definition as the number of partitions of 4 elements into ''k'' non-empty unlabeled subsets.


In contrast, the sum <math display="inline">\sum_{i=0}^{n} i^k</math> in the standard basis is given by [[Faulhaber's formula]], which in general is more complex.
In contrast, the sum <math display="inline">\sum_{i=0}^n i^k</math> in the standard basis is given by [[Faulhaber's formula]], which in general is more complicated.


==As inverse matrices==
==As inverse matrices==
The Stirling numbers of the first and second kinds can be considered inverses of one another:
The Stirling numbers of the first and second kinds can be considered inverses of one another:


:<math>\sum_{j\geq 0} s(n,j) S(j,k) = \sum_{j\geq 0} (-1)^{n-j} \biggl[{n \atop j}\biggr] \biggl\{{\!j\! \atop \!k\!}\biggr\} = \delta_{nk}</math>
:<math>\sum_{j=k}^n s(n,j) S(j,k) = \sum_{j=k}^n (-1)^{n-j} \biggl[{n \atop j}\biggr] \biggl\{{\!j\! \atop \!k\!}\biggr\} = \delta_{n,k}</math>


and
and


:<math>\sum_{j\geq 0} S(n,j) s(j,k) = \sum_{j\geq 0} (-1)^{j-k} \biggl\{{\!n\! \atop \!j\!}\biggr\} \biggl[{j \atop k}\biggr]= \delta_{nk},</math>
:<math>\sum_{j=k}^n S(n,j) s(j,k) = \sum_{j=k}^n (-1)^{j-k} \biggl\{{\!n\! \atop \!j\!}\biggr\} \biggl[{j \atop k}\biggr]= \delta_{n,k},</math>


where <math>\delta_{nk}</math> is the [[Kronecker delta]]. These two relationships may be understood to be matrix inverse relationships. That is, let ''s'' be the [[lower triangular matrix]] of Stirling numbers of the first kind, whose matrix elements
where <math>\delta_{nk}</math> is the [[Kronecker delta]]. These two relationships may be understood to be matrix inverse relationships. That is, let ''s'' be the [[lower triangular matrix]] of Stirling numbers of the first kind, whose matrix elements
Line 112: Line 110:
| title=Handbook of Number Theory II | publisher=[[Kluwer Academic Publishers]] | year=2004
| title=Handbook of Number Theory II | publisher=[[Kluwer Academic Publishers]] | year=2004
| url=https://books.google.com/books?id=B2WZkvmFKk8C&dq=%22Stirling+numbers+of+the+third+kind%22&pg=PA464 | isbn=9781402025464 | page=464}}</ref>
| url=https://books.google.com/books?id=B2WZkvmFKk8C&dq=%22Stirling+numbers+of+the+third+kind%22&pg=PA464 | isbn=9781402025464 | page=464}}</ref>
By convention, <math>L(0,0)=1</math> and <math>L(n,k)=0</math> if <math>n>k</math> or <math>k = 0 < n</math>.
By convention, <math>L(0,0)=1</math> and <math>L(n,k)=0</math> if <math>n<k</math> or <math>k = 0 < n</math>.


These numbers are coefficients expressing falling factorials in terms of rising factorials and vice versa:
These numbers are coefficients expressing falling factorials in terms of rising factorials and vice versa:
Line 120: Line 118:
In particular, one formula is the inverse of the other, thus:
In particular, one formula is the inverse of the other, thus:


: <math>\sum_{j} L(n,j) \cdot (-1)^{j-k} L(j,k) = \delta_{nk}.</math>
: <math>\sum_{j=k}^n (-1)^{j-k} L(n,j) L(j,k) = \delta_{n,k}.</math>


Similarly, composing for example the change of basis from <math>x^{(n)}</math> to <math>x^n</math> with the change of basis from <math>x^n</math> to <math>(x)_{n}</math> gives the change of basis directly from <math>x^{(n)}</math> to <math>(x)_{n}</math>:
Similarly, composing the change of basis from <math>x^{(n)}</math> to <math>x^n</math> with the change of basis from <math>x^n</math> to <math>(x)_{n}</math> gives the change of basis directly from <math>x^{(n)}</math> to <math>(x)_{n}</math>:


:<math> L(n,k) = \sum_{j} \biggl[{n \atop j}\biggr] \biggl\{{\!j\! \atop \!k\!}\biggr\} ,</math>
:<math> L(n,k) = \sum_{j=k}^n \biggl[{n \atop j}\biggr] \biggl\{{\!j\! \atop \!k\!}\biggr\} ,</math>


In terms of matrices, if <math>L</math> denotes the matrix with entries <math>L_{nk}=L(n,k)</math> and <math>L^{-}</math> denotes the matrix with entries <math>L^{-}_{nk}=(-1)^{n-k}L(n,k)</math>, then one is the inverse of the other: <math> L^{-} = L^{-1}</math>.
and similarly for other compositions. In terms of matrices, if <math>L</math> denotes the matrix with entries <math>L_{nk}=L(n,k)</math> and <math>L^{-}</math> denotes the matrix with entries <math>L^{-}_{nk}=(-1)^{n-k}L(n,k)</math>, then one is the inverse of the other: <math> L^{-} = L^{-1}</math>.
Similarly, composing the matrix of unsigned Stirling numbers of the first kind with the matrix of Stirling numbers of the second kind gives the Lah numbers: <math>L = |s| \cdot S</math>.
Composing the matrix of unsigned Stirling numbers of the first kind with the matrix of Stirling numbers of the second kind gives the Lah numbers: <math>L = |s| \cdot S</math>.


The numbers <math display=inline>\left\{{\!n\! \atop \!k\!}\right\}, \left[{n \atop k}\right] , L(n,k)</math> can be defined as the number of partitions of ''n'' elements into ''k'' non-empty unlabeled subsets, each of which is unordered, [[Cyclic order|cyclically ordered]], or linearly ordered, respectively. In particular, this implies the following inequalities:
[[Enumerative combinatorics|Enumeratively]], <math display="inline">\left\{{\!n\! \atop \!k\!}\right\}, \left[{n \atop k}\right] , L(n,k)</math> can be defined as the number of partitions of ''n'' elements into ''k'' non-empty unlabeled subsets, where each subset is endowed with no order, a [[cyclic order]], or a linear order, respectively. In particular, this implies the inequalities:
: <math>\biggl\{{\!n\! \atop \!k\!}\biggr\} \leq \biggl[{n \atop k}\biggr] \leq L(n,k).</math>
: <math>\biggl\{{\!n\! \atop \!k\!}\biggr\} \leq \biggl[{n \atop k}\biggr] \leq L(n,k).</math>

==Inversion relations and the Stirling transform==

For any pair of sequences, <math>\{f_n\}</math> and <math>\{g_n\}</math>, related by a finite sum Stirling number formula given by

:<math>g_n = \sum_{k=0}^{n} \left\{\begin{matrix} n \\ k \end{matrix} \right\} f_k, </math>

for all integers <math>n \geq 0</math>, we have a corresponding [[generating function transformation#Inversion relations and generating function identities|inversion formula]] for <math>f_n</math> given by

:<math>f_n = \sum_{k=0}^{n} \left[\begin{matrix} n \\ k \end{matrix} \right] (-1)^{n-k} g_k. </math>

The lower indices could be any integer between <math display="inline">0</math> and <math display="inline">n</math>.

These inversion relations between the two sequences translate into functional equations between the sequence [[generating function|exponential generating functions]] given by the [[Stirling transform|Stirling (generating function) transform]] as

:<math>\widehat{G}(z) = \widehat{F}\left(e^z-1\right)</math>

and

:<math>\widehat{F}(z) = \widehat{G}\left(\log(1+z)\right). </math>

For <math>D = d/dx</math>, the [[differential operators]] <math>x^nD^n</math> and <math>(xD)^n</math> are related by the following formulas for all integers <math>n \geq 0</math>:<ref>''Concrete Mathematics'' exercise 13 of section 6. Note that this formula immediately implies the first positive-order Stirling number transformation given in the main article on [[generating function transformation#Derivative transformations|generating function transformations]].</ref>

:<math>
\begin{align}
(xD)^n &= \sum_{k=0}^n S(n, k) x^k D^k \\
x^n D^n &= \sum_{k=0}^n s(n, k) (xD)^k = (xD)_n = xD(xD - 1)\ldots (xD - n + 1)
\end{align}
</math>

Another pair of "''inversion''" relations involving the [[Stirling numbers]] relate the [[finite difference|forward differences]] and the ordinary <math>n^{th}</math> [[derivative]]s of a function, <math>f(x)</math>, which is analytic for all <math>x</math> by the formulas<ref>{{cite journal|last1=Olver|first1=Frank|first2=Daniel|last2=Lozier|first3=Ronald|last3=Boisvert|first4=Charles|last4=Clark|title=NIST Handbook of Mathematical Functions|journal=NIST Handbook of Mathematical Functions|date=2010|url=https://www.nist.gov/publications/nist-handbook-mathematical-functions}} (Section 26.8)</ref>

:<math>\frac{1}{k!} \frac{d^k}{dx^k} f(x) = \sum_{n=k}^{\infty} \frac{s(n, k)}{n!} \Delta^n f(x)</math>

:<math>\frac{1}{k!} \Delta^k f(x) = \sum_{n=k}^{\infty} \frac{S(n, k)}{n!} \frac{d^n}{dx^n} f(x). </math>

== Similar properties ==
{| class="wikitable"
|+Table of similarities
![[Stirling numbers of the first kind]]
![[Stirling numbers of the second kind]]
|-
|<math> \left[{n+1\atop k}\right] = n \left[{n\atop k}\right] + \left[{n\atop k-1}\right]</math>
|<math>\left\{{n+1\atop k}\right\} = k \left\{{ n \atop k }\right\} + \left\{{n\atop k-1}\right\} </math>
|-
|<math>\sum_{k=0}^n \left[{n\atop k}\right] = n!</math>
|<math>\sum_{k=0}^n \left\{ {n \atop k} \right\} = B_n</math>, where <math>B_n </math> is the n-th [[Bell number]]
|-
|<math>\sum_{k=0}^n \left[{n\atop k}\right] x^k = x^{(n)}</math>, where <math>\{x^{(n)}\}_{n\in\N} </math> is the [[Falling and rising factorials|rising factorials]]
|<math>\sum_{k=0}^n \left\{ {n \atop k} \right\} x^k = T_n(x)</math>, where <math>\{T_n\}_{n\in\N} </math> is the [[Touchard polynomials]]
|-
|<math> \left[{n\atop 0}\right] = \delta_n,\ \left[{n\atop n-1}\right] = \binom{n}{2},\ \left[{n\atop n}\right] = 1</math>
|<math> \left\{{n\atop 0}\right\} = \delta_n,\ \left\{{n\atop n-1}\right\} = \binom{n}{2},\ \left\{{n\atop n}\right\} = 1</math>
|-
|<math>\left[{n+1\atop k+1}\right] = \sum_{j=k}^n \left[{n\atop j}\right] \binom{j}{k} </math>
|<math>\left\{{n+1\atop k+1}\right\} = \sum_{j=k}^n \binom{n}{j} \left\{{ j \atop k }\right\} </math>
|-
|<math>\left[\begin{matrix} n+1 \\ k+1 \end{matrix} \right] = \sum_{j=k}^n \frac{n!}{j!} \left[{j \atop k} \right]</math>
|<math>
\left\{{n+1\atop k+1}\right\} = \sum_{j=k}^n (k+1)^{n-j} \left\{{j \atop k}\right\} </math>
|-
|<math>\left[{ n+k+1 \atop n }\right] = \sum_{j=0}^k (n+j) \left[{n+j \atop j}\right]</math>
|<math>\left\{{n+k+1 \atop k}\right\} = \sum_{j=0}^k j \left\{{ n+j \atop j }\right\}</math>
|-
|<math>\left[{n \atop l+m } \right] \binom{l+m}{l} = \sum_k \left[{k \atop l} \right] \left[{n-k \atop m } \right] \binom{n}{k} </math>
|<math>\left\{{n \atop l+m } \right\} \binom{l+m}{l} = \sum_k \left\{{k \atop l} \right\} \left\{{n-k \atop m } \right\} \binom{n}{k} </math>
|-
|<math>\left[{n+k \atop n} \right] \underset{n \to \infty}{\sim} \frac{n^{2k}}{2^k k!}. </math>
|<math> \left\{{n+k \atop n}\right\} \underset{n \to \infty}{\sim} \frac{n^{2k}}{2^k k!}.</math>
|-
|<math>\sum_{n=k}^\infty\left[{n\atop k}\right] \frac{x^n}{n!} = \frac{(-\log(1-x))^k}{k!}.</math>
|<math> \sum_{n=k}^\infty \left\{ {n \atop k} \right\} \frac{x^n}{n!} = \frac{(e^x-1)^k}{k!}.</math>
|-
|<math> \left[{n \atop k} \right] = \sum_{0 \leq i_1 < \ldots < i_{n-k} < n} i_1 i_2 \cdots i_{n-k}.</math>
|<math>
\left\{ {n \atop k} \right\} = \sum_{
\begin{array}{c}
c_1 + \ldots + c_k = n-k\\
c_1, \ldots,\ c_k\ \geq\ 0
\end{array}
} 1^{c_1} 2^{c_2} \cdots k^{c_k}
</math>
|}
See the specific articles for details.


==Symmetric formulae==
==Symmetric formulae==


Abramowitz and Stegun give the following symmetric formulae that relate the Stirling numbers of the first and second kind.<ref>{{Citation|editor-last1=Abramowitz|editor-first1=Milton|editor-last2=Stegun|editor-first2=Irene A.|last1=Goldberg|first1=K.|last2=Newman|first2=M|last3=Haynsworth|first3=E.|title=Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th printing|contribution=Stirling Numbers of the First Kind, Stirling Numbers of the Second Kind|publisher=Dover|year=1972|pages=824–825|location=New York}}</ref>
Abramowitz and Stegun give the following symmetric formulae that relate the Stirling numbers of the first and second kind.<ref>{{Citation |last1=Goldberg |first1=K. |title=Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th printing |pages=824–825 |year=1972 |editor-last1=Abramowitz |editor-first1=Milton |contribution=Stirling Numbers of the First Kind, Stirling Numbers of the Second Kind |location=New York |publisher=Dover |last2=Newman |first2=M |last3=Haynsworth |first3=E. |editor-last2=Stegun |editor-first2=Irene A.}}</ref>


:<math>s(n,k) = \sum_{j=0}^{n-k} (-1)^j {n-1+j \choose n-k+j} {2n-k \choose n-k-j} S(n-k+j,j)</math>
:<math> \left[{ n \atop k } \right] = \sum_{j=n}^{2n-k} (-1)^{j-k} \binom{j-1}{k-1} \binom{2n-k}{j} \left\{{ j-k \atop j-n} \right\} </math>


and
and


:<math>
:<math>S(n,k) = \sum_{j=0}^{n-k} (-1)^j {n-1+j \choose n-k+j} {2n-k \choose n-k-j} s(n-k+j,j).</math>
\left\{{n \atop k}\right\} = \sum_{j=n}^{2n-k} (-1)^{j-k} \binom{j-1}{k-1} \binom{2n-k}{j} \left[{j-k \atop j-n } \right]
</math>


==Stirling numbers with negative integral values==
==Stirling numbers with negative integral values==
Line 192: Line 276:
| 0
| 0
| 1
| 1
|-
|}
|}


Donald Knuth<ref name=":1" /> defined the more general Stirling numbers by extending a [[Stirling numbers of the second kind#Recurrence relation|recurrence relation]] to all integers. In this approach, <math display=inline> \left[{n \atop k}\right]</math> and <math display=inline>\left\{{\!n\! \atop \!k\!}\right\}</math> are zero if ''n'' is negative and ''k'' is nonnegative, or if ''n'' is nonnegative and ''k'' is negative, and so we have, for ''any'' integers ''n'' and ''k'',
Donald Knuth<ref name=":1" /> defined the more general Stirling numbers by extending a [[Stirling numbers of the second kind#Recurrence relation|recurrence relation]] to all integers. In this approach, <math display=inline> \left[{n \atop k}\right]</math> and <math display=inline>\left\{{\!n\! \atop \!k\!}\right\}</math> are zero if ''n'' is negative and ''k'' is nonnegative, or if ''n'' is nonnegative and ''k'' is negative, and so we have, for ''any'' integers ''n'' and ''k'',
Line 204: Line 287:
= \frac{(-1)^{n+1}}{n!}\sum_{i=1}^{n}\frac{(-1)^{i+1}}{i^k} \binom ni </math>,
= \frac{(-1)^{n+1}}{n!}\sum_{i=1}^{n}\frac{(-1)^{i+1}}{i^k} \binom ni </math>,


For example, <math display=inline>\left[{-5 \atop k}\right] = \frac1{120}\Bigl(5-\frac{10}{2^k}+\frac{10}{3^k}-\frac 5{4^k}+\frac 1{5^k}\Bigr).</math> This leads to the following table of values of <math display=inline>\left[{-n \atop k}\right]</math>.
For example, <math display=inline>\left[{-5 \atop k}\right] = \frac1{120}\Bigl(5-\frac{10}{2^k}+\frac{10}{3^k}-\frac 5{4^k}+\frac 1{5^k}\Bigr).</math> This leads to the following table of values of <math display=inline>\left[{n \atop k}\right]</math> for negative integral ''n''.
{| cellspacing="0" cellpadding="5" style="text-align:center;" class="wikitable"
{| cellspacing="0" cellpadding="5" style="text-align:center;" class="wikitable"
|-
|-
Line 248: Line 331:
| <math>\tfrac{874853}{25920000}</math>
| <math>\tfrac{874853}{25920000}</math>
| <math>\tfrac{58067611}{1555200000}</math>
| <math>\tfrac{58067611}{1555200000}</math>
|-
|}

|}
In this case <math display=inline>\sum_{n=1}^{\infty}\left[{-n \atop -k}\right]=B_{k} </math> where <math>B_{k}</math> is a [[Bell number]], and so one may define the negative Bell numbers by <math display=inline>\sum_{n=1}^{\infty}\left[{-n \atop k}\right]=:B_{-k}</math>.


In this case <math display=inline>\sum_{n=-1}^{-\infty}\left[{-n \atop -k}\right]=B_{k} </math> where <math>B_{k}</math> is a [[Bell number]], and so one may define the negative Bell numbers by <math display=inline>\sum_{n=-1}^{-\infty}\left[{-n \atop k}\right]=B_{-k}</math>. For example, this produces <math display=inline>\sum_{n=-1}^{-\infty}\left[{-n \atop 2}\right]=B_{-2}=0.421773\ldots</math>.
For example, this produces <math display=inline>\sum_{n=1}^{\infty}\left[{-n \atop 1}\right]=B_{-1}=\frac 1e\sum_{j=1}^\infty\frac1{j\cdot j!}=\frac 1e\int_0^1\frac{e^t-1}{t}dt=0.4848291\dots</math>, generally <math display=inline>B_{-k}=\frac 1e\sum_{j=1}^\infty\frac1{j^kj!} </math>.


==See also==
==See also==
Line 257: Line 341:
* [[Catalan number]]
* [[Catalan number]]
* [[Cycles and fixed points]]
* [[Cycles and fixed points]]
* [[Lah number]]
* [[Pochhammer symbol]]
* [[Pochhammer symbol]]
* [[Polynomial sequence]]
* [[Polynomial sequence]]
* [[Stirling transform]]
* [[Touchard polynomials]]
* [[Touchard polynomials]]
* [[Stirling permutation]]
* [[Stirling permutation]]
Line 308: Line 390:
* {{cite journal| first1=Michael Z. | last1=Spivey | title=Combinatorial sums and finite differences
* {{cite journal| first1=Michael Z. | last1=Spivey | title=Combinatorial sums and finite differences
|doi=10.1016/j.disc.2007.03.052 | journal=Discrete Math.
|doi=10.1016/j.disc.2007.03.052 | journal=Discrete Math.
|year=2007 | volume=307 | number=24 | pages=3130–3146| doi-access=free }}
|year=2007 | volume=307 | number=24 | pages=3130–3146| doi-access=free | citeseerx=10.1.1.134.8665 }}
* {{Cite OEIS|sequencenumber=A008275|name=Stirling numbers of first kind}}
* {{Cite OEIS|sequencenumber=A008275|name=Stirling numbers of first kind}}
* {{Cite OEIS|sequencenumber=A008277|name=Stirling numbers of 2nd kind}}
* {{Cite OEIS|sequencenumber=A008277|name=Stirling numbers of 2nd kind}}

Latest revision as of 03:13, 9 April 2024

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).[1] They were rediscovered and given a combinatorial meaning by Masanobu Saka in 1782.[2]

Two different sets of numbers bear this name: the Stirling numbers of the first kind and the Stirling numbers of the second kind. Additionally, Lah numbers are sometimes referred to as Stirling numbers of the third kind. Each kind is detailed in its respective article, this one serving as a description of relations between them.

A common property of all three kinds is that they describe coefficients relating three different sequences of polynomials that frequently arise in combinatorics. Moreover, all three can be defined as the number of partitions of n elements into k non-empty subsets, where each subset is endowed with a certain kind of order (no order, cyclical, or linear).

Notation[edit]

Several different notations for Stirling numbers are in use. Ordinary (signed) Stirling numbers of the first kind are commonly denoted:

Unsigned Stirling numbers of the first kind, which count the number of permutations of n elements with k disjoint cycles, are denoted:

Stirling numbers of the second kind, which count the number of ways to partition a set of n elements into k nonempty subsets:[3]

Abramowitz and Stegun use an uppercase and a blackletter , respectively, for the first and second kinds of Stirling number. The notation of brackets and braces, in analogy to binomial coefficients, was introduced in 1935 by Jovan Karamata and promoted later by Donald Knuth. (The bracket notation conflicts with a common notation for Gaussian coefficients.[4]) The mathematical motivation for this type of notation, as well as additional Stirling number formulae, may be found on the page for Stirling numbers and exponential generating functions.

Another infrequent notation is and .

Expansions of falling and rising factorials[edit]

Stirling numbers express coefficients in expansions of falling and rising factorials (also known as the Pochhammer symbol) as polynomials.

That is, the falling factorial, defined as is a polynomial in x of degree n whose expansion is

with (signed) Stirling numbers of the first kind as coefficients.

Note that by convention, because it is an empty product. The notations for the falling factorial and for the rising factorial are also often used.[5] (Confusingly, the Pochhammer symbol that many use for falling factorials is used in special functions for rising factorials.)

Similarly, the rising factorial, defined as is a polynomial in x of degree n whose expansion is

with unsigned Stirling numbers of the first kind as coefficients. One of these expansions can be derived from the other by observing that

Stirling numbers of the second kind express the reverse relations:

and

As change of basis coefficients[edit]

Considering the set of polynomials in the (indeterminate) variable x as a vector space, each of the three sequences

is a basis. That is, every polynomial in x can be written as a sum for some unique coefficients (similarly for the other two bases). The above relations then express the change of basis between them, as summarized in the following commutative diagram:

A diagram of how different Stirling numbers give coefficients for changing one basis of polynomials to another
A diagram of how different Stirling numbers give coefficients for changing one basis of polynomials to another

The coefficients for the two bottom changes are described by the Lah numbers below. Since coefficients in any basis are unique, one can define Stirling numbers this way, as the coefficients expressing polynomials of one basis in terms of another, that is, the unique numbers relating with falling and rising factorials as above.

Falling factorials define, up to scaling, the same polynomials as binomial coefficients: . The changes between the standard basis and the basis are thus described by similar formulas:

.

Example[edit]

Expressing a polynomial in the basis of falling factorials is useful for calculating sums of the polynomial evaluated at consecutive integers. Indeed, the sum of falling factorials with fixed k can expressed as another falling factorial (for )

This can be proved by induction.

For example, the sum of fourth powers of integers up to n (this time with n included), is:

Here the Stirling numbers can be computed from their definition as the number of partitions of 4 elements into k non-empty unlabeled subsets.

In contrast, the sum in the standard basis is given by Faulhaber's formula, which in general is more complicated.

As inverse matrices[edit]

The Stirling numbers of the first and second kinds can be considered inverses of one another:

and

where is the Kronecker delta. These two relationships may be understood to be matrix inverse relationships. That is, let s be the lower triangular matrix of Stirling numbers of the first kind, whose matrix elements The inverse of this matrix is S, the lower triangular matrix of Stirling numbers of the second kind, whose entries are Symbolically, this is written

Although s and S are infinite, so calculating a product entry involves an infinite sum, the matrix multiplications work because these matrices are lower triangular, so only a finite number of terms in the sum are nonzero.

Lah numbers[edit]

The Lah numbers are sometimes called Stirling numbers of the third kind.[6] By convention, and if or .

These numbers are coefficients expressing falling factorials in terms of rising factorials and vice versa:

and

As above, this means they express the change of basis between the bases and , completing the diagram. In particular, one formula is the inverse of the other, thus:

Similarly, composing the change of basis from to with the change of basis from to gives the change of basis directly from to :

and similarly for other compositions. In terms of matrices, if denotes the matrix with entries and denotes the matrix with entries , then one is the inverse of the other: . Composing the matrix of unsigned Stirling numbers of the first kind with the matrix of Stirling numbers of the second kind gives the Lah numbers: .

Enumeratively, can be defined as the number of partitions of n elements into k non-empty unlabeled subsets, where each subset is endowed with no order, a cyclic order, or a linear order, respectively. In particular, this implies the inequalities:

Inversion relations and the Stirling transform[edit]

For any pair of sequences, and , related by a finite sum Stirling number formula given by

for all integers , we have a corresponding inversion formula for given by

The lower indices could be any integer between and .

These inversion relations between the two sequences translate into functional equations between the sequence exponential generating functions given by the Stirling (generating function) transform as

and

For , the differential operators and are related by the following formulas for all integers :[7]

Another pair of "inversion" relations involving the Stirling numbers relate the forward differences and the ordinary derivatives of a function, , which is analytic for all by the formulas[8]

Similar properties[edit]

Table of similarities
Stirling numbers of the first kind Stirling numbers of the second kind
, where is the n-th Bell number
, where is the rising factorials , where is the Touchard polynomials

See the specific articles for details.

Symmetric formulae[edit]

Abramowitz and Stegun give the following symmetric formulae that relate the Stirling numbers of the first and second kind.[9]

and

Stirling numbers with negative integral values[edit]

The Stirling numbers can be extended to negative integral values, but not all authors do so in the same way.[10][11][12] Regardless of the approach taken, it is worth noting that Stirling numbers of first and second kind are connected by the relations:

when n and k are nonnegative integers. So we have the following table for :

k
n
−1 −2 −3 −4 −5
−1 1 1 1 1 1
−2 0 1 3 7 15
−3 0 0 1 6 25
−4 0 0 0 1 10
−5 0 0 0 0 1

Donald Knuth[12] defined the more general Stirling numbers by extending a recurrence relation to all integers. In this approach, and are zero if n is negative and k is nonnegative, or if n is nonnegative and k is negative, and so we have, for any integers n and k,

On the other hand, for positive integers n and k, David Branson[11] defined and (but not or ). In this approach, one has the following extension of the recurrence relation of the Stirling numbers of the first kind:

,

For example, This leads to the following table of values of for negative integral n.

k
n
0 1 2 3 4
−1 1 1 1 1 1
−2
−3
−4
−5

In this case where is a Bell number, and so one may define the negative Bell numbers by .

For example, this produces , generally .

See also[edit]

Citations[edit]

  1. ^ Mansour & Schork 2015, p. 5.
  2. ^ Mansour & Schork 2015, p. 4.
  3. ^ Ronald L. Graham, Donald E. Knuth, Oren Patashnik (1988) Concrete Mathematics, Addison-Wesley, Reading MA. ISBN 0-201-14236-8, p. 244.
  4. ^ Donald Knuth
  5. ^ Aigner, Martin (2007). "Section 1.2 - Subsets and binomial coefficients". A Course in Enumeration. Springer. pp. 561. ISBN 978-3-540-39032-9.
  6. ^ Sándor, Jozsef; Crstici, Borislav (2004). Handbook of Number Theory II. Kluwer Academic Publishers. p. 464. ISBN 9781402025464.
  7. ^ Concrete Mathematics exercise 13 of section 6. Note that this formula immediately implies the first positive-order Stirling number transformation given in the main article on generating function transformations.
  8. ^ Olver, Frank; Lozier, Daniel; Boisvert, Ronald; Clark, Charles (2010). "NIST Handbook of Mathematical Functions". NIST Handbook of Mathematical Functions. (Section 26.8)
  9. ^ Goldberg, K.; Newman, M; Haynsworth, E. (1972), "Stirling Numbers of the First Kind, Stirling Numbers of the Second Kind", in Abramowitz, Milton; Stegun, Irene A. (eds.), Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th printing, New York: Dover, pp. 824–825
  10. ^ Loeb, Daniel E. (1992) [Received 3 Nov 1989]. "A generalization of the Stirling numbers". Discrete Mathematics. 103 (3): 259–269. doi:10.1016/0012-365X(92)90318-A.
  11. ^ a b Branson, David (August 1994). "An extension of Stirling numbers" (PDF). The Fibonacci Quarterly. Archived (PDF) from the original on 2011-08-27. Retrieved Dec 6, 2017.
  12. ^ a b D.E. Knuth, 1992.

References[edit]

  • Rosen, Kenneth H., ed. (2018), Handbook of Discrete and Combinatorial Mathematics, CRC Press, ISBN 978-1-5848-8780-5
  • Mansour, Toufik; Schork, Mathias (2015), Commutation Relations, Normal Ordering, and Stirling Numbers, CRC Press, ISBN 978-1-4665-7989-7

Further reading[edit]