Dmitri Khomukha and Binomial coefficient: Difference between pages

From Wikipedia, the free encyclopedia
(Difference between pages)
Content deleted Content added
No edit summary
 
Undid revision 244367132 by 155.198.157.19 (talk)
 
Line 1: Line 1:
In [[mathematics]], the '''binomial coefficient''' <math> \tbinom nk </math> is the [[coefficient]] of the ''x''<sup>&nbsp;''k''</sup> term in the [[polynomial]] expansion of the [[binomial]] [[exponentiation|power]] (1&nbsp;+&nbsp;''x'')<sup>&nbsp;''n''</sup>.
{{Football player infobox
| playername= Dmitri Khomukha
| fullname = Dmitri Ivanovich Khomukha
| nickname =
| image=
| dateofbirth = {{Birth date and age|1969|8|23|mf=y}}
| cityofbirth = [[Ashgabat]]
| countryofbirth = [[Soviet Union]]
| height = {{height|meters=1.78}}
| currentclub = Retired
| position = [[Midfielder]]
| youthyears =
| youthclubs =
| years = 1985-1987<br />1988<br />1989<br />1989<br />1989-1994<br />1994<br />1995-1996<br />1997-2000<br />2001-2003<br />2004-2005
| clubs = [[Kolkhozchi Ashgabat]]<br />[[CSKA-2 Moscow]]<br />[[FC Karpaty Lviv|SKA Karpaty Lviv]]<br />[[CSKA-2 Moscow|Chaika-CSKA Moscow]]<br />[[FC Metalist Kharkiv]]<br />[[FC Erzu Grozny]]<br />[[FC Zenit Saint Petersburg|FC Zenit St. Petersburg]]<br />[[PFC CSKA Moscow]]<br />[[FC Shinnik Yaroslavl]]<br />[[FC Terek Grozny]]
| caps(goals) = 47 (1)<br />35 (3)<br />17 (1)<br />10 (0)<br />91 (3)<br />14 (2)<br />61 (4)<br />116 (27)<br />69 (16)<br />36 (4)
| nationalyears =
| nationalteam = [[Turkmenistan national football team|Turkmenistan]]
| nationalcaps(goals) =
| pcupdate = 13:33, 10 October 2008 (UTC)
| ntupdate =
}}


In [[combinatorics]], <math>\tbinom nk</math> is often called the '''choose function''' of ''n'' and ''k''; <math>\tbinom nk</math> is the number of ''k''-element [[subset]]s (the ''k''-[[combination]]s) of an ''n''-element set; that is, the number of ways that ''k'' things can be 'chosen' from a set of ''n'' things.
'''Dmitri Ivanovich Khomukha''' ({{lang-ru|Дмитрий Иванович Хомуха}}) (born [[August 23]], [[1969]] in [[Ashgabat]]) is a retired [[Turkmenistan|Turkmenistani]] footballer. He is [[Ukraine|Ukrainian]] by ethnicity and also has a [[Russia]]n citizenship.


==External links==
==Definition==
Given a non-negative integer ''n'' and an integer ''k'', the binomial coefficient is defined to be the natural number
* {{de icon}} [http://www.transfermarkt.de/de/spieler/24143/khomukhadmitri/profil.html Player page on transfermarkt.de]


:<math>
{{Lifetime|1969||Khomukha, Dmitri}}
{n \choose k} = \frac{n \cdot (n-1) \cdots (n-k+1)}
[[Category:Turkmenistani footballers]]
{k \cdot (k-1) \cdots 1} = \frac{n!}{k!(n-k)!} \quad \mbox{if}\ 0\leq k\leq n \qquad (1) </math>
[[Category:Turkmenistan international footballers]]
[[Category:FC Karpaty Lviv players]]
[[Category:Metalist Kharkiv footballers]]
[[Category:FC Zenit Saint Petersburg players]]
[[Category:PFC CSKA Moscow players]]
[[Category:FC Shinnik Yaroslavl players]]
[[Category:FC Terek Grozny players]]


and
{{Turkmenistan-footy-bio-stub}}
:<math> {n \choose k} = 0 \quad \mbox{if } k < 0 \mbox{ or } k>n</math>


where ''n''! denotes the [[factorial]] of ''n''.
[[ru: Хомуха, Дмитрий Иванович]]

Alternatively, a recursive definition can be written as

:<math>
{n \choose k} = {n-1 \choose k-1} + {n-1 \choose k}</math>
where
:<math>
{n \choose 0} = {n \choose n} = 1</math>

The notation <math>\tbinom nk</math> was introduced by [[Albert von Ettinghausen]] in 1826,<ref>[[Nicholas J. Higham]], ''Handbook of writing for the mathematical sciences'', SIAM. ISBN 0898714206, p. 25</ref> although these numbers were already known centuries before that (see [[Pascal's triangle]]). Alternative notations include C(''n'', ''k''), <sub>''n''</sub>''C''<sub>''k''</sub> or <math>C^{k}_{n}</math>, in all of which the C stands for [[combination]]s or ''choices''. Indeed, the function <math>(n,k)\mapsto\tbinom nk</math> is often called the '''choose function''', and <math>\tbinom nk</math> is often read as "''n'' choose ''k''".

The binomial coefficients are the [[coefficient]]s of the [[series expansion]] of a power of a binomial, hence the name:

:<math> (1+x)^n = \sum_{k=0}^\infty {n \choose k} x^k. \qquad (2) </math>

If the exponent ''n'' is a nonnegative integer then this infinite series is actually a finite sum as all terms with k>n are zero, but if the exponent ''n'' is negative or a non-integer, then it is an infinite series.
(See the articles on [[combination]] and on [[binomial theorem]]).
=== Combinatorial interpretation ===
The importance of the binomial coefficients (and the motivation for the alternate name 'choose') lies in the fact that <math>{\tbinom n k}</math> is the number of ways that ''k'' objects can be chosen from among ''n'' objects, regardless of order. More formally,

:<math>{\tbinom n k}</math> is the number of ''k''-element subsets of an ''n''-element set. <math>\qquad (1a)</math>

In fact, this property is often chosen as an '''alternative definition''' of the binomial coefficient, since from (1a) one may derive (1) as a corollary by a straightforward [[combinatorial proof#Benefit of the approach|combinatorial proof]]. For a colloquial demonstration, note that in the formula
:<math> {n \choose k} = \frac{n \cdot (n-1) \cdots (n-k+1)}{k \cdot (k-1) \cdots 1}, </math>
the numerator gives the number of ways to fill the ''k'' slots using the ''n'' options, where the slots are distinguishable from one another. Thus a pizza with mushrooms added before sausage is considered to be different from a pizza with sausage added before mushrooms. The denominator eliminates these repetitions because if the ''k'' slots are indistinguishable, then all of the ''k''! ways of arranging them are considered identical.

In the context of computer science, it also helps to see <math>{\tbinom n k}</math> as the number of strings consisting of ones and zeros with ''k'' ones and ''n''&minus;''k'' zeros. For each ''k''-element subset, K, of an ''n''-element set, N, the [[indicator function]], 1<sub>K</sub> : N→{0,1}, where 1<sub>K</sub>(''x'') = 1 whenever ''x'' in K and 0 otherwise, produces a unique bit string of length ''n'' with exactly ''k'' ones by feeding 1<sub>K</sub> with the ''n'' elements in a specific order.<ref>[http://planetmath.org/encyclopedia/BinomialCoefficient.html PlanetMath: binomial coefficient<!-- Bot generated title -->]</ref>

==Example==

:<math> {7 \choose 3} = \frac{7!}{3!(7-3)!}
= \frac{7 \cdot 6 \cdot 5 \cdot 4 \cdot 3 \cdot 2 \cdot 1}{(3 \cdot 2 \cdot 1)(4 \cdot 3 \cdot 2 \cdot 1)}
= \frac{7\cdot 6 \cdot 5}{3\cdot 2\cdot 1}
=\frac{210}6
= 35.</math>

The calculation of the binomial coefficient is conveniently arranged like this: ((((5/1)·6)/2)·7)/3 = (((5·6)/2)·7)/3 = ((30/2)·7)/3 = (15·7)/3 = 105/3 = 35, alternately dividing and multiplying with increasing integers. Each division produces an integer result which is itself a binomial coefficient.

== Derivation from binomial expansion ==

For exponent 1, (1+''x'')<sup>1</sup> is 1+''x''. For exponent 2, (1+''x'')<sup>2</sup> is (1+''x'')·(1+''x''), which forms terms as follows. The first factor supplies either a 1 or a ''x''; likewise for the second factor. Thus to form 1, the only possibility is to choose 1 from both factors;
To form ''x''<sup>2</sup>, the only possibility is to choose ''x'' from both factors. However, the ''x'' term can be formed by 1 from the first and ''x'' from the second factor, or ''x'' from the first and 1 from the second factor; thus it acquires a coefficient of 2. Proceeding to exponent 3, (1+''x'')<sup>3</sup> reduces to (1+''x'')<sup>2</sup>·(1+''x''), where we already know that (1+''x'')<sup>2</sup>= 1+2''x''+''x''<sup>2</sup>, giving an initial expansion of (1+''x'')·(1+2''x''+''x''<sup>2</sup>). Again the extremes, 1 and ''x''<sup>3</sup> arise in a unique way. However, the ''x'' term is either 1·2''x'' or ''x''·1, for a coefficient of 3; likewise ''x''<sup>2</sup> arises in two ways, summing the coefficients 2 and 1 to give 3.

This suggests an [[mathematical induction|induction]]. Thus for exponent ''n'', each term of (1+''x'')<sup>''n''</sup> has ''n''−''k'' factors of 1 and ''k'' factors of ''x''. If ''k'' is 0 or ''n'', the term ''x''<sup>''k''</sup> arises in only one way, and we get the terms 1 and ''x''<sup>''n''</sup>. So <math>{\tbinom n 0}=1</math> and <math>{\tbinom n n}=1.</math> If ''k'' is neither 0 nor ''n'', then the term ''x''<sup>''k''</sup> arises in (1+''x'')<sup>''n''</sup>=(1+''x'')·(1+''x'')<sup>''n''&minus;1</sup> in two ways, from 1·''x''<sup>''k''</sup> and from ''x''·''x''<sup>''k''&minus;1</sup>, summing the coefficients <math>{\tbinom {n-1} k}+{\tbinom {n-1}{k-1}}</math> to give <math>{\tbinom n k}</math>. This is the origin of Pascal's triangle, discussed below.

Another perspective is that to form ''x''<sup>''k''</sup> from ''n'' factors of (1+''x''), we must choose ''x'' from ''k'' of the factors and 1 from the rest. To count the possibilities, consider all ''n''! [[permutation]]s of the factors. Represent each permutation as a shuffled list of the numbers from 1 to ''n''. Select a 1 from the first ''n''−''k'' factors listed, and an ''x'' from the remaining ''k'' factors; in this way each permutation contributes to the term ''x''<sup>''k''</sup>. For example, the list 〈<u>4,1</u>,<u style="text-decoration:overline">2,3</u>〉 selects 1 from factors 4 and 1, and selects ''x'' from factors 2 and 3, as one way to form the term ''x''<sup>2</sup> like this: "(<u>1</u> + ''x'')·(1 + <u style="text-decoration:overline">''x'' </u>)·(1 + <u style="text-decoration:overline">''x'' </u>)·(<u>1</u> + ''x'')". But the distinct list 〈<u>1,4</u>,<u style="text-decoration:overline">3,2</u>〉 makes exactly the same selection; the binomial coefficient formula must remove this redundancy. The ''n''−''k'' factors for 1 have (''n''−''k'')! permutations, and the ''k'' factors for ''x'' have ''k''! permutations. Therefore ''n''!/(''n''−''k'')!''k''! is the number of distinct ways to form the term ''x''<sup>''k''</sup>.

A simpler explanation follows:
One can pick a random element out of ''n'' in exactly ''n'' ways, a second random element in ''n''&minus;1 ways, and so forth. Thus, ''k'' elements can be picked out of ''n'' in ''n''·(''n''&minus;1)···(''n''&minus;''k''+1)
ways. In this calculation, however, each order-independent selection occurs ''k''! times, as a list of ''k'' elements can be permuted in so many ways. Thus eq. (1) is obtained.

== Pascal's triangle ==

[[Pascal's rule]] is the important [[recurrence relation]]
:<math> {n \choose k} + {n \choose k+1} = {n+1 \choose k+1}, \qquad (3) </math>
which follows directly from the definition:
:<math>\begin{align} {n \choose k} + {n \choose k+1}
&{}= \frac{n!}{k!(n-k)!} + \frac{n!}{(k+1)!(n-(k+1))!} \\
&{} = \frac{n!(k+1)}{k!(n-k)!(k+1)} + \frac{n!(n-k)}{(k+1)!(n-(k+1))!(n-k)}\\
&{} = \frac{n!(k+1 + n-k)}{(k+1)!(n-k)!} \\
&{} = \frac{(n+1)!}{(k+1)!((n+1)-(k+1))!} \\
&{} = {n+1 \choose k+1}
\end{align} </math>

The recurrence relation just proved can be used to prove by [[mathematical induction]] that <math> \tbinom n k</math> is a natural number for all ''n'' and ''k'', (equivalent to the statement that k! divides the product of k consecutive integers), a fact that is not immediately obvious from the definition.

=== Combinatorial proof of Pascal's Rule ===

Let us count the ways of choosing k+1 objects from a set of size n+1. Paint one of the n+1 objects red. The subset of size k+1 either contains the red object or does not. There are n choose k+1 subsets that do not contain the red object (we must choose k+1 non-red objects from the n that are not red), and n choose k subsets that do contain the red object (after we have chosen the red object, it remains to choose k more from the remaining n). Hence
<math> {n \choose k} + {n \choose k+1} = {n+1 \choose k+1}, \qquad (3) </math>

Pascal's rule also gives rise to [[Pascal's triangle]]:
:{|
|-
|0: || || || || || || || || ||1|| || || || || || || ||
|-
|1: || || || || || || || ||1|| ||1|| || || || || || ||
|-
|2: || || || || || || ||1|| ||2|| ||1|| || || || || ||
|-
|3: || || || || || ||1|| ||3|| ||3|| ||1|| || || || ||
|-
|4: || || || || ||1|| ||4|| ||6|| ||4|| ||1|| || || ||
|-
|5: || || || ||1|| ||5|| ||10|| ||10|| ||5|| ||1|| || ||
|-
|6: || || ||1|| ||6|| ||15|| ||20|| ||15|| ||6|| ||1|| ||
|-
|7: || ||1&nbsp;|| ||7&nbsp;|| ||21|| ||35|| ||35|| ||21|| ||7&nbsp;|| ||1&nbsp;||
|-
|8: ||1&nbsp;|| ||8&nbsp;|| ||28|| ||56|| ||70|| ||56|| ||28|| ||8&nbsp;|| ||1&nbsp;
|} <!--There is a wider cell made with &nbsp; in 1-digit columns, so triangle becomes more graphically symmetrical -->

Row number ''n'' contains the numbers <math> \tbinom n k</math> for ''k'' = 0,…,''n''. It is constructed by starting with ones at the outside and then always adding two adjacent numbers and writing the sum directly underneath. This method allows the quick calculation of binomial coefficients without the need for fractions or multiplications. For instance, by looking at row number 5 of the triangle, one can quickly read off that
:(''x'' + ''y'')<sup>5</sup> = '''1''' ''x''<sup>5</sup> + '''5''' ''x''<sup>4</sup>''y'' + '''10''' ''x''<sup>3</sup>''y''<sup>2</sup> + '''10''' ''x''<sup>2</sup>''y''<sup>3</sup> + '''5''' ''x'' ''y''<sup>4</sup> + '''1''' ''y''<sup>5</sup>.
The differences between elements on other diagonals are the elements in the previous diagonal, as a consequence of the recurrence relation (3) above.

In the 1303 AD treatise ''Precious Mirror of the Four Elements'', [[Zhu Shijie]] mentioned the triangle as an ancient method for evaluating binomial coefficients indicating that the method was known to [[Chinese mathematician]]s five centuries before [[Blaise Pascal|Pascal]].

== Combinatorics and statistics ==

Binomial coefficients are of importance in [[combinatorics]], because they provide ready formulas for certain frequent counting problems:
* There are <math>\tbinom n k</math> ways to choose ''k'' elements from a set of ''n'' elements. See [[Combination]].
* There are <math>\tbinom {n+k-1}k</math> ways to choose ''k'' elements from a set of ''n'' if repetitions are allowed. See [[Multiset]].
* There are <math> \tbinom {n+k} k</math> [[string (computer science)|strings]] containing ''k'' ones and ''n'' zeros.
* There are <math> \tbinom {n+1} k</math> strings consisting of ''k'' ones and ''n'' zeros such that no two ones are adjacent.
* The [[Catalan number]]s are <math>\frac {\tbinom{2n}n}{n+1}.</math>
* The [[binomial distribution]] in [[statistics]] is <math>\tbinom n k p^k (1-p)^{n-k} \!.</math>
* The formula for a [[Bézier curve]].

== Formulas involving binomial coefficients ==

When n is an integer
:<math> \tbinom n k= \tbinom n {n-k},\qquad\qquad(4)</math>

This follows from (2) by using (1&nbsp;+&nbsp;''x'')<sup>''n''</sup> = ''x''<sup>''n''</sup>·(1&nbsp;+&nbsp;''x''<sup>&minus;1</sup>)<sup>''n''</sup>.
It is reflected in the symmetry of [[Pascal's triangle]].

Another formula is
:<math> \sum_{k=0}^n \tbinom n k = 2^n, \qquad\qquad(5)</math>

it is obtained from (2) using ''x'' = 1. This is equivalent to saying that the elements in one row of Pascal's triangle always add up to two raised to an integer power. A combinatorial interpretation of this fact is given by counting subsets of size 0, size 1, size 2, and so on up to size ''n'' of a set ''S'' of ''n'' elements. Since we count the number of subsets of size ''i'' for 0 ≤ ''i'' ≤ ''n'', this sum must be equal to the number of subsets of ''S'', which is known to be 2<sup>''n''</sup>.

The formula
:<math> \sum_{k=1}^n k \tbinom n k = n 2^{n-1} \qquad(6)</math>
follows from (2), after [[derivative|differentiating]] with respect to ''x'' and then substituting ''x'' = 1.

[[Vandermonde's identity]]
:<math> \sum_j \tbinom m j \tbinom{n-m}{k-j} = \tbinom n k \qquad (7a) </math>
is found by expanding (1&nbsp;+&nbsp;''x'')<sup>''m''</sup>&nbsp;(1&nbsp;+&nbsp;''x'')<sup>''n''&minus;''m''</sup> = (1&nbsp;+&nbsp;''x'')<sup>''n''</sup> with (2). As <math>\tbinom n k</math> is zero if ''k'' > ''n'', the sum is finite for integer n and m. Equation (7a) generalizes equation (3). It holds for arbitrary, complex-valued <math>m</math> and <math>n</math>, the [[Chu-Vandermonde identity]].

A related formula is

:<math> \sum_m \tbinom m j \tbinom {n-m}{k-j}= \tbinom {n+1}{k+1}. \qquad (7b) </math>

While equation (7a) is true for all values of ''m'', equation (7b) is true for all values of ''j''.

From expansion (7a) using ''n''=2''m'', ''k'' = ''m'', and (4), one finds
:<math> \sum_{j=0}^m \tbinom m j ^2 = \tbinom {2m} m. \qquad (8)</math>

Denote by ''F''(''n''&nbsp;+&nbsp;1) the [[Fibonacci number]]s. We obtain a formula about the diagonals of Pascal's triangle
:<math> \sum_{k=0}^n \tbinom {n-k} k = F(n+1). \qquad (9) </math>

This can be proved by [[mathematical induction|induction]] using (3).

Also using (3) and induction, one can show that
:<math> \sum_{j=k}^n \tbinom j k = \tbinom {n+1}{k+1}. \qquad (10) </math>

Again by (3) and induction, one can show that for ''k'' = 0, ... , ''n''&minus;1
:<math> \sum_{j=0}^k (-1)^j\tbinom n j = (-1)^k\tbinom {n-1}k \qquad(11)</math>

as well as
:<math> \sum_{j=0}^n (-1)^j\tbinom n j = 0 \qquad(12)</math>

which is itself a special case of the result that for any integer ''k'' = 1, ..., ''n'' &minus; 1,
:<math> \sum_{j=0}^n (-1)^j\tbinom n j j^k = 0 \qquad(13) </math>
which can be shown by differentiating (2) ''k'' times and setting ''x'' = &minus;1.

The [[infinite series]]
:<math>\sum_{j=0}^\infty \frac 1 {\tbinom {n+j}n}=\frac n{n-1}\qquad(14)</math>
is convergent for ''n'' ≥ 2.
It is the limiting case of the finite sum
:<math>\sum_{j=0}^k{\tbinom {n+j}n}^{-1}=(1-n^{-1})^{-1}( 1-{\tbinom {n+k}{n-1}}^{-1}).</math>
This formula is proved by [[mathematical induction]] on ''k''.

== Combinatorial identities involving binomial coefficients ==

Some identities have combinatorial proofs:

:<math>\sum_{k=q}^n \tbinom n k \tbinom k q = 2^{n-q}\tbinom n q\qquad(15)</math>

for <math>{n} \geq {q}.</math> The combinatorial proof goes as follows: the left side counts the number of ways of selecting a subset of <math>[n]</math> of at least ''q'' elements, and marking ''q'' elements among those selected. The right side counts the same parameter, because there are <math>\tbinom n q</math> ways of choosing a set of ''q'' marks and they occur in all subsets that additionally contain some subset of the remaining elements, of which there are <math>2^{n-q}.</math>

This reduces to (6) when <math>q=1.</math>

The identity (8) also has a combinatorial proof. The identity reads

:<math>\sum_{k=0}^n \tbinom n k ^2 = \tbinom {2n} n.</math>

Suppose you have <math>2n</math> empty squares arranged in a row and you want to mark (select) ''n'' of them. There are <math>\tbinom {2n}n</math> ways to do this. On the other hand, you may select your ''n'' squares by selecting ''k'' squares from among the first ''n'' and <math>n-k</math> squares from the remaining ''n'' squares. This gives
:<math>\sum_{k=0}^n\tbinom n k\tbinom n{n-k} = \tbinom {2n} n.</math>
Now apply (4) to get the result.

== Generating functions ==
The binomial coefficients can also be derived from the labelled case of the [[Fundamental Theorem of Combinatorial Enumeration]]. This is done by ''defining'' <math>C(n, k)</math> to be the number of ways of partitioning <math>[n]</math> into two subsets, the first of which has size ''k''. These partitions form a combinatorial class with the specification

:<math>\mathfrak{S}_2(\mathfrak{P}(\mathcal{Z})) =
\mathfrak{P}(\mathcal{Z}) \mathfrak{P}(\mathcal{Z}).</math>

Hence the exponential [[generating function]] ''B'' of the sum function of the binomial coefficients is given by
:<math> B(z) = \exp{z} \exp{z} = \exp(2z)\,.</math>

This immediately yields
:<math> \sum_{k=0}^{n} {n \choose k} = n! [z^n] \exp (2z) = 2^n,</math>

as expected. We mark the first subset with <math>\mathcal{U}</math> in order to obtain the binomial coefficients themselves, giving

:<math> \mathfrak{P}(\mathcal{U} \; \mathcal{Z}) \mathfrak{P}(\mathcal{Z}).</math>

This yields the bivariate generating function

:<math>B(z, u) = \exp uz \exp z\,.</math>

Extracting coefficients, we find that

:<math>{n \choose k} = n! [u^k] [z^n] \exp uz \exp z =
n! [z^n] \frac{z^k}{k!} \exp z</math>

or

:<math>
\frac{n!}{k!} [z^{n-k}] \exp z =
\frac{n!}{k! \, (n-k)!},</math>

again as expected. This derivation closely parallels that of the [[Stirling number]]s of the first and second kind, motivating the binomial-style notation that is used for these numbers.

== Divisors of binomial coefficients ==

The [[prime number|prime]] divisors of <math>\tbinom n k</math> can be interpreted as follows: if ''p'' is a prime number and ''p''<sup>''r''</sup> is the highest power of ''p'' which divides <math>\tbinom n k</math>, then ''r'' is equal to the number of natural numbers ''j'' such that the [[fractional part]] of ''k''/''p''<sup>''j''</sup> is bigger than the fractional part of ''n''/''p''<sup>''j''</sup>. In particular, <math>\tbinom n k</math> is always divisible by ''n''/[[greatest common divisor|gcd]](''n'',''k'').

A somewhat surprising result by [[David Singmaster]] (1974) is that any integer divides [[almost all]] binomial coefficients. More precisely, fix an integer ''d'' and let ''f''(''N'') denote the number of binomial coefficients <math>\tbinom n k</math> with ''n'' < ''N'' such that ''d'' divides <math>\tbinom n k</math>. Then
:<math> \lim_{N\to\infty} \frac{f(N)}{N(N+1)/2} = 1. </math>
Since the number of binomial coefficients <math>\tbinom n k</math> with ''n'' < ''N'' is ''N''(''N''+1) / 2, this implies that the density of binomial coefficients divisible by ''d'' goes to 1.

== Bounds for binomial coefficients ==

The following bounds for <math>\tbinom n k</math> hold:

<math>\left(\frac{n}{k}\right)^k \le {n \choose k} \le \frac{n^k}{k!} \le \left(\frac{n\cdot e}{k}\right)^k</math>

==Generalizations==
===Generalization to multinomials===

Binomial coefficients can be generalized to '''multinomial coefficients'''.
They are defined to be the number:
:<math>{n\choose k_1,k_2,\ldots,k_r} =\frac{n!}{k_1!k_2!\cdots k_r!}</math>
where
:<math>\sum_{i=1}^rk_i=n</math>

While the binomial coefficients represent the coefficients of (''x''+''y'')<sup>''n''</sup>, the multinomial coefficients
represent the coefficients of the polynomial
:(''x''<sub>1</sub> + ''x''<sub>2</sub> + ... + ''x''<sub>''r''</sub>)<sup>''n''</sup>.
See [[multinomial theorem]]. The case ''r'' = 2 gives binomial coefficients:
:<math>{n\choose k_1,k_2}={n\choose k_1, n-k_1}={n\choose k_1}= {n\choose k_2}</math>

The combinatorial interpretation of multinomial coefficients is distribution of ''n'' distinguishable elements over ''r'' (distinguishable) containers, each containing exactly ''k<sub>i</sub>'' elements, where ''i'' is the index of the container.

Multinomial coefficients have many properties similar to these of binomial coefficients, for example the recurrence relation:
:<math>{n\choose k_1,k_2,\ldots,k_r} ={n-1\choose k_1-1,k_2,\ldots,k_r}+{n-1\choose k_1,k_2-1,\ldots,k_r}+\ldots+{n-1\choose k_1,k_2,\ldots,k_r-1}</math>
and symmetry:
:<math>{n\choose k_1,k_2,\ldots,k_r} ={n\choose k_{\sigma_1},k_{\sigma_2},\ldots,k_{\sigma_r}}</math>
where <math>(\sigma_i)</math> is a permutation of (1,2,...,''r'').

=== Generalization to negative integers ===
If <math>k \geq 0</math>, then
<math> {n \choose k} = \frac{n(n-1) \dots (n-k+1)}{1 . 2 \dots k}= (-1)^k {-n+k-1 \choose k}</math> extends to all <math> n </math>.

The binomial coefficient extends to <math>k \leq 0</math> via
:<math>
{n \choose k}=
\begin{cases}
(-1)^{n-k} {-k-1 \choose n-k} \quad \mbox{if } n \geq k,\\
(-1)^{n-k} {-k-1 \choose -n-1} \quad \mbox{if } n \leq -1.
\end{cases}</math>

Notice in particular, that
:<math>{n \choose k}=0 \quad \mbox{iff }
\begin{cases}
n \geq 0 \mbox{ and } n < k, \\
n \geq 0 \mbox{ and } k < 0, \\
n < 0 \mbox{ and } n < k < 0.
\end{cases}</math>

This gives rise to the Pascal Hexagon or Pascal Windmill.
* {{cite book | author=Hilton, Holton and Pedersen | title=Mathematical Reflections | publisher=Springer | year=1997 | id=ISBN 0-387-94770-1}}

=== Generalization to real and complex argument ===
The binomial coefficient <math>{z\choose k}</math> can be defined for any [[complex number]] ''z'' and any [[natural number]] ''k'' as follows:
:<math>{z\choose k} = \prod_{n=1}^{k}{z-k+n\over n}= \frac{z(z-1)(z-2)\cdots (z-k+1)}{k!}. \qquad (14) </math>

This generalization is known as the '''generalized binomial coefficient''' and is used in the formulation of the [[binomial theorem]] and satisfies properties (3) and (7).

Alternatively, the infinite product
:<math>(-1)^k {z \choose k}= {-z+k-1 \choose k} = \frac{1}{\Gamma(-z)} \frac{1}{(k+1)^{z+1}} \prod_{j=k+1} \frac{(1+\frac{1}{j})^{-z-1}}{1-\frac{z+1}{j}}</math>
may be used to generalize the binomial coefficient. This formula discloses that asymptotically <math>{z \choose k} \approx \frac{(-1)^k}{\Gamma(-z) k^{z+1}}</math> as <math>k \to \infty</math>.

For fixed ''k'', the expression <math>f(z)={z\choose k}</math> is a [[polynomial]] in ''z'' of degree ''k'' with [[rational number|rational]] coefficients.
''f''(''z'') is the unique polynomial of degree ''k'' satisfying

:''f''(0) = ''f''(1) = ... = ''f''(''k'' &minus; 1) = 0 and ''f''(''k'') = 1.

Any polynomial ''p''(''z'') of degree ''d'' can be written in the form

:<math> p(z) = \sum_{k=0}^{d} a_k {z\choose k}. </math>

This is important in the theory of [[difference equation]]s and [[finite difference]]s, and can be seen as a discrete analog of [[Taylor's theorem]]. It is closely related to [[Newton's polynomial]]. Alternating sums of this form may be expressed as the [[Nörlund-Rice integral]].

In particular, one can express the product of binomial coefficients as such a linear combination:

:<math> {x\choose m} {x\choose n} = \sum_{k=0}^m {m+n-k\choose k,m-k,n-k} {x\choose m+n-k}</math>

where the connection coefficients are [[Multinomial theorem|multinomial coefficients]]. In terms of labelled combinatorial objects, the connection coefficients represent the number of ways to assign ''m+n-k'' labels to a pair of labelled combinatorial objects of weight ''m'' and ''n'' respectively, that have had their first ''k'' labels identified, or glued together, in order to get a new labelled combinatorial object of weight ''m+n-k''. (That is, to separate the labels into 3 portions to be applied to the glued part, the unglued part of the first object, and the unglued part of the second object.) In this regard, binomial coefficients are to exponential generating series what [[falling factorial]]s are to ordinary generating series.

==== Newton's binomial series ====
Newton's [[binomial series]], named after [[Sir Isaac Newton]], is one of the simplest [[Newton series]]:

:<math> (1+z)^{\alpha} = \sum_{n=0}^{\infty}{\alpha\choose n}z^n = 1+{\alpha\choose1}z+{\alpha\choose 2}z^2+\cdots.</math>

The identity can be obtained by showing that both sides satisfy the differential equation (1+''z'') ''f'''(''z'') = α ''f''(''z'').
The [[radius of convergence]] of this series is 1. An alternative expression is

:<math>\frac{1}{(1-z)^{\alpha+1}} = \sum_{n=0}^{\infty}{n+\alpha \choose n}z^n</math>

where the identity

:<math>{n \choose k} = (-1)^k {k-n-1 \choose k}</math>

is applied.

The formula for the binomial series was etched onto Newton's gravestone in [[Westminster Abbey]] in 1727.

==== Two real or complex valued arguments ====
The binomial coefficient is generalized to two real or complex valued arguments using [[gamma function]] or [[Beta function]] via
:<math>{x \choose y}:= \frac{\Gamma(x+1)}{\Gamma(y+1) \Gamma(x-y+1)}= \frac{1}{(x+1) \Beta(x-y+1,y+1)}.</math>
This definition inherits these following additional properties from <math>\Gamma</math>:
:<math>{x \choose y}= \frac{\sin (y \pi)}{\sin(x \pi)} {-y-1 \choose -x-1}= \frac{\sin((x-y) \pi)}{\sin (x \pi)} {y-x-1 \choose y};</math>
moreover,
:<math>{x \choose y} \cdot {y \choose x}= \frac{\sin((x-y) \pi)}{(x-y) \pi}</math>.

=== Generalization to ''q''-series ===

The binomial coefficient has a [[q-analog]] generalization known as the [[Gaussian binomial]].

=== Generalization to infinite cardinals ===

The definition of the binomial coefficient can be generalized to [[Cardinal Number|infinite cardinals]] by defining:

:<math>{\alpha \choose \beta} = | \{ B \subseteq A : |B| = \beta \} |</math>

where A is some set with [[cardinality]] <math>\alpha</math>. One can show that the generalized binomial coefficient is well-defined, in the sense that no matter what set we choose to represent the cardinal number <math>\alpha</math>, <math>{\alpha \choose \beta}</math> will remain the same. For finite cardinals, this definition coincides with the standard definition of the binomial coefficient.

Assuming the [[Axiom of Choice]], one can show that <math>{\alpha \choose \alpha} = 2^{\alpha}</math> for any infinite cardinal <math>\alpha</math>.

==Binomial coefficient in programming languages==
The notation <math> {n \choose k} </math> is convenient in handwriting but inconvenient for [[typewriter]]s and [[computer terminal]]s. Many [[programming language]]s do not offer a standard [[subroutine]] for computing the binomial coefficient, but for example the [[J programming language]] uses the exclamation mark: k ! n .

Naive implementations, such as the following snippet in [[C (programming language)|C]]:
<pre><nowiki>
int choose(int n, int k) {
return factorial(n) / (factorial(k) * factorial(n - k));
}
</nowiki></pre>

are prone to overflow errors, severely restricting the range of input values. A direct implementation of the first definition works well:

<pre><nowiki>
unsigned long long choose(unsigned n, unsigned k) {
if (k > n)
return 0;

if (k > n/2)
k = n-k; // faster

long double accum = 1;
for (unsigned i = 1; i <= k; i++)
accum = accum * (n-k+i) / i;

return accum + 0.5; // avoid rounding error
}
</nowiki></pre>

==See also==
* [[Combination]]
* [[Central binomial coefficient]]
* [[Binomial transform]]
* [[Table of Newtonian series]]
* [[List of factorial and binomial topics]]
* [[Multiplicities of entries in Pascal's triangle]]
* [[Binomial theorem]]

== References ==

<references/>
* ''This article incorporates material from the following [[PlanetMath]] articles, which are licensed under the [[Wikipedia:Text of the GNU Free Documentation License|GFDL]]: [http://planetmath.org/?op=getobj&amp;from=objects&amp;id=273 Binomial Coefficient], [http://planetmath.org/?op=getobj&amp;from=objects&amp;id=4074 Bounds for binomial coefficients], [http://planetmath.org/?op=getobj&amp;from=objects&amp;id=6744 Proof that C(n,k) is an integer], [http://planetmath.org/?op=getobj&amp;from=objects&amp;id=6309 Generalized binomial coefficients].''
* {{cite book | first=Donald E.| last=Knuth | authorlink=Donald Knuth
| title=The Art of Computer Programming, Volume 1: ''Fundamental Algorithms'' | edition=Third Edition | publisher=Addison-Wesley
| year=1997 | isbn= 0-201-89683-4 | pages=52–74
}}
* {{citation | first=Ronald L. | last=Graham | authorlink= Ronald Graham
| first2=Donald E .| last2=Knuth | author2-link= Donald Knuth
| first3=Oren | last3=Patashnik | author3-link= Oren Patashnik
| title=Concrete Mathematics | publisher=Addison Wesley | year=1989 | isbn= 0-201-14236-8 | pages=153–242
}}
* {{cite journal | first=David | last=Singmaster | authorlink=David Singmaster
| title=Notes on binomial coefficients. III. Any integer divides almost all binomial coefficients
| journal=J. London Math. Soc. (2) | volume=8 | year=1974 | pages=555–560
| doi=10.1112/jlms/s2-8.3.555
}}
* {{cite book | first=Victor | last=Bryant | authorlink=Victor Bryant
| title=Aspects of combinatorics | publisher= Cambridge University Press | year=1993
}}

[[Category:Combinatorics]]
[[Category:Factorial and binomial topics]]
[[Category:Integer sequences]]
[[Category:Triangles of numbers]]

[[bn:দ্বিপদী সহগ]]
[[bg:Биномен коефициент]]
[[ca:Coeficient binomial]]
[[cv:Биномлă коэффициентсем]]
[[cs:Kombinační číslo]]
[[da:Binomialkoefficient]]
[[de:Binomialkoeffizient]]
[[es:Coeficiente binomial]]
[[eo:Binoma koeficiento]]
[[fr:Coefficient binomial]]
[[ko:이항계수]]
[[it:Coefficiente binomiale]]
[[lt:Deriniai]]
[[nl:Binomiaalcoëfficiënt]]
[[no:Binomialkoeffisient]]
[[pl:Symbol Newtona]]
[[ro:Coeficient binomial]]
[[ru:Биномиальный коэффициент]]
[[sl:Binomski koeficient]]
[[sr:Биномни коефицијент]]
[[fi:Binomikerroin]]
[[sv:Binomialkoefficient]]
[[uk:Біноміальний коефіцієнт]]
[[zh:二項式係數]]

Revision as of 13:36, 10 October 2008

In mathematics, the binomial coefficient is the coefficient of the x k term in the polynomial expansion of the binomial power (1 + x) n.

In combinatorics, is often called the choose function of n and k; is the number of k-element subsets (the k-combinations) of an n-element set; that is, the number of ways that k things can be 'chosen' from a set of n things.

Definition

Given a non-negative integer n and an integer k, the binomial coefficient is defined to be the natural number

and

where n! denotes the factorial of n.

Alternatively, a recursive definition can be written as

where

The notation was introduced by Albert von Ettinghausen in 1826,[1] although these numbers were already known centuries before that (see Pascal's triangle). Alternative notations include C(n, k), nCk or , in all of which the C stands for combinations or choices. Indeed, the function is often called the choose function, and is often read as "n choose k".

The binomial coefficients are the coefficients of the series expansion of a power of a binomial, hence the name:

If the exponent n is a nonnegative integer then this infinite series is actually a finite sum as all terms with k>n are zero, but if the exponent n is negative or a non-integer, then it is an infinite series. (See the articles on combination and on binomial theorem).

Combinatorial interpretation

The importance of the binomial coefficients (and the motivation for the alternate name 'choose') lies in the fact that is the number of ways that k objects can be chosen from among n objects, regardless of order. More formally,

is the number of k-element subsets of an n-element set.

In fact, this property is often chosen as an alternative definition of the binomial coefficient, since from (1a) one may derive (1) as a corollary by a straightforward combinatorial proof. For a colloquial demonstration, note that in the formula

the numerator gives the number of ways to fill the k slots using the n options, where the slots are distinguishable from one another. Thus a pizza with mushrooms added before sausage is considered to be different from a pizza with sausage added before mushrooms. The denominator eliminates these repetitions because if the k slots are indistinguishable, then all of the k! ways of arranging them are considered identical.

In the context of computer science, it also helps to see as the number of strings consisting of ones and zeros with k ones and nk zeros. For each k-element subset, K, of an n-element set, N, the indicator function, 1K : N→{0,1}, where 1K(x) = 1 whenever x in K and 0 otherwise, produces a unique bit string of length n with exactly k ones by feeding 1K with the n elements in a specific order.[2]

Example

The calculation of the binomial coefficient is conveniently arranged like this: ((((5/1)·6)/2)·7)/3 = (((5·6)/2)·7)/3 = ((30/2)·7)/3 = (15·7)/3 = 105/3 = 35, alternately dividing and multiplying with increasing integers. Each division produces an integer result which is itself a binomial coefficient.

Derivation from binomial expansion

For exponent 1, (1+x)1 is 1+x. For exponent 2, (1+x)2 is (1+x)·(1+x), which forms terms as follows. The first factor supplies either a 1 or a x; likewise for the second factor. Thus to form 1, the only possibility is to choose 1 from both factors; To form x2, the only possibility is to choose x from both factors. However, the x term can be formed by 1 from the first and x from the second factor, or x from the first and 1 from the second factor; thus it acquires a coefficient of 2. Proceeding to exponent 3, (1+x)3 reduces to (1+x)2·(1+x), where we already know that (1+x)2= 1+2x+x2, giving an initial expansion of (1+x)·(1+2x+x2). Again the extremes, 1 and x3 arise in a unique way. However, the x term is either 1·2x or x·1, for a coefficient of 3; likewise x2 arises in two ways, summing the coefficients 2 and 1 to give 3.

This suggests an induction. Thus for exponent n, each term of (1+x)n has nk factors of 1 and k factors of x. If k is 0 or n, the term xk arises in only one way, and we get the terms 1 and xn. So and If k is neither 0 nor n, then the term xk arises in (1+x)n=(1+x)·(1+x)n−1 in two ways, from 1·xk and from x·xk−1, summing the coefficients to give . This is the origin of Pascal's triangle, discussed below.

Another perspective is that to form xk from n factors of (1+x), we must choose x from k of the factors and 1 from the rest. To count the possibilities, consider all n! permutations of the factors. Represent each permutation as a shuffled list of the numbers from 1 to n. Select a 1 from the first nk factors listed, and an x from the remaining k factors; in this way each permutation contributes to the term xk. For example, the list 〈4,1,2,3〉 selects 1 from factors 4 and 1, and selects x from factors 2 and 3, as one way to form the term x2 like this: "(1 + x)·(1 + x )·(1 + x )·(1 + x)". But the distinct list 〈1,4,3,2〉 makes exactly the same selection; the binomial coefficient formula must remove this redundancy. The nk factors for 1 have (nk)! permutations, and the k factors for x have k! permutations. Therefore n!/(nk)!k! is the number of distinct ways to form the term xk.

A simpler explanation follows: One can pick a random element out of n in exactly n ways, a second random element in n−1 ways, and so forth. Thus, k elements can be picked out of n in n·(n−1)···(nk+1) ways. In this calculation, however, each order-independent selection occurs k! times, as a list of k elements can be permuted in so many ways. Thus eq. (1) is obtained.

Pascal's triangle

Pascal's rule is the important recurrence relation

which follows directly from the definition:

The recurrence relation just proved can be used to prove by mathematical induction that is a natural number for all n and k, (equivalent to the statement that k! divides the product of k consecutive integers), a fact that is not immediately obvious from the definition.

Combinatorial proof of Pascal's Rule

Let us count the ways of choosing k+1 objects from a set of size n+1. Paint one of the n+1 objects red. The subset of size k+1 either contains the red object or does not. There are n choose k+1 subsets that do not contain the red object (we must choose k+1 non-red objects from the n that are not red), and n choose k subsets that do contain the red object (after we have chosen the red object, it remains to choose k more from the remaining n). Hence

Pascal's rule also gives rise to Pascal's triangle:

0: 1
1: 1 1
2: 1 2 1
3: 1 3 3 1
4: 1 4 6 4 1
5: 1 5 10 10 5 1
6: 1 6 15 20 15 6 1
7: 21 35 35 21
8: 28 56 70 56 28

Row number n contains the numbers for k = 0,…,n. It is constructed by starting with ones at the outside and then always adding two adjacent numbers and writing the sum directly underneath. This method allows the quick calculation of binomial coefficients without the need for fractions or multiplications. For instance, by looking at row number 5 of the triangle, one can quickly read off that

(x + y)5 = 1 x5 + 5 x4y + 10 x3y2 + 10 x2y3 + 5 x y4 + 1 y5.

The differences between elements on other diagonals are the elements in the previous diagonal, as a consequence of the recurrence relation (3) above.

In the 1303 AD treatise Precious Mirror of the Four Elements, Zhu Shijie mentioned the triangle as an ancient method for evaluating binomial coefficients indicating that the method was known to Chinese mathematicians five centuries before Pascal.

Combinatorics and statistics

Binomial coefficients are of importance in combinatorics, because they provide ready formulas for certain frequent counting problems:

  • There are ways to choose k elements from a set of n elements. See Combination.
  • There are ways to choose k elements from a set of n if repetitions are allowed. See Multiset.
  • There are strings containing k ones and n zeros.
  • There are strings consisting of k ones and n zeros such that no two ones are adjacent.
  • The Catalan numbers are
  • The binomial distribution in statistics is
  • The formula for a Bézier curve.

Formulas involving binomial coefficients

When n is an integer

This follows from (2) by using (1 + x)n = xn·(1 + x−1)n. It is reflected in the symmetry of Pascal's triangle.

Another formula is

it is obtained from (2) using x = 1. This is equivalent to saying that the elements in one row of Pascal's triangle always add up to two raised to an integer power. A combinatorial interpretation of this fact is given by counting subsets of size 0, size 1, size 2, and so on up to size n of a set S of n elements. Since we count the number of subsets of size i for 0 ≤ in, this sum must be equal to the number of subsets of S, which is known to be 2n.

The formula

follows from (2), after differentiating with respect to x and then substituting x = 1.

Vandermonde's identity

is found by expanding (1 + x)m (1 + x)nm = (1 + x)n with (2). As is zero if k > n, the sum is finite for integer n and m. Equation (7a) generalizes equation (3). It holds for arbitrary, complex-valued and , the Chu-Vandermonde identity.

A related formula is

While equation (7a) is true for all values of m, equation (7b) is true for all values of j.

From expansion (7a) using n=2m, k = m, and (4), one finds

Denote by F(n + 1) the Fibonacci numbers. We obtain a formula about the diagonals of Pascal's triangle

This can be proved by induction using (3).

Also using (3) and induction, one can show that

Again by (3) and induction, one can show that for k = 0, ... , n−1

as well as

which is itself a special case of the result that for any integer k = 1, ..., n − 1,

which can be shown by differentiating (2) k times and setting x = −1.

The infinite series

is convergent for n ≥ 2. It is the limiting case of the finite sum

This formula is proved by mathematical induction on k.

Combinatorial identities involving binomial coefficients

Some identities have combinatorial proofs:

for The combinatorial proof goes as follows: the left side counts the number of ways of selecting a subset of of at least q elements, and marking q elements among those selected. The right side counts the same parameter, because there are ways of choosing a set of q marks and they occur in all subsets that additionally contain some subset of the remaining elements, of which there are

This reduces to (6) when

The identity (8) also has a combinatorial proof. The identity reads

Suppose you have empty squares arranged in a row and you want to mark (select) n of them. There are ways to do this. On the other hand, you may select your n squares by selecting k squares from among the first n and squares from the remaining n squares. This gives

Now apply (4) to get the result.

Generating functions

The binomial coefficients can also be derived from the labelled case of the Fundamental Theorem of Combinatorial Enumeration. This is done by defining to be the number of ways of partitioning into two subsets, the first of which has size k. These partitions form a combinatorial class with the specification

Hence the exponential generating function B of the sum function of the binomial coefficients is given by

This immediately yields

as expected. We mark the first subset with in order to obtain the binomial coefficients themselves, giving

This yields the bivariate generating function

Extracting coefficients, we find that

or

again as expected. This derivation closely parallels that of the Stirling numbers of the first and second kind, motivating the binomial-style notation that is used for these numbers.

Divisors of binomial coefficients

The prime divisors of can be interpreted as follows: if p is a prime number and pr is the highest power of p which divides , then r is equal to the number of natural numbers j such that the fractional part of k/pj is bigger than the fractional part of n/pj. In particular, is always divisible by n/gcd(n,k).

A somewhat surprising result by David Singmaster (1974) is that any integer divides almost all binomial coefficients. More precisely, fix an integer d and let f(N) denote the number of binomial coefficients with n < N such that d divides . Then

Since the number of binomial coefficients with n < N is N(N+1) / 2, this implies that the density of binomial coefficients divisible by d goes to 1.

Bounds for binomial coefficients

The following bounds for hold:

Generalizations

Generalization to multinomials

Binomial coefficients can be generalized to multinomial coefficients. They are defined to be the number:

where

While the binomial coefficients represent the coefficients of (x+y)n, the multinomial coefficients represent the coefficients of the polynomial

(x1 + x2 + ... + xr)n.

See multinomial theorem. The case r = 2 gives binomial coefficients:

The combinatorial interpretation of multinomial coefficients is distribution of n distinguishable elements over r (distinguishable) containers, each containing exactly ki elements, where i is the index of the container.

Multinomial coefficients have many properties similar to these of binomial coefficients, for example the recurrence relation:

and symmetry:

where is a permutation of (1,2,...,r).

Generalization to negative integers

If , then extends to all .

The binomial coefficient extends to via

Notice in particular, that

This gives rise to the Pascal Hexagon or Pascal Windmill.

  • Hilton, Holton and Pedersen (1997). Mathematical Reflections. Springer. ISBN 0-387-94770-1.

Generalization to real and complex argument

The binomial coefficient can be defined for any complex number z and any natural number k as follows:

This generalization is known as the generalized binomial coefficient and is used in the formulation of the binomial theorem and satisfies properties (3) and (7).

Alternatively, the infinite product

may be used to generalize the binomial coefficient. This formula discloses that asymptotically as .

For fixed k, the expression is a polynomial in z of degree k with rational coefficients.

f(z) is the unique polynomial of degree k satisfying

f(0) = f(1) = ... = f(k − 1) = 0 and f(k) = 1.

Any polynomial p(z) of degree d can be written in the form

This is important in the theory of difference equations and finite differences, and can be seen as a discrete analog of Taylor's theorem. It is closely related to Newton's polynomial. Alternating sums of this form may be expressed as the Nörlund-Rice integral.

In particular, one can express the product of binomial coefficients as such a linear combination:

where the connection coefficients are multinomial coefficients. In terms of labelled combinatorial objects, the connection coefficients represent the number of ways to assign m+n-k labels to a pair of labelled combinatorial objects of weight m and n respectively, that have had their first k labels identified, or glued together, in order to get a new labelled combinatorial object of weight m+n-k. (That is, to separate the labels into 3 portions to be applied to the glued part, the unglued part of the first object, and the unglued part of the second object.) In this regard, binomial coefficients are to exponential generating series what falling factorials are to ordinary generating series.

Newton's binomial series

Newton's binomial series, named after Sir Isaac Newton, is one of the simplest Newton series:

The identity can be obtained by showing that both sides satisfy the differential equation (1+z) f'(z) = α f(z).

The radius of convergence of this series is 1. An alternative expression is

where the identity

is applied.

The formula for the binomial series was etched onto Newton's gravestone in Westminster Abbey in 1727.

Two real or complex valued arguments

The binomial coefficient is generalized to two real or complex valued arguments using gamma function or Beta function via

This definition inherits these following additional properties from :

moreover,

.

Generalization to q-series

The binomial coefficient has a q-analog generalization known as the Gaussian binomial.

Generalization to infinite cardinals

The definition of the binomial coefficient can be generalized to infinite cardinals by defining:

where A is some set with cardinality . One can show that the generalized binomial coefficient is well-defined, in the sense that no matter what set we choose to represent the cardinal number , will remain the same. For finite cardinals, this definition coincides with the standard definition of the binomial coefficient.

Assuming the Axiom of Choice, one can show that for any infinite cardinal .

Binomial coefficient in programming languages

The notation is convenient in handwriting but inconvenient for typewriters and computer terminals. Many programming languages do not offer a standard subroutine for computing the binomial coefficient, but for example the J programming language uses the exclamation mark: k ! n .

Naive implementations, such as the following snippet in C:

int choose(int n, int k)  {
    return factorial(n) / (factorial(k) * factorial(n - k));
} 

are prone to overflow errors, severely restricting the range of input values. A direct implementation of the first definition works well:

unsigned long long choose(unsigned n, unsigned k) {
    if (k > n)
        return 0;

    if (k > n/2)
        k = n-k; // faster

    long double accum = 1;
    for (unsigned i = 1; i <= k; i++)
         accum = accum * (n-k+i) / i;

    return accum + 0.5; // avoid rounding error
}

See also

References

  1. ^ Nicholas J. Higham, Handbook of writing for the mathematical sciences, SIAM. ISBN 0898714206, p. 25
  2. ^ PlanetMath: binomial coefficient