2019년 2월 28일 목요일

binomial coefficients

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

For example, the fourth power of 1 + x is
and the binomial coefficient  is the coefficient of the x2 term.
Arranging the numbers  in successive rows for  gives a triangular array called Pascal's triangle, satisfying the recurrence relation
The binomial coefficients occur in many areas of mathematics, and especially in combinatorics. The symbol  is usually read as "n choose k" because there are  ways to choose an (unordered) subset of k elements from a fixed set of n elements. For example, there are  ways to choose 2 elements from  namely  and 
The binomial coefficients can be generalized to  for any complex number z and integer k ≥ 0, and many of their properties continue to hold in this more general form.
History and notation[edit]
Andreas von Ettingshausen introduced the notation  in 1826,[1] although the numbers were known centuries earlier (see Pascal's triangle). The earliest known detailed discussion of binomial coefficients is in a tenth-century commentary, by Halayudha, on an ancient Sanskrit text, Pingala's Chandaḥśāstra. In about 1150, the Indian mathematician Bhaskaracharya gave an exposition of binomial coefficients in his book Līlāvatī.[2]
Alternative notations include C(nk)nCknCkCknCnk, and Cn,k in all of which the C stands for combinations or choices. Many calculators use variants of the C notation because they can represent it on a single-line display. In this form the binomial coefficients are easily compared to k-permutations of n, written as P(nk), etc.

Definition and interpretations[edit]

For natural numbers (taken to include 0) n and k, the binomial coefficient  can be defined as the coefficient of the monomial Xk in the expansion of (1 + X)n. The same coefficient also occurs (if k ≤ n) in the binomial formula




()
(valid for any elements x,y of a commutative ring), which explains the name "binomial coefficient".
Another occurrence of this number is in combinatorics, where it gives the number of ways, disregarding order, that k objects can be chosen from among n objects; more formally, the number of k-element subsets (or k-combinations) of an n-element set. This number can be seen as equal to the one of the first definition, independently of any of the formulas below to compute it: if in each of the n factors of the power (1 + X)n one temporarily labels the term X with an index i (running from 1 to n), then each subset of k indices gives after expansion a contribution Xk, and the coefficient of that monomial in the result will be the number of such subsets. This shows in particular that  is a natural number for any natural numbers n and k. There are many other combinatorial interpretations of binomial coefficients (counting problems for which the answer is given by a binomial coefficient expression), for instance the number of words formed of n bits (digits 0 or 1) whose sum is k is given by , while the number of ways to write  where every ai is a nonnegative integer is given by . Most of these interpretations are easily seen to be equivalent to counting k-combinations.

Computing the value of binomial coefficients[edit]

Several methods exist to compute the value of  without actually expanding a binomial power or counting k-combinations.

Recursive formula[edit]

One method uses the recursive, purely additive, formula
with initial/boundary values
The formula follows from considering the set {1,2,3,…,n} and counting separately (a) the k-element groupings that include a particular set element, say "i", in every group (since "i" is already chosen to fill one spot in every group, we need only choose k − 1 from the remaining n − 1) and (b) all the k-groupings that don't include "i"; this enumerates all the possible k-combinations of n elements. It also follows from tracing the contributions to Xk in (1 + X)n−1(1 + X). As there is zero Xn+1 or X−1 in (1 + X)n, one might extend the definition beyond the above boundaries to include  = 0 when either k > n or k < 0. This recursive formula then allows the construction of Pascal's triangle, surrounded by white spaces where the zeros, or the trivial coefficients, would be.

Multiplicative formula[edit]

A more efficient method to compute individual binomial coefficients is given by the formula
where the numerator of the first fraction  is expressed as a falling factorial power. This formula is easiest to understand for the combinatorial interpretation of binomial coefficients. The numerator gives the number of ways to select a sequence of k distinct objects, retaining the order of selection, from a set of n objects. The denominator counts the number of distinct sequences that define the same k-combination when order is disregarded.
Due to the symmetry of the binomial coefficient with regard to k and nk, calculation may be optimised by setting the upper limit of the product above to the smaller of k and nk.

Factorial formula[edit]

Finally, though computationally unsuitable, there is the compact form, often used in proofs and derivations, which makes repeated use of the familiar factorial function:
where n! denotes the factorial of n. This formula follows from the multiplicative formula above by multiplying numerator and denominator by (n − k)!; as a consequence it involves many factors common to numerator and denominator. It is less practical for explicit computation (in the case that k is small and n is large) unless common factors are first cancelled (in particular since factorial values grow very rapidly). The formula does exhibit a symmetry that is less evident from the multiplicative formula (though it is from the definitions)




(1)
which leads to a more efficient multiplicative computational routine. Using the falling factorial notation,

Generalization and connection to the binomial series[edit]

The multiplicative formula allows the definition of binomial coefficients to be extended[3] by replacing n by an arbitrary number α (negative, real, complex) or even an element of any commutative ringin which all positive integers are invertible:
With this definition one has a generalization of the binomial formula (with one of the variables set to 1), which justifies still calling the  binomial coefficients:




(2)
This formula is valid for all complex numbers α and X with |X| < 1. It can also be interpreted as an identity of formal power series in X, where it actually can serve as definition of arbitrary powers of power series with constant coefficient equal to 1; the point is that with this definition all identities hold that one expects for exponentiation, notably
If α is a nonnegative integer n, then all terms with k > n are zero, and the infinite series becomes a finite sum, thereby recovering the binomial formula. However, for other values of α, including negative integers and rational numbers, the series is really infinite.

Pascal's triangle[edit]

1000th row of Pascal's triangle, arranged vertically, with grey-scale representations of decimal digits of the coefficients, right-aligned. The left boundary of the image corresponds roughly to the graph of the logarithm of the binomial coefficients, and illustrates that they form a log-concave sequence.




(3)
which can be used to prove by mathematical induction that  is a natural number for all n and k, a fact that is not immediately obvious from formula (1).
Pascal's rule also gives rise to Pascal's triangle:
0:1
1:11
2:121
3:1331
4:14641
5:15101051
6:1615201561
7:21353521
8:2856705628
Row number n contains the numbers  for k = 0, ..., n. It is constructed by first placing 1s in the outermost positions, and then filling each inner position with the sum of the two numbers directly above. 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

Combinatorics and statistics[edit]

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 elements 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.[4]
  • The Catalan numbers are 
  • The binomial distribution in statistics is 

Binomial coefficients as polynomials[edit]

For any nonnegative integer k, the expression  can be simplified and defined as a polynomial divided by k!:
this presents a polynomial in t with rational coefficients.
As such, it can be evaluated at any real or complex number t to define binomial coefficients with such first arguments. These "generalized binomial coefficients" appear in Newton's generalized binomial theorem.
For each k, the polynomial  can be characterized as the unique degree k polynomial p(t) satisfying p(0) = p(1) = ... = p(k − 1) = 0 and p(k) = 1.
Its coefficients are expressible in terms of Stirling numbers of the first kind:
The derivative of  can be calculated by logarithmic differentiation:

Binomial coefficients as a basis for the space of polynomials[edit]

Over any field of characteristic 0 (that is, any field that contains the rational numbers), each polynomial p(t) of degree at most d is uniquely expressible as a linear combination  of binomial coefficients. The coefficient ak is the kth difference of the sequence p(0), p(1), …, p(k). Explicitly,[5]




(4)

Integer-valued polynomials[edit]

Each polynomial  is integer-valued: it has an integer value at all integer inputs . (One way to prove this is by induction on k, using Pascal's identity.) Therefore, any integer linear combination of binomial coefficient polynomials is integer-valued too. Conversely, (4) shows that any integer-valued polynomial is an integer linear combination of these binomial coefficient polynomials. More generally, for any subring R of a characteristic 0 field K, a polynomial in K[t] takes values in R at all integers if and only if it is an R-linear combination of binomial coefficient polynomials.

Example[edit]

The integer-valued polynomial 3t(3t + 1)/2 can be rewritten as

Identities involving binomial coefficients[edit]

The factorial formula facilitates relating nearby binomial coefficients. For instance, if k is a positive integer and n is arbitrary, then




(5)
and, with a little more work,
Moreover, the following may be useful:
For constant n, we have the following recurrence:

Sums of the binomial coefficients[edit]

The formula




(∗∗)
says the elements in the nth row of Pascal's triangle always add up to 2 raised to the nth power. This is obtained from the binomial theorem () by setting x = 1 and y = 1. The formula also has a natural combinatorial interpretation: the left side sums the number of subsets of {1,...,n} of sizes k = 0,1,...,n, giving the total number of subsets. (That is, the left side counts the power set of {1,...,n}.) However, these subsets can also be generated by successively choosing or excluding each element 1,...,n; the n independent binary choices (bit-strings) allow a total of  choices. The left and right sides are two ways to count the same collection of subsets, so they are equal.
The formulas




(6)
and  follow from (2) after differentiating with respect to x (twice in the latter) and then substituting x = 1.
The Chu–Vandermonde identity, which holds for any complex-values m and n and any non-negative integer k, is




(7)
and can be found by examination of the coefficient of  in the expansion of (1 + x)m (1 + x)n − m = (1 + x)n using equation (2). When m = 1, equation (7) reduces to equation (3). In the special case n = 2mk = m, using (1), the expansion (7) becomes (as seen in Pascal's triangle at right)
Pascal's triangle, rows 0 through 7. Equation 8for m = 3 is illustrated in rows 3 and 6 as 




(8)
where the term on the right side is a central binomial coefficient.
Another form of the Chu–Vandermonde identity, which applies for any integers jk, and n satisfying 0 ≤ j ≤ k ≤ n, is




(9)
The proof is similar, but uses the binomial series expansion (2) with negative integer exponents. When j = k, equation (9) gives the hockey-stick identity
Let F(n) denote the n-th Fibonacci number. Then
This can be proved by induction using (3) or by Zeckendorf's representation. A combinatorial proof is given below.

Multisections of sums[edit]

For integers s and t such that  series multisection gives the following identity for the sum of binomial coefficients:
For small s, these series have particularly nice forms; for example,[6]

Partial sums[edit]

Although there is no closed formula for partial sums
of binomial coefficients,[7] one can again use (3) and induction to show that for k = 0, ..., n − 1,
,
with special case[8]
for n > 0. This latter result is also a special case of the result from the theory of finite differences that for any polynomial P(x) of degree less than n,[9]
Differentiating (2k times and setting x = −1 yields this for , when 0 ≤ k < n, and the general case follows by taking linear combinations of these.
When P(x) is of degree less than or equal to n,




(10)
where  is the coefficient of degree n in P(x).
More generally for (10),
where m and d are complex numbers. This follows immediately applying (10) to the polynomial Q(x):=P(m + dx) instead of P(x), and observing that Q(x) has still degree less than or equal to n, and that its coefficient of degree n is dnan.
The series  is convergent for k ≥ 2. This formula is used in the analysis of the German tank problem. It follows from  which is proved by induction on M.

Identities with combinatorial proofs[edit]

Many identities involving binomial coefficients can be proved by combinatorial means. For example, for nonnegative integers , the identity
(which reduces to (6) when q = 1) can be given a double counting proof, as follows. The left side counts the number of ways of selecting a subset of [n] = {1, 2, …, n} with at least q elements, and marking q elements among those selected. The right side counts the same thing, because there are  ways of choosing a set of q elements to mark, and  to choose which of the remaining elements of [n] also belong to the subset.
In Pascal's identity
both sides count the number of k-element subsets of [n]: the two terms on the right side group them into those that contain element n and those that do not.
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 ksquares from among the first n and  squares from the remaining n squares; any k from 0 to n will work. This gives
Now apply (5) to get the result.
If one denotes by F(i) the sequence of Fibonacci numbers, indexed so that F(0) = F(1) = 1, then the identity
has the following combinatorial proof.[10] One may show by induction that F(n) counts the number of ways that a n × 1 strip of squares may be covered by 2 × 1 and 1 × 1 tiles. On the other hand, if such a tiling uses exactly k of the 2 × 1 tiles, then it uses n − 2k of the 1 × 1 tiles, and so uses n − k tiles total. There are  ways to order these tiles, and so summing this coefficient over all possible values of k gives the identity.

Sum of coefficients row[edit]

The number of k-combinations for all k, is the sum of the nth row (counting from 0) of the binomial coefficients. These combinations are enumerated by the 1 digits of the set of base 2 numbers counting from 0 to , where each digit position is an item from the set of n.

Dixon's identity[edit]

or, more generally,
where ab, and c are non-negative integers.

Continuous identities[edit]

Certain trigonometric integrals have values expressible in terms of binomial coefficients:
For 
These can be proved by using Euler's formula to convert trigonometric functions to complex exponentials, expanding using the binomial theorem, and integrating term by term.

Generating functions[edit]

Ordinary generating functions[edit]

For a fixed n, the ordinary generating function of the sequence  is
For a fixed k, the ordinary generating function of the sequence  is
The bivariate generating function of the binomial coefficients is
Another, symmetric, bivariate generating function of the binomial coefficients is

Exponential generating function[edit]

A symmetric exponential bivariate generating function of the binomial coefficients is:

Divisibility properties[edit]

In 1852, Kummer proved that if m and n are nonnegative integers and p is a prime number, then the largest power of p dividing  equals pc, where c is the number of carries when m and n are added in base p. Equivalently, the exponent of a prime p in  equals the number of nonnegative integers j such that the fractional part of k/pj is greater than the fractional part of n/pj. It can be deduced from this that  is divisible by n/gcd(n,k). In particular therefore it follows that p divides  for all positive integers r and s such that s < pr. However this is not true of higher powers of p: for example 9 does not divide .
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.
Binomial coefficients have divisibility properties related to least common multiples of consecutive integers. For example:[11]
 divides .
 is a multiple of .
Another fact: An integer n ≥ 2 is prime if and only if all the intermediate binomial coefficients
are divisible by n.
Proof: When p is prime, p divides
 for all 0 < k < p
because  is a natural number and p divides the numerator but not the denominator. When n is composite, let p be the smallest prime factor of n and let k = n/p. Then 0 < p < n and
otherwise the numerator k(n − 1)(n − 2)×...×(n − p + 1) has to be divisible by n = k×p, this can only be the case when (n − 1)(n − 2)×...×(n − p + 1) is divisible by p. But n is divisible by p, so p does not divide n − 1, n − 2, ..., n − p + 1 and because p is prime, we know that p does not divide (n − 1)(n − 2)×...×(n − p + 1) and so the numerator cannot be divisible by n.

Bounds and asymptotic formulas[edit]

The following bounds for  hold for all values of n and k such that 1 ≤ k ≤ n:
.
The first inequality follows from the fact that  and each of these  terms in this product is . A similar argument can be made to show . The final strict inequality follows from multiplying  by the Taylor series for .
From the divisibility properties we can infer that
,
where both equalities can be achieved.[11]

Both n and k large[edit]

Stirling's approximation yields the following approximation, when  are sufficiently large:
Because the inequality forms of Stirling's formula also bound the factorials, slight variants on the above asymptotic approximation give exact bounds.
In particular, when is sufficiently large:
and 
and, in general, for m ≥ 2 and n ≥ 1,[why?]

n much larger than k[edit]

When  is large and  is much smaller than , one can also write
This gives the following analogue of Stirling's formula:
If more precision is desired, one can approximate  with an integral, obtaining
For  and , and these approximations yield 12.312 and 12.133 respectively.
If n is large and k is linear in n, various precise asymptotic estimates exist for the binomial coefficient . For example, if  then[12]
where d = n − 2k.

Sums of binomial coefficients[edit]

A simple and rough upper bound for the sum of binomial coefficients can be obtained using the binomial theorem:
For  and  both much larger than 1, Stirling's approximation also yields the following asymptotic approximation:
where  is the binary entropy of . More precisely, for all integers  with , we can estimate the sum of the first  binomial coefficients as follows:[13]

Generalized binomial coefficients[edit]

The infinite product formula for the Gamma function also gives an expression for binomial coefficients
which yields the asymptotic formulas
as .
This asymptotic behaviour is contained in the approximation
as well. (Here  is the k-th harmonic number and  is the Euler–Mascheroni constant.)
Further, the asymptotic formula
hold true, whenever  and  for some complex number .

Generalizations[edit]

Generalization to multinomials[edit]

Binomial coefficients can be generalized to multinomial coefficients 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
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 those of binomial coefficients, for example the recurrence relation:
and symmetry:
where  is a permutation of (1,2,...,r).

Taylor series[edit]

Using Stirling numbers of the first kind the series expansion around any arbitrarily chosen point  is

Binomial coefficient with n = 1/2[edit]

The definition of the binomial coefficients can be extended to the case where  is real and  is integer.
In particular, the following identity holds for any non-negative integer  :
This shows up when expanding  into a power series using the Newton binomial series :

Identity for the product of binomial coefficients[edit]

One can express the product of binomial coefficients as a linear combination of binomial coefficients:
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 to get a new labelled combinatorial object of weight m + n − k. (That is, to separate the labels into three portions to apply 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.

Partial fraction decomposition[edit]

The partial fraction decomposition of the reciprocal is given by

Newton's binomial series[edit]

Newton's binomial series, named after Sir Isaac Newton, is a generalization of the binomial theorem to infinite series:
The identity can be obtained by showing that both sides satisfy the differential equation (1 + zf'(z) = α f(z).
The radius of convergence of this series is 1. An alternative expression is
where the identity
is applied.

Multiset (rising) binomial coefficient[edit]

Binomial coefficients count subsets of prescribed size from a given set. A related combinatorial problem is to count multisets of prescribed size with elements drawn from a given set, that is, to count the number of ways to select a certain number of elements from a given set with the possibility of selecting the same element repeatedly. The resulting numbers are called multiset coefficients;[14] the number of ways to "multichoose" (i.e., choose with replacement) k items from an n element set is denoted .
To avoid ambiguity and confusion with n's main denotation in this article,
let f = n = r + (k – 1) and r = f – (k – 1).
Multiset coefficients may be expressed in terms of binomial coefficients by the rule
One possible alternative characterization of this identity is as follows: We may define the falling factorial as
,
and the corresponding rising factorial as
;
so, for example,
.
Then the binomial coefficients may be written as
,
while the corresponding multiset coefficient is defined by replacing the falling with the rising factorial:
.

Generalization to negative integers[edit]

For any n,
In particular, binomial coefficients evaluated at negative integers are given by signed multiset coefficients. In the special case , this reduces to 
For example, if n = -4 and k = 7, then r = 4 and f = 10:

Two real or complex valued arguments[edit]

The binomial coefficient is generalized to two real or complex valued arguments using the gamma function or beta function via
This definition inherits these following additional properties from :
moreover,
The resulting function has been little-studied, apparently first being graphed in (Fowler 1996). Notably, many binomial identities fail:  but  for n positive (so  negative). The behavior is quite complex, and markedly different in various octants (that is, with respect to the x and y axes and the line ), with the behavior for negative x having singularities at negative integer values and a checkerboard of positive and negative regions:
  • in the octant  it is a smoothly interpolated form of the usual binomial, with a ridge ("Pascal's ridge").
  • in the octant  and in the quadrant  the function is close to zero.
  • in the quadrant  the function is alternatingly very large positive and negative on the parallelograms with vertices
  • in the octant  the behavior is again alternatingly very large positive and negative, but on a square grid.
  • in the octant  it is close to zero, except for near the singularities.

Generalization to q-series[edit]

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

Generalization to infinite cardinals[edit]

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[edit]

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 both the APL programming language and the (related) J programming language use the exclamation mark: k ! n .
Naive implementations of the factorial formula, such as the following snippet in Python:
from math import factorial
def binomialCoefficient(n, k):
    return factorial(n) // (factorial(k) * factorial(n - k))
are very slow and are useless for calculating factorials of very high numbers (in languages such as C or Java they suffer from overflow errors because of this reason). A direct implementation of the multiplicative formula works well:
def binomialCoefficient(n, k):
    if k < 0 or k > n:
        return 0
    if k == 0 or k == n:
        return 1
    k = min(k, n - k) # take advantage of symmetry
    c = 1
    for i in range(k):
        c = c * (n - i) / (i + 1)
    return c
(In Python, range(k) produces a list from 0 to k–1.)
Pascal's rule provides a recursive definition which can also be implemented in Python, although it is less efficient:
def binomialCoefficient(n, k):
    if k < 0 or k > n:
        return 0
    if k > n - k: # take advantage of symmetry
        k = n - k
    if k == 0 or n <= 1:
     return 1
    return binomialCoefficient(n-1, k) + binomialCoefficient(n-1, k-1)
The example mentioned above can be also written in functional style. The following Scheme example uses the recursive definition
Rational arithmetic can be easily avoided using integer division
The following implementation uses all these ideas
(define (binomial n k)
;; Helper function to compute C(n,k) via forward recursion
  (define (binomial-iter n k i prev)
    (if (>= i k)
      prev
     (binomial-iter n k (+ i 1) (/ (* (- n i) prev) (+ i 1)))))
;; Use symmetry property C(n,k)=C(n, n-k)
  (if (< k (-  n k))
    (binomial-iter n k 0 1)
    (binomial-iter n (- n k) 0 1)))
When computing  in a language with fixed-length integers, the multiplication by  may overflow even when the result would fit. The overflow can be avoided by dividing first and fixing the result using the remainder:
Implementation in the C language:
#include <limits.h>

unsigned long binomial(unsigned long n, unsigned long k) {
  unsigned long c = 1, i;
  
  if (k > n-k) // take advantage of symmetry
    k = n-k;
  
  for (i = 1; i <= k; i++, n--) {
    if (c/i > UINT_MAX/n) // return 0 on overflow
      return 0;
      
    c = c / i * n + c % i * n / i;  // split c * n / i into (c / i * i + c % i) * n / i
  }
  
  return c;
}
Another way to compute the binomial coefficient when using large numbers is to recognize that
where  denotes the natural logarithm of the gamma function at . It is a special function that is easily computed and is standard in some programming languages such as using log_gamma in MaximaLogGamma in Mathematicagammaln in MATLAB and Python's SciPy module, lngamma in PARI/GP or lgamma in C, R,[15] and Julia. Roundoff error may cause the returned value to not be an integer.
Wikipedia

댓글 없음: