Fact-checked by Grok 2 weeks ago

Binomial coefficient

In , the binomial coefficient, denoted \binom{n}{k} or {}_nC_k and read as "n choose k," is the number of distinct ways to select k unordered elements from a set of n elements, without regard to the order of selection. This combinatorial quantity is given by the formula \binom{n}{k} = \frac{n!}{k!(n-k)!} for non-negative integers n and k with $0 \leq k \leq n, where ! denotes the function. Binomial coefficients form the entries of Pascal's triangle, where each number in row n (starting from row 0) is \binom{n}{k} for k = 0 to n, and adjacent entries satisfy the recurrence relation \binom{n+1}{k} = \binom{n}{k} + \binom{n}{k-1}. They are central to the binomial theorem, which states that for any non-negative integer n and variables x and y, (x + y)^n = \sum_{k=0}^{n} \binom{n}{k} x^{n-k} y^k. Key properties include symmetry (\binom{n}{k} = \binom{n}{n-k}), boundary values (\binom{n}{0} = \binom{n}{n} = 1), and the fact that the sum of the coefficients in row n equals $2^n, representing the total number of subsets of an n-element set. These coefficients appear extensively in , probability, , and other fields, underpinning concepts like in problems and expansions in generating functions.

Definition and Notation

Basic Definition

The binomial coefficient, denoted \binom{n}{k}, represents the number of ways to choose k items from a set of n distinct items without regard to the order of selection, where n and k are non-negative integers with n \geq k \geq 0. This quantity is also known as a in . The explicit formula for the binomial coefficient is given by \binom{n}{k} = \frac{n!}{k!(n-k)!}, where n! denotes the of n, defined as the product of all positive integers up to n (with $0! = 1). By convention, \binom{n}{k} = 0 if k > n or k < 0, reflecting that it is impossible to select more items than available or a negative number of items. In particular, \binom{n}{0} = 1 and \binom{n}{n} = 1, as there is exactly one way to choose nothing or everything from the set. Binomial coefficients arise naturally in the binomial theorem, which states that for any non-negative integer n and variables x and y, (x + y)^n = \sum_{k=0}^n \binom{n}{k} x^{n-k} y^k. This expansion motivates the coefficients as the multipliers that count the distinct terms formed by choosing k factors of y (and thus n-k factors of x) in the product. A key property is the symmetry \binom{n}{k} = \binom{n}{n-k}, which follows directly from the factorial formula and underscores the equivalence of choosing k items or the complementary n-k items to leave behind. Binomial coefficients form the entries of Pascal's triangle, a triangular array where each interior value is the sum of the two values directly above it.

Historical Development and Notation

The concept of binomial coefficients traces its origins to ancient Indian mathematics, particularly in the work of Pingala around the 3rd century BCE. In his treatise Chandaḥśāstra on Sanskrit prosody, Pingala explored combinatorial patterns for poetic meters using recursive algorithms that effectively computed what are now recognized as binomial coefficients, marking an early systematic approach to such enumerations. Independently, Chinese mathematicians also developed these ideas. Around the 11th century, Jia Xian described a triangular array of binomial coefficients in his work Shi Suo Suan Fa, using it to extract roots via binomial expansions, which was later popularized by Yang Hui in the 13th century. This combinatorial foundation was further developed in the Islamic world during the medieval period. In the 10th century, the Persian mathematician Al-Karaji provided one of the earliest explicit formulations of binomial coefficients in his algebraic texts, including a description of their triangular arrangement akin to Pascal's triangle, and applied mathematical induction to related expansions. Building on this, Omar Khayyam in the 11th century utilized binomial expansions to solve cubic equations, demonstrating practical applications of these coefficients in algebraic geometry. European engagement with binomial coefficients gained prominence in the 17th century through Blaise Pascal's Traité du triangle arithmétique (1654), where he systematically analyzed the "arithmetic triangle" of coefficients, connecting them to probability problems in games of chance and establishing their role in early statistical reasoning. The binomial theorem, which expresses powers of binomials using these coefficients, had been generalized by Isaac Newton around 1665 for non-integer exponents, though its integer form built on prior traditions. The modern notation for binomial coefficients, \binom{n}{k}, was introduced by Andreas von Ettingshausen in his 1826 work Die Combinatorische Analysis, providing a compact symbolic representation that facilitated further analysis. Prior to this, expressions relied on verbal descriptions or product forms; by the 18th century, Leonhard Euler and contemporaries had popularized factorial-based representations, such as \frac{n!}{k!(n-k)!}, to denote these coefficients explicitly in analytic contexts. Alternative notations, including C(n,k) and the phrase "n choose k", emerged later in the 19th and 20th centuries for combinatorial literature.

Computation Methods

Factorial Formula

The factorial formula provides the classical explicit expression for the binomial coefficient \binom{n}{k} when n and k are nonnegative integers with $0 \leq k \leq n. It is given by \binom{n}{k} = \frac{n!}{k!(n-k)!}, where the factorial n! denotes the product of all positive integers from 1 to n, specifically n! = n \times (n-1) \times \cdots \times 1, and by convention $0! = 1. For example, \binom{5}{2} = \frac{5!}{2!(5-2)!} = \frac{120}{2 \times 6} = 10. This formula arises as a derivation from the binomial theorem, which states that (x + y)^n = \sum_{k=0}^{n} \binom{n}{k} x^{n-k} y^k; equating coefficients of corresponding powers of x and y yields the factorial expression after expanding both sides. By convention, \binom{n}{k} = 0 when k < 0, k > n, or n < 0 with k \geq 0, though extensions to negative upper indices exist via analytic continuation (discussed in later sections). Computing binomial coefficients via the factorial formula poses challenges for large n, as factorials grow rapidly and can cause numerical overflow in fixed-precision arithmetic, such as in programming implementations where intermediate values exceed the representable range (e.g., beyond 64-bit integers for n > 20).

Multiplicative Formula

The multiplicative formula expresses the binomial coefficient \binom{n}{k} as a finite product of k fractional terms, which is particularly useful for direct computation when k \leq n/2 (otherwise, replace k with n-k to minimize the product length). This formula is given by \binom{n}{k} = \prod_{i=1}^k \frac{n-k+i}{i}, where the product starts from 1 and each term \frac{n-k+i}{i} is computed sequentially. This representation derives from the falling factorial form \binom{n}{k} = \frac{n^{\underline{k}}}{k!}, where n^{\underline{k}} = n(n-1)\cdots(n-k+1), avoiding the need to compute full factorials n!, k!, and (n-k)!. To compute this efficiently in practice, initialize the result to 1 and iteratively update it by multiplying by the numerator n-k+i and then dividing by the denominator i at each step i = 1 to k. To ensure exact integer arithmetic and prevent from large intermediate values, reduce the fraction before each multiplication by dividing both the current result and the upcoming numerator by their (gcd) with the denominator, or perform the division immediately if it divides evenly. This step-by-step process keeps intermediate results bounded closely to the final value, unlike the factorial formula which requires handling much larger numbers (e.g., n! can be exponentially bigger than \binom{n}{k}). For example, to compute \binom{10}{3}, set k=3 (since $3 \leq 5) and proceed as follows:
  • Start with result = 1.
  • For i=1: multiply by \frac{10-3+1}{1} = \frac{8}{1}, so result = 1 \times 8 / 1 = 8.
  • For i=2: multiply by \frac{10-3+2}{2} = \frac{9}{2}. First, 8 \times 9 = 72, then 72 / 2 = 36 (exact division).
  • For i=3: multiply by \frac{10-3+3}{3} = \frac{10}{3}. First, 36 \times 10 = 360, then 360 / 3 = 120 (exact division).
The result is 120, obtained without computing any factorial larger than necessary and with intermediates (8, 36, 120) far smaller than those in the factorial approach (e.g., 10! = 3,628,800). This method yields exact integers for integer inputs $0 \leq k \leq n and is foundational for algorithmic implementations in computer science.

Recursive Formulas

One fundamental recursive relation for binomial coefficients is Pascal's identity, which states that for non-negative integers n and k with $1 \leq k \leq n-1, \binom{n}{k} = \binom{n-1}{k} + \binom{n-1}{k-1}. This identity arises from the additive structure of combinations and holds with base cases \binom{n}{0} = 1 and \binom{n}{n} = 1 for all n \geq 0, as well as \binom{n}{k} = 0 for k > n or k < 0. The recursive identity enables a dynamic programming approach to compute \binom{n}{k} by constructing a table where each entry is filled row by row, starting from the base cases along the edges. In this method, the table B is built iteratively for i from 0 to n and j from 0 to \min(i, k), using the recurrence B = B[i-1] + B[i-1][j-1] after initializing B{{grok:render&&&type=render_inline_citation&&&citation_id=0&&&citation_type=wikipedia}} = 1 and B = 1. This avoids redundant computations by storing intermediate results, making it suitable for calculating multiple coefficients up to \binom{n}{k}. For illustration, consider computing \binom{4}{2} via the recursion tree: it branches to \binom{3}{2} + \binom{3}{1}, where \binom{3}{2} further splits into \binom{2}{2} + \binom{2}{1} = 1 + 2 = 3 and \binom{3}{1} into \binom{2}{1} + \binom{2}{0} = 2 + 1 = 3, yielding $3 + 3 = 6. The additive buildup in such trees highlights how larger coefficients emerge from summing smaller ones, though naive recursion without memoization leads to exponential time due to overlapping subproblems. The table-based dynamic programming method achieves a time complexity of O(nk), as it performs a constant-time addition for each of the O(nk) entries, with space usage of O(nk) or optimizable to O(k) by retaining only the previous row. Recursive formulations extend beyond integers through the binomial series, where for real \alpha, the expansion (1 + x)^\alpha = \sum_{k=0}^\infty \binom{\alpha}{k} x^k (with \binom{\alpha}{k} = \frac{\alpha (\alpha-1) \cdots (\alpha-k+1)}{k!}) generalizes the identity for convergent |x| < 1, connecting binomial coefficients to power series analysis.

Visual and Combinatorial Interpretations

Pascal's Triangle

Pascal's triangle is a triangular array where each entry is a binomial coefficient, constructed by starting with row 0 containing a single 1, and each subsequent entry in row n (for n \geq 1) being the sum of the two entries directly above it from row n-1. This recursive construction yields the binomial coefficients \binom{n}{k} as the entry in row n, position k (with $0 \leq k \leq n), aligning the rows with the coefficients in the expansion of (x + y)^n. The first few rows of Pascal's triangle illustrate this structure:
Row nEntries \binom{n}{k} for k = 0 to n
01
11 1
21 2 1
31 3 3 1
41 4 6 4 1
This array exhibits symmetry, where \binom{n}{k} = \binom{n}{n-k}, reflecting the balanced nature of binomial expansions. The sum of entries in row n equals $2^n, providing a preview of the through cumulative row sums that build powers of 2. Additionally, the sums along the shallow diagonals produce the , as the nth diagonal sums to the (n+1)th F_{n+1}, following the recursive relation shared with binomial coefficients. Blaise Pascal formalized the triangle in his 1665 treatise Traité du triangle arithmétique, where he explored its properties in the context of probability calculations and games of chance, such as dividing stakes in interrupted games, linking the entries to combinatorial counts. One practical application is counting paths in a grid: the number of distinct paths from the top-left to the bottom-right of an m \times n grid, moving only right or down, equals \binom{m+n}{m}, directly corresponding to an entry in the triangle.

Combinatorial Interpretations

The binomial coefficient \binom{n}{k} fundamentally counts the number of ways to select a subset of k elements from a set of n distinct elements, without regard to order. This interpretation arises in discrete mathematics as the cardinality of the collection of all k-element subsets of an n-element set. For instance, \binom{n}{k} represents the number of ways to choose k members for a committee from n eligible individuals. Equivalently, it counts the number of binary sequences of length n with exactly k ones (and n-k zeros), corresponding to the positions selected for the ones. In the context of sampling without replacement from a finite population, the binomial coefficient appears in the hypergeometric distribution, where \binom{K}{k} denotes the number of ways to draw k successes from K available successes in the population. Combinatorial identities involving binomial coefficients, such as Vandermonde's convolution \sum_{i=0}^{k} \binom{m}{i} \binom{n}{k-i} = \binom{m+n}{k}, can be proved via bijections that equate different ways of selecting subsets. Specifically, both sides count the number of k-subsets from a set of m+n elements partitioned into two disjoint subsets of sizes m and n, by considering the possible intersections with each partition. An extension of this subset-counting view interprets \binom{n}{k} as the number of lattice paths from the point (0,0) to (n-k, k) in the plane, using exactly n-k rightward steps (1,0) and k upward steps (0,1). Each such path corresponds to a sequence of n steps where the positions of the upward moves are chosen in \binom{n}{k} ways.

Statistical Applications

Binomial coefficients play a central role in the binomial distribution, which models the number of successes in a fixed number of independent Bernoulli trials, each with success probability p. The probability mass function for a binomial random variable X \sim \operatorname{Bin}(n, p) is given by P(X = k) = \binom{n}{k} p^k (1-p)^{n-k}, for k = 0, 1, \dots, n, where \binom{n}{k} enumerates the ways to choose k successes out of n trials. This distribution arises directly from sequences of Bernoulli trials, where each trial is a random experiment with two outcomes: success (probability p) or failure (probability $1-p). The binomial coefficient \binom{n}{k} counts the number of favorable sequences with exactly k successes, scaled by the probability of each such sequence. A classic example is repeated coin flips, where a fair coin (p = 0.5) models heads as success. The probability of exactly k heads in n flips is P(X = k) = \binom{n}{k} (0.5)^n, such as \binom{10}{5} / 2^{10} \approx 0.246 for 5 heads in 10 flips. The expected value and variance of X derive from sums involving binomial coefficients. Specifically, E[X] = np, obtained by E[X] = \sum_{k=0}^n k \binom{n}{k} p^k (1-p)^{n-k} = np \sum_{k=1}^n \binom{n-1}{k-1} p^{k-1} (1-p)^{n-k} = np (p + 1-p)^{n-1} = np, using the identity k \binom{n}{k} = n \binom{n-1}{k-1}. The variance is \operatorname{Var}(X) = np(1-p), derived similarly via E[X^2] - (E[X])^2. In hypothesis testing, binomial coefficients underpin the exact binomial test for a population proportion p, testing H_0: p = p_0 against alternatives by computing the p-value as the sum of probabilities under H_0 for outcomes as extreme as or more extreme than observed. This is particularly useful for small samples or when np_0 or n(1-p_0) is below 5, avoiding normal approximations. For confidence intervals on binomial proportions, methods like the Wilson score interval adjust the observed proportion \hat{p} = k/n using \binom{n}{k}-based likelihoods to achieve better coverage than the naive Wald interval \hat{p} \pm z \sqrt{\hat{p}(1-\hat{p})/n}. The Wilson interval is \tilde{p} \pm z \sqrt{\tilde{p}(1-\tilde{p})/(n + z^2)}, where \tilde{p} = (k + z^2/2)/(n + z^2), recommended for most practical settings. For large n, the de Moivre-Laplace theorem provides a normal approximation to the binomial distribution: (X - np)/\sqrt{np(1-p)} \approx N(0,1), enabling asymptotic inference like approximate tests and intervals when exact computations with binomial coefficients are infeasible. This central limit theorem variant facilitates analysis of large-scale Bernoulli processes, such as quality control or polling.

Polynomial and Algebraic Properties

Binomial Coefficients as Polynomials

The binomial coefficient can be generalized to non-integer arguments by defining \binom{x}{k} = \frac{x(x-1)\cdots(x-k+1)}{k!} for a fixed positive integer k and variable x, which expresses it explicitly as a polynomial in x of degree exactly k. This form arises from the falling factorial (x)_k = x(x-1)\cdots(x-k+1) divided by the constant k!, preserving the combinatorial meaning when x is a non-negative integer greater than or equal to k. For instance, when k=2, the expression simplifies to \binom{x}{2} = \frac{x(x-1)}{2} = \frac{x^2 - x}{2}, a quadratic polynomial whose coefficients are rational numbers. In general, expanding \binom{x}{k} yields a monic polynomial with leading term \frac{x^k}{k!} and lower-degree terms involving , but the product form highlights its direct connection to successive decrements in the argument. When x = n for a non-negative integer n \geq k, this polynomial evaluates to the standard binomial coefficient \binom{n}{k}, recovering the combinatorial count of ways to choose k elements from n. This integer evaluation underscores the polynomial's role as an interpolating function for the discrete binomial coefficients. A key property in discrete calculus is that the forward difference operator \Delta, defined by \Delta f(x) = f(x+1) - f(x), satisfies \Delta \binom{x}{k} = \binom{x}{k-1} for k \geq 1, with \Delta \binom{x}{0} = 0. This relation follows from the identity for falling factorials, \Delta (x)_k = k (x)_{k-1}, since dividing by the constants k! and (k-1)! yields the result; it parallels the differentiation rule \frac{d}{dx} x^k = k x^{k-1} in continuous calculus and facilitates summation by parts and discrete integration techniques.

Integer-Valued Polynomials

Integer-valued polynomials are those with rational coefficients that map the integers to integers, forming the ring \operatorname{Int}(\mathbb{Z}) = \{ f \in \mathbb{Q} \mid f(\mathbb{Z}) \subseteq \mathbb{Z} \}. Every such polynomial f of degree n can be uniquely expressed as f(x) = \sum_{k=0}^n c_k \binom{x}{k}, where the c_k are integers and \binom{x}{k} = \frac{x(x-1)\cdots(x-k+1)}{k!} denotes the binomial polynomials. This representation highlights that the binomial polynomials \{\binom{x}{k}\}_{k=0}^\infty form a \mathbb{Z}-basis for \operatorname{Int}(\mathbb{Z}) as a free \mathbb{Z}-module. The uniqueness of this expansion follows from the fact that the binomial basis provides a complete and linearly independent spanning set over \mathbb{Z}, analogous to a in the context of this module structure. For instance, the monomial x^2 admits the decomposition x^2 = 2\binom{x}{2} + \binom{x}{1}, demonstrating how higher powers are rewritten in this non-monomial basis while preserving integer values on \mathbb{Z}. This contrasts with the standard power basis, as the coefficients c_k ensure integrality without requiring integer coefficients in the monomial form. These polynomials find significant applications in finite differences and interpolation. The forward difference operator \Delta f(x) = f(x+1) - f(x) satisfies \Delta \binom{x}{k} = \binom{x}{k-1}, leading to the Newton series expansion f(x) = \sum_{k=0}^\infty \Delta^k f(0) \binom{x}{k}, which converges for polynomials and provides a discrete analog of . In interpolation, given integer values at n+1 consecutive points, there exists a unique integer-valued polynomial of degree at most n interpolating those points, expressible via the binomial basis.

Basis for Polynomial Spaces

The vector space of polynomials of degree at most n with rational coefficients, denoted \mathbb{P}_n(\mathbb{Q}), has dimension n+1. The set \left\{ \binom{x}{0}, \binom{x}{1}, \dots, \binom{x}{n} \right\}, where \binom{x}{k} = \frac{x(x-1)\cdots(x-k+1)}{k!}, forms a basis for this space. The polynomials in this set are linearly independent. One proof relies on their leading coefficients: the polynomial \binom{x}{k} has degree k and leading coefficient $1/k!, which is nonzero for each k; thus, in any linear combination \sum_{k=0}^n c_k \binom{x}{k} = 0, the coefficients of the highest-degree terms force c_n = 0, and proceeding inductively yields all c_k = 0. Alternatively, linear independence follows from evaluating the polynomials at the points x = 0, 1, \dots, n: the resulting (n+1) \times (n+1) matrix has entries M_{ij} = \binom{i}{j} (with rows indexed by i and columns by j), known as the , which is invertible as it is lower triangular with 1s on the diagonal. The change of basis from the standard monomial basis \{1, x, x^2, \dots, x^n\} to the binomial basis is given explicitly by the relation x^m = \sum_{k=0}^m S(m,k) \, k! \, \binom{x}{k} for $0 \leq m \leq n, where S(m,k) are the Stirling numbers of the second kind, counting the number of ways to partition m elements into k nonempty subsets. The corresponding transformation matrix is lower triangular with entries involving these scaled Stirling numbers. This basis finds key applications in numerical analysis, particularly in Newton interpolation for data on equidistant points. For points x_i = x_0 + i h (i = 0, \dots, n), the interpolating polynomial takes the forward difference form p(x_0 + s h) = \sum_{k=0}^n \binom{s}{k} \Delta^k f(x_0), where \Delta is the forward difference operator and s = (x - x_0)/h; the coefficients \Delta^k f(x_0) are readily computed from function values, making the binomial basis computationally stable for such approximations.

Key Identities

Sums and Basic Identities

One of the fundamental identities involving binomial coefficients arises from the binomial theorem, which expands (1 + x)^n = \sum_{k=0}^n \binom{n}{k} x^k. Substituting x = 1 yields the total sum \sum_{k=0}^n \binom{n}{k} = 2^n, representing the total number of subsets of an n-element set. The alternating sum follows similarly by substituting x = -1: (1 - 1)^n = \sum_{k=0}^n (-1)^k \binom{n}{k}. For n > 0, this equals 0, highlighting the balance between even and odd indexed terms. Partial sums \sum_{k=0}^m \binom{n}{k} for $0 < m < n generally lack a closed form but appear in applications like cumulative distribution functions. A key identity for related partial sums is the hockey-stick identity: \sum_{k=r}^n \binom{k}{r} = \binom{n+1}{r+1}. This is proved algebraically via Pascal's recurrence \binom{k}{r} = \binom{k+1}{r+1} - \binom{k}{r+1}, which telescopes the sum to \binom{n+1}{r+1} - \binom{r}{r+1} = \binom{n+1}{r+1}. Sums restricted to residue classes modulo m partition the total sum into m parts. The roots of unity filter provides the formula \sum_{\substack{0 \leq k \leq n \\ k \equiv r \pmod{m}}} \binom{n}{k} = \frac{1}{m} \sum_{j=0}^{m-1} \omega^{-j r} (1 + \omega^j)^n, where \omega = e^{2\pi i / m} is a primitive m-th root of unity. This derives from applying the binomial theorem to each (1 + \omega^j)^n and averaging with the orthogonality of roots of unity.

Combinatorial Proofs

Combinatorial proofs establish binomial identities by demonstrating that both sides count the same combinatorial object in different ways, often through double counting or bijections, thereby avoiding algebraic manipulation. A fundamental example is the , which states that \sum_{k=0}^{r} \binom{m}{k} \binom{n}{r-k} = \binom{m+n}{r} for non-negative integers m, n, and r. To prove this combinatorially, consider selecting a committee of r people from a group consisting of m men and n women. The left side counts the ways by fixing k as the number of men selected (for k = 0 to r), choosing those k men from m and the remaining r-k women from n. The right side counts the total ways to choose r from the combined m+n people, establishing equality via this double counting. Another basic identity is the sum of the binomial coefficients in a row of Pascal's triangle: \sum_{k=0}^{n} \binom{n}{k} = 2^n. Combinatorially, the right side counts the total number of subsets of an n-element set, as each element can either be included or excluded independently, yielding $2^n possibilities. The left side achieves the same count by partitioning the subsets according to their size k, where \binom{n}{k} is the number of k-element subsets; summing over all k thus equals the total. The hockey-stick identity, \sum_{k=r}^{n} \binom{k}{r} = \binom{n+1}{r+1}, admits a combinatorial proof via committee selection with a distinguished member. The right side counts the ways to choose r+1 elements from \{1, 2, \dots, n+1\} and designate the largest as the chair. For a fixed chair j (where j = r+1 to n+1), the remaining r members are chosen from \{1, 2, \dots, j-1\}, which has \binom{j-1}{r} ways. Summing over possible j yields the left side, matching the total. The Chu-Vandermonde identity generalizes the above to \sum_{k} \binom{a}{k} \binom{b}{n-k} = \binom{a+b}{n} for general complex numbers a and b, often arising in the context of hypergeometric series as a consequence of the binomial theorem for (1+x)^a (1+x)^b = (1+x)^{a+b}. Combinatorial interpretations exist for integer cases via signed sets or lattice paths. Bijective proofs, a subset of combinatorial arguments, establish identities by constructing explicit one-to-one correspondences between sets counted by each side, such as matching committees of different compositions in the Vandermonde case, providing intuitive verifications without enumeration.

Advanced Identities

Dixon's identity is a notable advanced relation among products of three binomial coefficients. For nonnegative integers a, b, and c, it states that \sum_{k=-a}^{a} (-1)^k \binom{a+b}{a+k} \binom{b+c}{b+k} \binom{c+a}{c+k} = \frac{(a+b+c)!}{a! \, b! \, c!}. This identity was originally proved by A. C. Dixon in 1903 using trigonometric integrals. A hypergeometric proof represents the left side as a well-poised {}_3F_2 series at argument 1, which evaluates to the right side via Dougall's theorem for balanced hypergeometric series. For small values, such as a = b = c = 1, the sum over k = -1, 0, 1 yields (-1)^{-1} \binom{2}{0}^3 + \binom{2}{1}^3 + (-1)^{1} \binom{2}{2}^3 = -1 + 8 - 1 = 6, matching the right side \frac{3!}{1! \, 1! \, 1!} = 6. When a = b = c = 2, the sum equals \frac{6!}{2!^3} = 90. A continuous analog of emerges through the , providing integral representations that generalize to non-integer arguments. The is defined as B(x, y) = \int_0^1 t^{x-1} (1-t)^{y-1} \, dt = \frac{\Gamma(x) \Gamma(y)}{\Gamma(x+y)} for \Re(x) > 0, \Re(y) > 0. For positive integers n and k \leq n, \binom{n}{k}^{-1} = (n+1) \int_0^1 t^k (1-t)^{n-k} \, dt, since B(k+1, n-k+1) = \frac{k! (n-k)!}{(n+1)!}. This links discrete binomial coefficients to continuous integrals and extends to the generalized binomial coefficient \binom{\alpha}{\beta} = \frac{\Gamma(\alpha+1)}{\Gamma(\beta+1) \Gamma(\alpha - \beta + 1)} via the reflection formula for the gamma function. The ratio of binomial coefficients can be expressed in terms of falling factorials, facilitating evaluations in hypergeometric contexts. Specifically, \frac{\binom{a}{k}}{\binom{b}{k}} = \prod_{i=0}^{k-1} \frac{a - i}{b - i} = \frac{(a)_k}{(b)_k}, where (z)_k = z (z-1) \cdots (z-k+1) is the falling factorial. This form arises in the terminating Chu-Vandermonde identity for {}_2F_1(-n, b; c; 1) = \frac{(c-b)_n}{(c)_n}, which generalizes summation identities to ratio expressions for partial sums. For example, with a=5, b=3, k=2, the ratio \frac{\binom{5}{2}}{\binom{3}{2}} = \frac{10}{3} = \frac{5 \cdot 4}{3 \cdot 2}. For congruences, Lucas' theorem offers an efficient method to compute \binom{m}{n} \mod p for prime p using base-p digits. If the base-p expansions are m = \sum m_i p^i, n = \sum n_i p^i with $0 \leq m_i, n_i < p, then \binom{m}{n} \equiv \prod_i \binom{m_i}{n_i} \pmod{p}, with the product zero if n_i > m_i for any i. This criterion determines when p divides \binom{m}{n} without full computation. The theorem was established by Édouard Lucas in 1878 as part of periodic functions theory. For example, with p=3, m=10=101_3, n=4=11_3=011_3, \binom{1}{0} \binom{0}{1} \binom{1}{1} = 1 \cdot 0 \cdot 1 = 0, and indeed $210 \equiv 0 \mod 3. Another case, m=7=21_3, n=3=10_3, \binom{2}{1} \binom{1}{0} = 2 \cdot 1 = 2 \equiv 2 \mod 3, matching \binom{7}{3} = 35 \equiv 2 \mod 3.

Generating Functions

Ordinary Generating Functions

The ordinary generating function for the binomial coefficients \binom{n}{k} in the variable x, for fixed nonnegative integer n, is given by the binomial theorem: (1 + x)^n = \sum_{k=0}^n \binom{n}{k} x^k. This expansion directly encodes the coefficients as the number of ways to choose k elements from a set of n elements, providing a polynomial representation that facilitates algebraic manipulations in combinatorics. A prominent example is the generating function for the central binomial coefficients \binom{2n}{n}, which counts the number of ways to choose n elements from $2n elements: \sum_{n=0}^\infty \binom{2n}{n} x^n = \frac{1}{\sqrt{1 - 4x}}, valid for |x| < 1/4. This closed form arises from the binomial series expansion of (1 - 4x)^{-1/2} and appears in enumerative problems such as lattice path counting. Coefficients from ordinary s involving binomial coefficients can be extracted using complex analysis techniques, such as Cauchy's integral formula. For a generating function f(x) = \sum_{n=0}^\infty a_n x^n, the coefficient a_n = [x^n] f(x) is given by a_n = \frac{1}{2\pi i} \oint \frac{f(z)}{z^{n+1}} \, dz, where the contour is a small circle around the origin enclosing no singularities of f(z)/z^{n+1}. Alternatively, the residue theorem computes a_n as the residue of f(z)/z^{n+1} at z = 0, enabling precise evaluation even for infinite series like the central binomial generating function. Ordinary generating functions prove useful in deriving and solving recurrence relations through operations like series multiplication, which corresponds to convolution of coefficients. If A(x) = \sum a_n x^n and B(x) = \sum b_n x^n, their product A(x) B(x) = \sum c_n x^n satisfies c_n = \sum_{k=0}^n a_k b_{n-k}, allowing to model recursive combinatorial structures; for instance, multiplying the generating function for \binom{n}{k} by another series yields recurrences for higher-order selections or compositions. A concrete application is the generating function for subsets of a set with n elements, ordered by size: (1 + x)^n = \sum_{k=0}^n \binom{n}{k} x^k, where the coefficient of x^k counts the k-element subsets. This formalizes the enumeration of power set partitions by cardinality and extends to weighted variants by substituting variables.

Exponential Generating Functions

The exponential generating function provides a powerful analytic tool for studying binomial coefficients in the context of labeled combinatorial structures, where the $1/n! factor accounts for the labeling of elements. For a fixed integer k \geq 0, the sequence \binom{n}{k} for n = 0, 1, 2, \dots (with \binom{n}{k} = 0 for n < k) has exponential generating function \sum_{n=0}^\infty \binom{n}{k} \frac{x^n}{n!} = \frac{x^k}{k!} e^x. This formula arises because \binom{n}{k} = n^{\underline{k}}/k!, where n^{\underline{k}} = n(n-1)\cdots(n-k+1) is the falling factorial, and the exponential generating function for the falling factorial sequence n^{\underline{k}} is x^k e^x. Combinatorially, in the labeled setting, this EGF corresponds to structures consisting of a distinguished k-element set (with EGF \frac{x^k}{k!}) together with an arbitrary labeled set on the remaining elements (with EGF e^x). In labeled structures such as permutations, binomial coefficients appear prominently through the exponential formula, which constructs the generating function for composite objects from those of their connected components. For example, the bivariate exponential generating function marking the number of fixed points in permutations is \sum_{n,k \geq 0} \binom{n}{k} D_{n-k} \frac{x^n}{n!} u^k = e^{u x} \cdot e^{-x}/(1-x), where D_m denotes the number of derangements on m elements, but more directly, the generating function \sum_{n,k \geq 0} \binom{n}{k} u^k \frac{x^n}{n!} = e^{x(1+u)} encodes the binomial coefficients as coefficients of u^k. This bivariate form highlights how binomial coefficients count the ways to choose k labeled fixed points, with the remaining elements permuted arbitrarily. Similarly, for Cayley trees on labeled vertices, the exponential generating function T(x) = x e^{T(x)} involves implicit binomial structures through branch factorials, but the falling factorial connection ties directly to selecting rooted subtrees. The multiplication of exponential generating functions corresponds to the labeled product (disjoint union) of structures, preserving the combinatorial interpretation; for instance, if A(x) and B(x) are exponential generating functions for two classes of labeled objects, then A(x) B(x) generates the class of objects formed by taking one from each class on disjoint label sets. The exponential formula extends naturally to set partitions, where the exponential generating function for the Bell numbers (total partitions) is \sum_{n=0}^\infty B_n \frac{x^n}{n!} = e^{e^x - 1}, derived by exponentiating the generating function e^x - 1 for nonempty labeled sets as components. Binomial coefficients enter via the inclusion-exclusion representation of Stirling numbers of the second kind, S(n,k) = \frac{1}{k!} \sum_{j=0}^k (-1)^{k-j} \binom{k}{j} j^n, which count partitions into k nonempty subsets, with B_n = \sum_k S(n,k). A notable arithmetic property arising from this analytic framework is Touchard's congruence: for a prime p and nonnegative integer n, B_{n+p} \equiv B_n + B_{n+1} \pmod{p}. This congruence captures modular behavior in the growth of set partitions, linking back to the exponential structure. Derangements provide another example where binomial coefficients facilitate indirect computation via inclusion-exclusion: the number of derangements is D_n = \sum_{k=0}^n (-1)^k \binom{n}{k} (n-k)!, with exponential generating function \sum_{n=0}^\infty D_n \frac{x^n}{n!} = \frac{e^{-x}}{1-x}. In probabilistic terms, the number of fixed points in a uniform random permutation on n labels follows the distribution P(K=k) = \binom{n}{k} \frac{D_{n-k}}{n!}, which approximates a Poisson distribution with parameter \lambda = 1 as n \to \infty, reflecting the limiting independence of fixed-point indicators. This Poisson approximation underscores the role of exponential generating functions in bridging combinatorial enumeration with probabilistic limits for labeled structures.

Analytic Properties

Divisibility Properties

Binomial coefficients exhibit significant divisibility properties, particularly with respect to prime numbers. For a prime p and integer k with $1 \leq k \leq p-1, the binomial coefficient \binom{p}{k} is divisible by p. This follows from the factorial representation \binom{p}{k} = \frac{p!}{k!(p-k)!}, where p divides the numerator but not the denominator, as k! and (p-k)! involve no factors of p. A deeper insight into the exact power of a prime dividing a binomial coefficient is provided by Kummer's theorem. This theorem states that the p-adic valuation v_p\left(\binom{n}{k}\right), which is the highest power of prime p dividing \binom{n}{k}, equals the number of carries that occur when adding the base-p representations of k and n-k. For example, if n = 6, k = 1, and p = 2, the base-2 expansions are $6_{10} = 110_2, $1_{10} = 001_2, and $5_{10} = 101_2; adding $001_2 + 101_2 = 110_2 requires one carry, so v_2\left(\binom{6}{1}\right) = 1. Lucas' theorem extends this to compute binomial coefficients modulo a prime p. It asserts that if the base-p digits of n and k are n = n_m p^m + \cdots + n_0 and k = k_m p^m + \cdots + k_0, then \binom{n}{k} \equiv \prod_{i=0}^m \binom{n_i}{k_i} \pmod{p}, provided k_i \leq n_i for all i; otherwise, the product is zero. This allows efficient determination of whether p divides \binom{n}{k} by checking the digits: divisibility holds if there exists an i with k_i > n_i. For instance, for p=3, \binom{5}{2} \equiv \binom{1}{0} \binom{2}{2} \equiv 1 \cdot 1 \equiv 1 \pmod{3}, so 3 does not divide 10. These properties manifest clearly in the rows of , where the entries are binomial coefficients. In the p-th row (index starting at 0), the coefficients are \binom{p}{0} = 1, \binom{p}{1} = p, ..., \binom{p}{p-1} = p, \binom{p}{p} = 1, so all interior entries are divisible by the prime p. More generally, Lucas' theorem reveals fractal-like patterns modulo p, such as the Sierpinski triangle for p=2, where entries divisible by 2 appear in a self-similar structure. The Erdős–Ginzburg–Ziv theorem provides a related result on sums: in any sequence of $2m-1 integers, there exists a of exactly m elements whose sum is divisible by m. This has been applied to show divisibility properties for sums of coefficients in certain rows of .

Bounds and Asymptotic Formulas

coefficients satisfy various inequalities that provide upper and lower bounds on their magnitude, which are essential for analyzing their growth in combinatorial and probabilistic contexts. A fundamental upper bound is given by \binom{n}{k} \leq \left( \frac{en}{k} \right)^k for $1 \leq k \leq n, derived from the inequality k! \geq (k/e)^k applied to the expression of the binomial coefficient. A corresponding lower bound is \binom{n}{k} \geq (n/k)^k. For the central binomial coefficient, Stirling's approximation yields a precise asymptotic formula: \binom{2n}{n} \sim \frac{4^n}{\sqrt{\pi n}} as n \to \infty. This follows from applying Stirling's formula m! \sim \sqrt{2\pi m} (m/e)^m to (2n)! / (n!)^2. When k is fixed and n \to \infty, or more generally when k = o(\sqrt{n}), the binomial coefficient approximates \binom{n}{k} \sim \frac{n^k}{k!}. A refined version using Stirling's approximation is \binom{n}{k} \sim \frac{(ne/k)^k}{\sqrt{2\pi k}}. For k proportional to n, such as in the central regime where k/n \approx 1/2, tighter bounds involve the H(p) = -p \log_2 p - (1-p) \log_2 (1-p), yielding \frac{2^{n H(k/n)}}{n+1} \leq \binom{n}{k} \leq 2^{n H(k/n)}. This entropy bound captures the exponential growth and is particularly sharp near the center. In the large n limit with k near np for fixed p \in (0,1), the local provides a Gaussian : the probability mass \binom{n}{k} / 2^n behaves like \frac{1}{\sqrt{2\pi n p(1-p)}} \exp\left( -\frac{(k - np)^2}{2 n p(1-p)} \right), reflecting the N(np, np(1-p)) for the binomial sum. The total sum \sum_{k=0}^n \binom{n}{k} = 2^n exactly, with refinements from the above approximations showing concentration around k = n/2, where the central term dominates as \binom{n}{n/2} \sim 2^n / \sqrt{\pi n / 2}. For non-integer arguments, the generalized binomial coefficient \binom{\alpha}{k} = \frac{\Gamma(\alpha+1)}{\Gamma(k+1) \Gamma(\alpha - k + 1)} admits asymptotic expansions via on the , such as \ln \left( \frac{\Gamma(z + 1/2)}{\Gamma(z)} \right) \sim \frac{1}{2} \ln z + \sum_{j=0}^\infty (-1)^{j+1} \tilde{\beta}_j z^{-(2j+1)} for large z, enabling approximations for large non-integer \alpha.

Generalizations

Multinomial Coefficients

Multinomial coefficients generalize the concept of binomial coefficients to scenarios involving more than two categories or groups. The multinomial coefficient, denoted by \binom{n}{k_1, k_2, \dots, k_m}, is defined as \binom{n}{k_1, k_2, \dots, k_m} = \frac{n!}{k_1! \, k_2! \cdots k_m!}, where n is a non-negative and k_1, k_2, \dots, k_m are non-negative s satisfying \sum_{i=1}^m k_i = n. This formula counts the number of ways to a set of n distinct objects into m distinct subsets of sizes k_1, k_2, \dots, k_m, respectively. When m=2, the multinomial coefficient reduces to the binomial coefficient, as \binom{n}{k_1, k_2} = \binom{n}{k_1} where k_2 = n - k_1. This connection highlights how binomial coefficients serve as a special case of the more general multinomial structure, extending the paradigm to multiple labeled groups. Combinatorially, multinomial coefficients arise in problems of distributing n distinct items into m labeled bins, with k_i items in the i-th bin, yielding \binom{n}{k_1, k_2, \dots, k_m} distinct arrangements. For instance, the coefficient \binom{3}{1,1,1} equals 6, representing the ways to assign three distinct letters to three distinct positions, one each. This interpretation underscores their role in counting permutations of multisets. The multinomial theorem provides an algebraic expansion for powers of sums involving multiple variables: (x_1 + x_2 + \dots + x_m)^n = \sum_{\substack{k_1 + \dots + k_m = n \\ k_i \geq 0}} \binom{n}{k_1, k_2, \dots, k_m} \prod_{i=1}^m x_i^{k_i}, where the sum is over all non-negative integer solutions to the equation \sum k_i = n. This theorem, provable by induction on n or combinatorial arguments, generalizes the and appears in expansions like for multivariate functions. Multinomial coefficients can be computed iteratively using successive binomial coefficients, reflecting the step-by-step selection process: \binom{n}{k_1, k_2, \dots, k_m} = \binom{n}{k_1} \binom{n - k_1}{k_2} \cdots \binom{n - k_1 - \dots - k_{m-1}}{k_m}. This recursive relation facilitates numerical evaluation, particularly for large n, by leveraging efficient computation methods.

Negative and Non-Integer Arguments

The generalized binomial coefficient extends the concept to real or upper index α and non-negative lower index k, defined by the falling factorial product \binom{\alpha}{k} = \frac{\alpha (\alpha - 1) \cdots (\alpha - k + 1)}{k!} for k ≥ 1, with \binom{\alpha}{0} = 1. This definition preserves the combinatorial interpretation when α is a non-negative but allows for other values of α. When α = -n for positive integer n, the formula yields the negative binomial coefficient \binom{-n}{k} = (-1)^k \binom{n + k - 1}{k}, which arises in the expansion of (1 - x)^{-n} = \sum_{k=0}^\infty \binom{n + k - 1}{k} x^k for |x| < 1. This identity follows directly from substituting α = -n into the generalized definition and simplifying the product, relating it to the rising factorial or (n)_k = n(n+1)\cdots(n+k-1). Isaac Newton generalized the binomial theorem to non-integer exponents in his 1676 work, establishing the infinite series (1 + x)^\alpha = \sum_{k=0}^\infty \binom{\alpha}{k} x^k valid for |x| < 1 and complex α not a non-negative integer. This binomial series provides the Taylor expansion around x = 0 for functions like (1 + x)^\alpha, enabling approximations in analysis and physics. For half-integer values such as α = 1/2, the coefficients take the explicit form \binom{1/2}{k} = (-1)^{k-1} \frac{(2k-2)!}{2^{2k-1} (k-1)! k!} for k ≥ 1, derived from the product formula. These appear in the series for \sqrt{1 + x} and, through integration, contribute to expansions like that of \arcsin x, where the coefficients relate to central binomial terms in the integrand (1 - x^2)^{-1/2}.

q-Series and Other Extensions

The q-binomial coefficient, also known as the Gaussian binomial coefficient, provides a q-deformation of the classical binomial coefficient. It is defined as \binom{n}{k}_q = \prod_{i=1}^k \frac{[n-i+1]_q}{_q}, where _q = \frac{1 - q^m}{1 - q} denotes the q-analogue of the integer m, for q \neq 1 and nonnegative integers n and k with k \leq n. This polynomial in q reduces to the ordinary binomial coefficient \binom{n}{k} in the limit as q \to 1. These coefficients arise prominently in the theory of quantum groups, where they appear in the structure constants of quantum universal enveloping algebras and in the q-analogues of Clebsch-Gordan coefficients for representations. In combinatorics, q-binomial coefficients enumerate partitions fitting inside a k by (n-k) rectangle, with the exponent of q tracking statistics such as the area or major index of the partitions. They also serve as building blocks for basic hypergeometric series, which are q-analogues of the classical hypergeometric functions, facilitating identities like the q-binomial theorem: \prod_{i=0}^{n-1} (1 + q^i z) = \sum_{k=0}^n \binom{n}{k}_q z^k. Binomial coefficients extend to infinite cardinals in set theory via cardinal arithmetic. For infinite cardinals \kappa and \lambda with \lambda \leq \kappa, \binom{\kappa}{\lambda} is defined as the cardinality of the set of all subsets of a set of cardinality \kappa with exactly \lambda elements. This yields \binom{\kappa}{\lambda} = \kappa when \lambda is finite and \kappa infinite, and \binom{\kappa}{\kappa} = 2^\kappa, the power set cardinality. More generally, under the generalized continuum hypothesis or in models of ZFC, \binom{\kappa}{\lambda} = \kappa^\lambda when \lambda is infinite and \kappa \geq 2^\lambda, though exact values depend on the continuum hypothesis for certain pairs. (Jech, Set Theory, 3rd ed., 2003) A generalization to multisets involves rising factorials with a step parameter r > 0. Using the rising Pochhammer symbol (n)_k^r = n (n + r) \cdots (n + (k-1) r), the coefficient is defined as \binom{n}{k}_r = \frac{(n)_k^r}{k!}. When r=1, it recovers the standard combinations with repetition \binom{n+k-1}{k}. The reciprocal of the binomial coefficient admits partial fraction decompositions in certain contexts, particularly for generating functions or representations. For fixed n and variable k, expressions like \frac{1}{\binom{n}{k}} = (n+1) \int_0^1 t^{k} (1-t)^{n-k} \, dt can be derived, but direct partial fraction forms appear in sums or q-analogues, such as decomposing generating functions involving s into poles at roots of unity or binomial denominators, aiding evaluations of series like \sum_k \frac{1}{\binom{n}{k} x^k}. These decompositions are useful in for asymptotic analysis of sums.

References

  1. [1]
    Binomial Coefficient -- from Wolfram MathWorld
    The binomial coefficient (n; k) is the number of ways of picking k unordered outcomes from n possibilities, also known as a combination or combinatorial number.Missing: authoritative sources
  2. [2]
    1.3 Binomial coefficients
    The entries in Pascal's Triangle are the coefficients of the polynomial produced by raising a binomial to an integer power.
  3. [3]
    Binomial Theorem -- from Wolfram MathWorld
    The binomial theorem has several related results, including the binomial series identity, and is also known as binomial formula or expansion. It was known by ...
  4. [4]
    Pascal's Triangle -- from Wolfram MathWorld
    where (n; r) is a binomial coefficient. The triangle was studied by B. Pascal, in whose posthumous work it appeared in 1665 (Pascal 1665).
  5. [5]
    [PDF] Pingala and the Beginnings of Combinatorics in India - IISc Math
    Recorded maths in India begins with evidence from the archaeological remains of the Indus Valley (Harappan) civilisation. (IVC) (∼ 2600 − 1800 BCE).
  6. [6]
    (PDF) A HISTORY OF PIṄGALA'S COMBINATORICS - ResearchGate
    The tradition began with the formal theory of Sanskrit meters formulated by Piṅgala in the 2 nd century B.C.E. His recursive algorithms are the first example of ...
  7. [7]
    al-Karaji (953 - 1029) - Biography - MacTutor History of Mathematics
    Al-Karaji was an Islamic mathematician who wrote about the work of earlier mathematicians and who can be regarded as the first person to free algebra from ...
  8. [8]
    Muhammad Al-Karaji: A Mathematician Engineer from the Early 11th ...
    Jun 4, 2009 · (Source). In the Fakhri, Al-Karaji studied the successive powers of a binomial, developed it further in the Badi', and concluded his analysis ...
  9. [9]
    Omar Khayyam (1048 - 1131) - Biography - MacTutor
    In fact we can be fairly sure that Khayyam used a method of finding nth roots based on the binomial expansion, and therefore on the binomial coefficients.
  10. [10]
    Umar Khayyam - Stanford Encyclopedia of Philosophy
    Sep 6, 2011 · ... Khayyam had used the binomial theorem. \[(a + b)^n = a^n + na^{n-1}b + \cdots + nab^{n-1}+ b^n\]. in his method, improving on and extending ...
  11. [11]
    Blaise Pascal - Biography - MacTutor - University of St Andrews
    He discovered that the sum of the angles of a triangle are two right angles and, when his father found out, he relented and allowed Blaise a copy of Euclid. At ...
  12. [12]
    [PDF] Pascal's Treatise on the Arithmetical Triangle
    Pascal, however, was the first to connect binomial coefficients with combinatorial coefficients in probability. In fact, a major motivation for Pascal was a ...
  13. [13]
    Earliest Uses of Symbols in Probability and Statistics - MacTutor
    The modern notation, using parentheses and no fraction bar, appears in 1826 in Die Combinatorische Analyse by Andreas von Ettingshausen [Henry W. Gould]. ...
  14. [14]
    Introduction to the factorials and binomials - Wolfram Functions Site
    A special role in the history of the factorial and binomial belongs to L. Euler, who introduced the gamma function as the natural extension of factorial ( ) ...
  15. [15]
    [PDF] The Binomial Coefficient for Negative Arguments - arXiv
    Mar 30, 2015 · n k. =.. n! k!(n − k)! if 0 ≤ k ≤ n. 0 otherwise. (1.2). The value of the binomial coefficient (1.1) for negative integer n and integer k ...
  16. [16]
    [PDF] 6.1 Gamma Function, Beta Function, Factorials, Binomial Coefficients
    for each required factorial. Better to go back to basics, holding gammln in ... If your problem requires a series of related binomial coefficients, a good idea.
  17. [17]
    Binomial Coefficients - Algorithms for Competitive Programming
    Oct 22, 2024 · Binomial coefficients are the number of ways to select a set of elements from different elements without taking into account the order of arrangement of these ...
  18. [18]
    Pascal's Formula -- from Wolfram MathWorld
    Pascal's Formula ; (n; r), = (n!)/((n-r)!r!) ; = ((n-1)!n)/((n-r)!r ; = ((n-1)!(n-r))/((n-r) ; = ((n-1)!)/((n-r-1)! ; = (n-1; r)+(n-1; r-1.Missing: authoritative source
  19. [19]
    [PDF] Module 4 Dynamic Programming - Jackson State University
    Binomial coefficients are coefficients of the binomial formula: (a + b)n = C ... = 0 otherwise. Time Complexity: Θ(nm). Space Complexity: Θ(nm).
  20. [20]
    [PDF] fibonacci diagonals
    For example, highlighted in Pascal's triangle above are the 5th and 8th diagonals. We have. 1+3+1=5= F5 and. 1 + 6 + 10 + 4 = 21 = F8.
  21. [21]
    3.1 Pascal's Arithmetical Triangle - Oscar Levin
    However, we are interested in counting questions, so our main goal now is to observe how the numbers of Pascal's triangle are answers to a variety of counting ...
  22. [22]
    [PDF] Combinatorial interpretation of the binomial theorem - UMD MATH
    We have the definition of the binomial coefficient. (n k. ) = n! k!(n − k)! . n-choose-k. Let S(k, n) denote the collection of subsets of ...
  23. [23]
    [PDF] Combinatorial Identities: Binomial Coefficients, Pascal's Triangle ...
    Sep 9, 2011 · Binomial Coefficients: Interpretation. (n k. ) = the # of ways to choose a subset with k elements from a set with n elements. Equivalently: (n.
  24. [24]
    [PDF] Lecture Notes 1 Basic Properties of Binomial Coefficients - csail
    Apr 4, 2000 · Combinatorial proof: Consider choosing a committee of size r and a leader, from n people. One way is to pick the leader first (n ways), then ...
  25. [25]
    [PDF] Basic Combinatorics - UTK Math
    k is read “ n choose k,” or “the kth binomial coefficient of order n.” Certain values of n k are immediately obvious. For example, n. 0 = 1, since the only ...
  26. [26]
    7.4 - Hypergeometric Distribution | STAT 414 - STAT ONLINE
    Note that one of the key features of the hypergeometric distribution is that it is associated with sampling without replacement. ... binomial distribution.
  27. [27]
    BinomialCoefficients
    Base cases: If k = 0, then there is exactly one zero-element set of our n-element set—it's the empty set—and we have.
  28. [28]
    [PDF] Counting III
    Apr 12, 2010 · The second proof technique is called a “combinatorial proof.” 3 Vandermonde's Identity. Here's another useful identity: Claim 2 ...
  29. [29]
    [PDF] 2 - Strings and Binomial Coefficients - William T. Trotter
    Nov 14, 2017 · Explanation A lattice path corresponds to a choice of m horizontal moves in a sequence of m + n moves. Here the choices are: RUURRRURUR. Page 37 ...
  30. [30]
    [PDF] q-combinatorics
    Feb 7, 2019 · A combinatorial interpretation is the following: Claim 2 cn,k,α is the number of lattice walks from (0,0) to (k, n - k) (using the usual ...<|control11|><|separator|>
  31. [31]
    Lesson 10: The Binomial Distribution | STAT 414
    To understand the effect on the parameters and on the shape of a binomial distribution. To derive formulas for the mean and variance of a binomial random ...
  32. [32]
    The Binomial Distribution
    The binomial distribution for a random variable X with parameters n and p represents the sum of n independent variables Z which may assume the values 0 or 1.
  33. [33]
    [PDF] The Binomial Distribution - Academic Web
    So, for example, using a binomial distribution, we can determine the probability of getting 4 heads in 10 coin tosses. How does the binomial distribution do ...Missing: binary | Show results with:binary
  34. [34]
    Tech Tips: Binomial Distributions
    R: use the function dbinom(x=k, size=n, prob=p). As an example, to find the probability that one flips exactly 4 heads in 8 tosses of a fair coin: · Excel: use ...
  35. [35]
    [PDF] Important Probability Distributions
    Binomial probabilities can be computed using the Excel function BINOMDIST(). Two other examples are given in a separate Excel file. = V (X) = np(1 − p).
  36. [36]
    Binomial Test - University of Texas at Austin
    A binomial test uses sample data to determine if the population proportion of one level in a binary (or dichotomous) variable equals a specific claimed value.
  37. [37]
    8.1.2 - Hypothesis Testing | STAT 200
    2 - Hypothesis Testing. A hypothesis test for a proportion is used when you are comparing one group to a known or hypothesized population proportion value.
  38. [38]
    2.2 - Tests and CIs for a Binomial Parameter | STAT 504
    The most common choice for α is 0.05, which gives 95% confidence and multiplier z .025 = 1.96 . We should also mention here that one-sided confidence intervals ...<|separator|>
  39. [39]
    Confidence Intervals for the binomial parameter p - MIT Mathematics
    Sep 14, 2015 · But for the binomial case there are no precise confidence intervals. The binomial random variable X has just n + 1 possible values 0,1, ..., n, ...
  40. [40]
    [PDF] STAT 516 - Central Limit Theorem
    Apr 7, 2020 · Normal approximation to Binomial:de Moivre-Laplace. Central Limit Theorem. ▷ Let X = Xn ∼ Bin(n,p). Then, for any fixed p and real-valued x.
  41. [41]
    [PDF] The binomial distribution, and a normal approximation - ResearchGate
    In other words, the De Moivre–Laplace central limit theorem tells us that if n is large, then the binomial probability of having between a and b successes ...
  42. [42]
    [PDF] Chapter 3.3, 4.1, 4.3. Binomial Coefficient Identities - UCSD Math
    Prove by counting (k + 1)-element subsets of [n + 1] in two ways. First way: The number of such subsets is. (n+1 k+1. ) . Prof. Tesler. Binomial Coefficient ...
  43. [43]
    [PDF] Falling Factorials, Generating Functions, and Conjoint Ranking Tables
    Falling factorial powers satisfy. ∆xn = nxn−1 ,. 4. Page 5. where ∆ is the forward difference operator. A little more generally, using the two-variable.<|control11|><|separator|>
  44. [44]
    [PDF] 0.1. Linear transformations - MIT
    This is called the binomial basis of Pn. (a) Represent the maps Ta, Tb, Tg and Th of Exercise 0.2 with respect to the binomial bases. ( ...
  45. [45]
    Conversions between standard polynomial bases - Terence Tao
    Apr 7, 2019 · These coefficients describe how to convert a polynomial expressed in the {Q_n} basis into a polynomial expressed in the {R_k} basis.
  46. [46]
    [PDF] The Pascal Matrix Function and Its Applications to Bernoulli ...
    In this paper, a uni- fied approach by means of P[x] is presented in the study of Bernoulli polynomials and numbers and Euler polynomials and numbers. For any ...
  47. [47]
    [PDF] Introduction to Real Orthogonal Polynomials - DTIC
    DISCRETE EXTENSIONS. We turn now to orthogonal polynomial classes which use the discrete inner product introduced in Chapter 11, Section A.2. 53. Page 62. < f, ...
  48. [48]
    [PDF] Interpolation and Polynomial Approximation Divided Differences ...
    Another form, Newton's Forward Difference Formula is constructed by using the forward difference operator ∆: ∆f (xn) = f (xn+1) − f (xn) using this ...
  49. [49]
    Hockey Stick Identity | Brilliant Math & Science Wiki
    The hockey stick identity is an identity regarding sums of binomial coefficients ... Proofs. Inductive Proof of Hockey Stick Identity: Base Case: r = n r=n ...Missing: source | Show results with:source
  50. [50]
    [PDF] generatingfunctionology - Penn Math
    May 21, 1992 · Preface. This book is about generating functions and some of their uses in discrete mathematics. The subject is so vast that I have not ...
  51. [51]
    [PDF] Binomial Identities and Hypergeometric Series - cs.wisc.edu
    The Chu-Vandermonde identity: (n, -bc (c +) (2.7). The Pfaff-Saalschiitz identity: c, d (c)n(c + a + b)n (2.8) where d = 1 - a - b - n - c, that is, the ...
  52. [52]
    Chu-Vandermonde Identity -- from Wolfram MathWorld
    The Chu-Vandermonde identity _2F_1(-n,b;c;1)=((c-b)_n)/((c)_n) (1) (for n in Z^+) is a special case of Gauss's hypergeometric theorem _2F_1(a,b;c;1) ...
  53. [53]
    Lucas' theorem: its generalizations, extensions and applications (1878
    Sep 6, 2014 · In this article, consisting of six sections, we provide a historical survey of Lucas type congruences, generalizations of Lucas' theorem modulo prime powers.Missing: original | Show results with:original
  54. [54]
    [PDF] AC.pdf - Analytic Combinatorics
    Analytic combinatorics aims to enable precise quantitative predictions of the proper- ties of large combinatorial structures. The theory has emerged over ...
  55. [55]
    [PDF] On the divisibility of binomial coefficients
    which a prime power pa divides a binomial coefficient n k are given by Kummer's. Theorem [10] and also by a generalized form of Lucas' Theorem [5, 12]. Still ...
  56. [56]
    Über die Ergänzungssätze zu den allgemeinen Reciprocitätsgesetzen.
    Kummer, E.E.. "Über die Ergänzungssätze zu den allgemeinen Reciprocitätsgesetzen.." Journal für die reine und angewandte Mathematik 44 (1852): 93-146.Missing: Erganzungssätze pdf
  57. [57]
    [PDF] Eric Rowland - Binomial coefficients, valuations, and words
    Kummer's theorem is the first of many results to express arithmetic informa- tion about binomial coefficients in terms of base-p representations of integers.Missing: paper | Show results with:paper
  58. [58]
    [PDF] THÉORIE DES FONCTIONS NUMERIQUES SIMPLEMENT ...
    CE mémoire a pour objet l'étude des fonctions symétriques des racines d'une équation du second degré, et son application à la théorie des nombres premiers.
  59. [59]
    [PDF] arXiv:1704.05872v2 [math.NT] 24 Mar 2018
    Mar 24, 2018 · Kummer's theorem follows easily from Legendre's formula. (2) νp(m!) = m − σp(m) p − 1 for the p-adic valuation of m!. Our first ...
  60. [60]
    [PDF] arXiv:1701.04942v3 [math.CO] 29 Jun 2019
    Jun 29, 2019 · Paul Erdős, Abraham Ginzburg, and Abraham Ziv, A Theorem in additive number theory, Bull. Res. Council Israel 10F (1961), 41–43. MR 3618568.
  61. [61]
    [PDF] A brief note on estimates of binomial coefficients
    upper bound of 2n. replace it with an exponential expression by making use of Stirling's Approximation. 12In other words, n tends to infinity. Theorem 1 ( ...Missing: \binom<|separator|>
  62. [62]
    254A, Notes 0a: Stirling's formula - Terry Tao - WordPress.com
    Jan 2, 2010 · In this supplemental set of notes we derive some approximations for {n!}, when {n} is large, and in particular Stirling's formula.
  63. [63]
    [PDF] Useful Bounds
    Binomial Coefficients. A simple, but loose, bound: (nk) k. ≤ ( n k) ≤ ( ne k ) k . A slightly more advanced bound is. 2nH(k/n) n + 1 ≤ ( n k) ≤. 2nH(k/n),. (1).
  64. [64]
    275A, Notes 4: The central limit theorem
    - **Local Central Limit Theorem (LCLT) for Binomial Coefficients/Sums**:
  65. [65]
    [PDF] Asymptotic approximation of central binomial coefficients with ...
    We show that a well-known asymptotic series for the logarithm of the central binomial coefficient is strictly enveloping, so the error incurred in truncating ...
  66. [66]
    [PDF] Multinomial coefficients - Purdue Math
    Multinomial coefficients, defined as 𝑛 𝑛1 ,𝑛2 ,…,𝑛𝑟, count ways to divide 𝑛 distinct things into 𝑟 distinct groups of sizes 𝑛1 ,𝑛2 ,…,and 𝑛𝑟.
  67. [67]
    [PDF] Lecture 2
    Sep 14, 2020 · Multinomial Coefficient is the number of ways to sort N distinct objects into bins, each having ni objects. 2 Extension to Multinomial Theorem.
  68. [68]
    [PDF] Section 1.9. Multinomial Coefficients
    Jul 18, 2019 · Note. We can prove the Multinomial Theorem using mathematical induction. For k = 2, the Multinomial Theorem reduces to the Binomial Theorem.
  69. [69]
    [PDF] Lecture 1.4: Binomial and multinomial coefficients
    Multinomial coefficients generalize binomial coefficients (the case when r = 2). Not surprisingly, the Binomial Theorem generalizes to a Multinomial Theorem.
  70. [70]
    [PDF] Multinomial coefficients - Arizona Math
    Feb 16, 2011 · A multinomial coefficient is associated with each (finite) multiset taken from the set of natural numbers. Such a multi-set is given by a list ...
  71. [71]
    [PDF] 21-228 Week 3 Notes - andrew.cmu.ed
    Sep 16, 2001 · we can use multinomial coefficients to find the coefficients of (x1 + ··· + xm)n: Theorem 1.3 (Multinomial Theorem). (x1 + ··· + xm)n is the ...
  72. [72]
    [PDF] Binomial and Multinomial Coefficients
    Binomial and Multinomial Coefficients. Page 1. Binomial and Multinomial Coefficients. The. allows one to compute the number of combinations of things taken.
  73. [73]
    Generalized Binomial Coefficient -- from Wolfram MathWorld
    Binomial Coefficients. Generalized Binomial Coefficient. See. Binomial Coefficient · About MathWorld · MathWorld Classroom · Contribute · MathWorld Book ...
  74. [74]
    Negative Binomial Series -- from Wolfram MathWorld
    The series which arises in the binomial theorem for negative integer -n ,. (x+a)^(-n), = sum_(k=0)^(infty)(-n; k). (1). = sum_(k=0)^(infty)(-1)^k.
  75. [75]
    Binomial Series -- from Wolfram MathWorld
    There are several related series that are known as the binomial series. The most general is (x+a)^nu=sum_(k=0)^infty(nu; k)x^ka^(nu-k), (1) where (nu; ...
  76. [76]
    [PDF] The q-Binomial Coefficient for Negative Arguments and ... - arXiv
    Jan 11, 2023 · 1 Definitions and Basic Identities. Let the following definition of the q-binomial coefficient, also called the Gaussian polynomial, be given ...
  77. [77]
    [PDF] Braids, q-binomials and quantum groups - Cornell Mathematics
    For the q-binomial coefficients Pascal's identity says that h ni i = qn−ih n − 1 i − 1 i+ h n − 1 i i = h n − 1 i − 1 i. + qih n − 1 i i . Its generalization to ...
  78. [78]
    [PDF] THE q-SERIES IN COMBINATORICS; PERMUTATION STATISTICS ...
    May 5, 2011 · Partitions of integers and q-binomial coefficients. The interpreta- tion of the q-binomial coefficients in terms of partitions of integers (see.
  79. [79]
    [PDF] q-SPECIAL FUNCTIONS, BASIC HYPERGEOMETRIC SERIES AND ...
    Aug 10, 2018 · Abstract. In the lecture notes we start off with an introduction to the q-hypergeometric series, or basic hypergeometric series, and we ...
  80. [80]
    [PDF] arXiv:2012.12691v2 [math.HO] 2 May 2023
    May 2, 2023 · Given n, k ∈ N, the corresponding multiset coefficient is by definition: ... We have a remarkable expansion formula for the rising factorial ...
  81. [81]
    [PDF] arXiv:2306.10655v2 [math.CA] 2 Dec 2023
    Dec 2, 2023 · in α also generates a partial fraction form with respect to j which will be useful in ... reciprocal binomial coefficient and this decreases.