Fact-checked by Grok 2 weeks ago

Central binomial coefficient

The central binomial coefficient \binom{2n}{n} is the binomial coefficient that selects the middle entry in the $2nth row of (for nonnegative integer n), given explicitly by the formula \frac{(2n)!}{(n!)^2}. It counts the number of ways to choose n items from a set of $2n distinct items without regard to order, and equivalently, the number of monotonic paths along the edges of a grid with n \times n square cells. This coefficient also equals the sum of the squares of the binomial coefficients in the nth row of : \sum_{k=0}^n \binom{n}{k}^2 = \binom{2n}{n}. Central binomial coefficients play a fundamental role in and , serving as a cornerstone for counting problems and due to their rich structure and frequent appearances in series expansions. The ordinary for the sequence is \sum_{n=0}^\infty \binom{2n}{n} x^n = \frac{1}{\sqrt{1-4x}} for |x| < \frac{1}{4}, which arises in contexts like the enumeration of binary trees and random walks. They are intimately connected to the Catalan numbers C_n = \frac{1}{n+1} \binom{2n}{n}, which quantify diverse structures such as correctly matched parentheses and non-crossing partitions. Asymptotically, the central binomial coefficient grows like \binom{2n}{n} \sim \frac{4^n}{\sqrt{\pi n}} as n \to \infty, a result derived from Stirling's approximation to the factorials, with more refined expansions involving higher-order terms for precise estimates in large-n applications. These coefficients exhibit notable number-theoretic properties, such as never being prime for n > 1 and having only finitely many squarefree values (specifically 2, 6, and 70 for n=1,2,4), as established through advanced theorems in . Their study extends to inequalities, divisibility criteria by primes, and integrals, underscoring their utility across pure and .

Definition and Basics

Definition

The binomial coefficient \binom{m}{k} is defined as the number \frac{m!}{k!(m-k)!}, where m and k are non-negative integers with k \leq m, representing the number of ways to choose k elements from a set of m elements without regard to order. The nth central is the specific binomial coefficient \binom{2n}{n}, also denoted C(2n, n), given explicitly by the formula \binom{2n}{n} = \frac{(2n)!}{(n!)^2}. This quantity arises as the central entry in the (2n)th row of , where the entries are the binomial coefficients \binom{2n}{k} for k = 0, 1, \dots, 2n, and \binom{2n}{n} is the largest among them due to the and of the row. The term "central binomial coefficient" reflects its position at the center of the even-length row in . These coefficients appeared in early combinatorial studies during the , notably in Leonhard Euler's (1748), where binomial expansions and their applications in series were explored. They possess significant combinatorial interpretations, such as counting balanced parentheses or lattice paths, which are elaborated in subsequent sections.

Examples and Computation

The central binomial coefficient \binom{2n}{n} for small values of n illustrates its rapid growth. The sequence begins with \binom{0}{0} = 1, \binom{2}{1} = 2, \binom{4}{2} = 6, and continues as shown in the following table for n = 0 to $10:
n\binom{2n}{n}
01
12
26
320
470
5252
6924
73432
812870
948620
10184756
These values are cataloged in the as A000984. For small n, \binom{2n}{n} can be computed directly using the definition \binom{2n}{n} = \frac{(2n)!}{(n!)^2}, where factorials are calculated via successive multiplication. This approach is straightforward but becomes inefficient for larger n due to the need to handle increasingly large intermediate factorials. An alternative is the multiplicative formula \binom{2n}{n} = \prod_{k=1}^n \frac{n + k}{k}, which computes the value as a running product, reducing the size of intermediates compared to full factorials. A useful recursive relation for iterative computation is \binom{2n}{n} = \frac{4n - 2}{n} \binom{2n-2}{n-1}, with base case \binom{0}{0} = 1. To derive this, start with the ratio: \frac{\binom{2n}{n}}{\binom{2n-2}{n-1}} = \frac{\frac{(2n)!}{(n!)^2}}{\frac{(2n-2)!}{((n-1)!)^2}} = \frac{(2n)!}{(2n-2)!} \cdot \frac{((n-1)!)^2}{(n!)^2} = (2n)(2n-1) \cdot \frac{1}{n \cdot n} = \frac{2n(2n-1)}{n^2} = \frac{2(2n-1)}{n} = \frac{4n-2}{n}. Rearranging yields the recurrence. This relation enables computation by building up from smaller values, avoiding full evaluations and useful for programming implementations. Computing \binom{2n}{n} for large n presents challenges, such as in fixed-precision integer arithmetic, where values exceed standard data types like 64-bit integers beyond n \approx 30. In applications, such as studying divisibility or congruences a prime p, is employed to compute \binom{2n}{n} \mod p without full expansion, often using Lucas' theorem or specialized recurrences to avoid entirely. As of 2025, modern software supports efficient evaluation using . In 3.8 and later, the math.comb(2*n, n) function implements the multiplicative with built-in big integers for exact results up to very large n. Similarly, Mathematica's Binomial[2 n, n] leverages arbitrary-precision for exact values. For extremely large n (e.g., n > 10^6), where even arbitrary-precision multiplication becomes costly, (FFT)-based methods accelerate the computation of the product in the multiplicative by enabling rapid for the numerator and denominator terms. For estimating values without exact , asymptotic approximations provide scale, such as \binom{2n}{n} \sim \frac{4^n}{\sqrt{\pi n}} via Stirling's (detailed in later sections).

Combinatorial Interpretations

Classical Interpretations

The central binomial coefficient \binom{2n}{n} counts the number of ways to choose n elements from a set of $2n distinct elements. This interpretation follows directly from the general combinatorial definition of binomial coefficients as the number of combinations. Equivalently, it enumerates the number of strings of $2n containing exactly n zeros and n ones, where each string corresponds to selecting positions for the ones (or zeros). In lattice path enumeration, \binom{2n}{n} gives the number of monotonic paths from (0,0) to (n,n) consisting of n right steps (1,0) and n up steps (0,1), with no restrictions on crossing the main diagonal y = x. These paths represent all possible orders of taking the required steps to reach the endpoint. For comparison, the restricted case of paths that stay below or on the diagonal corresponds to the Catalan number \frac{1}{n+1} \binom{2n}{n}. The ballot theorem provides another classical context, where \binom{2n}{n} counts the total number of vote-counting sequences in an election between two candidates each receiving exactly n votes. Each sequence is an arrangement of n votes for one candidate and n for the other. The theorem itself specifies the number of such sequences in which one candidate remains strictly ahead throughout the count, given by \frac{1}{n+1} \binom{2n}{n}. A simple application arises in partitioning problems: \binom{2n}{n} is the number of ways to divide $2n distinct objects into two labeled groups of n each, such as assigning to two distinguishable teams. For unlabeled groups, where the partitions are unordered, the count is \frac{1}{2} \binom{2n}{n} when n > 0, since each is counted twice in the labeled case.

Extensions and Variations

One prominent extension of the central binomial coefficient involves q-analogues, where the Gaussian binomial coefficient \dbinom{2n}{n}_q serves as a deformation parameterized by q. This generalization counts the number of n-dimensional subspaces in a (2n)-dimensional vector space over the finite field \mathbb{F}_q, offering insights into combinatorial structures in finite geometries. Additionally, it is the generating function for the number of integer partitions whose Young diagrams fit within an n by n square (weighted by q to the power of the partition's size), connecting to q-series and partition theory. In probability theory, the central binomial coefficient \dbinom{2n}{n} arises as the numerator in the probability of observing exactly n heads in 2n independent flips of a , yielding \dbinom{2n}{n} / 4^n. For large n, this probability aligns with the , approximating the peak of the standardized and highlighting the concentration of the around its mean. Geometrically, the central binomial coefficient \dbinom{2n}{n} counts the number of lattice points in the n-dimensional defined by non-negative coordinates with the sum of coordinates at most n, equivalent to the n-dilate of the standard n-. This interpretation ties into Ehrhart theory, where the number of lattice points in the t-dilate of the simplex is given by the \dbinom{t + n}{n} evaluated at t = n, bridging and volume computations. In , it also emerges in enumerating points on certain hypersurfaces, such as those arising in toric varieties associated with scaled simplices. Variations appear in , where the number of perfect matchings in the K_{2n} is given by (2n-1)!! = \dbinom{2n}{n} \cdot n! / 2^n, linking the central binomial coefficient to on matchings.

Algebraic Properties

Identities and Recurrences

The central binomial coefficient satisfies several fundamental identities and recurrence relations that highlight its algebraic structure. One key identity is the of binomial coefficients, given by \sum_{k=0}^n \binom{n}{k}^2 = \binom{2n}{n}. This follows as a special case of Vandermonde's convolution identity, \sum_{k=0}^r \binom{m}{k} \binom{n}{r-k} = \binom{m+n}{r}, by setting m = n and r = n, since \binom{n}{n-k} = \binom{n}{k}. Vandermonde's identity itself admits a combinatorial proof: the right side counts the number of ways to choose r elements from a set of m+n elements by selecting k from the first m and r-k from the last n, which equals the left side. An algebraic proof uses the binomial theorem applied to (1+x)^{m+n} = (1+x)^m (1+x)^n and equating coefficients of x^r. Recurrence relations for the central binomial coefficient can be derived from Pascal's identity, \binom{n}{k} = \binom{n-1}{k} + \binom{n-1}{k-1}, which has a combinatorial interpretation: the k-subsets of an n-set are partitioned based on whether they include a distinguished element. Applying symmetry \binom{2n-1}{n} = \binom{2n-1}{n-1} and substituting into Pascal's identity for \binom{2n}{n} yields \binom{2n}{n} = 2 \binom{2n-1}{n}. Then, using Pascal's identity again on \binom{2n-1}{n} = \frac{2n-1}{n} \binom{2n-2}{n-1}, we obtain the ratio recurrence \frac{\binom{2n}{n}}{\binom{2n-2}{n-1}} = \frac{4n-2}{n}. This relation allows iterative computation of central binomial coefficients starting from \binom{0}{0} = 1. The central binomial coefficient also exhibits symmetry in the context of the binomial theorem. The expansion (1+x)^{2n} = \sum_{k=0}^{2n} \binom{2n}{k} x^k is symmetric in the coefficients \binom{2n}{k} = \binom{2n}{2n-k}, with the central term \binom{2n}{n} appearing at k=n. Setting x=1 gives \sum_{k=0}^{2n} \binom{2n}{k} = 4^n, underscoring the central term's prominence. A useful lower bound arises from this expansion: since the coefficients are unimodal and peak at , \binom{2n}{n} is the largest term, and there are $2n+1 terms summing to $4^n, it follows that \binom{2n}{n} \geq \frac{4^n}{2n+1}, with equality only for n=0. This bound provides scale for the growth of the central coefficient. Another important relation is the of central binomial coefficients, \sum_{k=0}^n \binom{2k}{k} \binom{2(n-k)}{n-k} = 4^n, which follows from the Cauchy product of generating functions, as the square of \sum \binom{2k}{k} x^k = \frac{1}{\sqrt{1-4x}} yields \frac{1}{1-4x} = \sum 4^n x^n.

Divisibility and Prime Factors

The p-adic valuation v_p\left(\binom{2n}{n}\right), which gives the highest power of a prime p dividing the central binomial coefficient, can be determined using Kummer's theorem. This theorem states that v_p\left(\binom{2n}{n}\right) equals the number of carries that occur when adding n + n in base p. Equivalently, by de Polignac's (Legendre's) formula for the valuation of factorials, v_p\left(\binom{2n}{n}\right) = \frac{2s_p(n) - s_p(2n)}{p-1}, where s_p(m) denotes the sum of the digits of m in base p. This expression counts the carries implicitly, as each carry reduces the total digit sum by a multiple of p-1. For the specific case p=2, the formula simplifies to v_2\left(\binom{2n}{n}\right) = s_2(n), where s_2(n) is the number of $1s in the binary expansion of n. This follows because &#36;2n in binary is the bits of n shifted left by one position (appending a $0), so s_2(2n) = s_2(n), yielding v_2\left(\binom{2n}{n}\right) = s_2(n). To see this via Lucas' theorem generalized to higher powers, note that the &#36;2-adic valuation arises from the iterative application of the mod-$2case combined with lifting; however, the digit-sum formula provides the exact exponent directly. For example, whenn=7 (binary &#36;111, so s_2(7)=3), \binom{14}{7}=3432=2^3 \times 429, confirming v_2=3. When n=8 (binary $1000, s_2(8)=1), \binom{16}{8}=12870=2^1 \times 6435, confirming v_2=1$. For odd primes p, the valuation v_p\left(\binom{2n}{n}\right) = \frac{2s_p(n) - s_p(2n)}{p-1} depends on carries when doubling n in base p. For instance, take p=3 and n=5 (base-$3: 12_3, s_3(5)=3; 2n=10=101_3, s_3(10)=2). Then v_3\left(\binom{10}{5}\right) = \frac{2 \cdot 3 - 2}{2} = 2, and indeed 252 = 3^2 \times 28. Another example: p=5, n=3 (base-&#36;5: $3_5, s_5(3)=3; $2n=6=11_5, s_5(6)=2), so v_5\left(\binom{6}{3}\right) = \frac{6-2}{4} = 1, and $20=5^1 \times 4. Regarding square-freeness, Erdős and Graham conjectured that \binom{2n}{n} is square-free (not divisible by p^2 for any prime p) only for small n, specifically n=1 ($2), n=2 ($6), and n=4 ($70). This was proved in by and Olivier Ramaré, who showed that for n>4, \binom{2n}{n} is always divisible by p^2 for some prime p. Their proof uses explicit bounds on exponential sums to demonstrate the scarcity of square-free coefficients, with implications for the distribution of prime powers in binomial coefficients and related Diophantine problems in . For n=3, \binom{6}{3}=20=2^2 \times 5 provides an early counterexample beyond the square-free cases, while for n=5, $252=2^2 \times 3^2 \times 7 is divisible by multiple squares.

Analytic Developments

Generating Functions

The ordinary generating function for the central binomial coefficients \binom{2n}{n} is given by \sum_{n=0}^\infty \binom{2n}{n} x^n = \frac{1}{\sqrt{1-4x}} for |x| < 1/4. This form arises from the generalized binomial theorem applied to the expansion of (1 + y)^\alpha, where \alpha = -1/2 and y = -4x. Specifically, the generalized binomial coefficient is \binom{-1/2}{n} = \frac{(-1/2)(-3/2)\cdots(-1/2 - n + 1)}{n!} = (-1)^n \frac{(2n-1)!!}{2^n n!}, and substituting y = -4x yields \binom{-1/2}{n} (-4x)^n = \binom{2n}{n} x^n, confirming the series expansion. The radius of convergence $1/4 follows from the nearest singularity of the right-hand side at x = 1/4. The exponential generating function is \sum_{n=0}^\infty \binom{2n}{n} \frac{x^n}{n!} = e^{2x} I_0(2x), where I_0(z) denotes the modified Bessel function of the first kind, defined by the series I_0(z) = \sum_{m=0}^\infty \frac{1}{(m!)^2} \left( \frac{z}{2} \right)^{2m}. This relation follows from the bivariate exponential generating function \sum_{r,s \geq 0} \binom{r+s}{r} \frac{x^r y^s}{r! s!} = e^{x+y} I_0(2 \sqrt{xy}); setting x = y extracts the central terms via the diagonal. The modified Bessel function I_0 satisfies the differential equation z^2 w'' + z w' - (z^2 + 0^2) w = 0 and appears in solutions to problems in heat conduction and wave propagation. The generating function for the squares of the central binomial coefficients is \sum_{n=0}^\infty \binom{2n}{n}^2 x^n = \frac{2}{\pi} K(4 \sqrt{x}), valid for |x| < 1/16, where K(k) = \int_0^{\pi/2} (1 - k^2 \sin^2 \theta)^{-1/2} \, d\theta is the complete elliptic integral of the first kind. This expression derives from the hypergeometric representation \sum_{n=0}^\infty \binom{2n}{n}^2 x^n = {}_2F_1(1/2, 1/2; 1; 16x) and the known identity K(k) = \frac{\pi}{2} {}_2F_1(1/2, 1/2; 1; k^2), with the argument k = 4 \sqrt{x} matching the parameter $16x = k^2. The complete elliptic integral K(k) has a branch point at k=1 and is real-valued for $0 \leq k < 1, with applications in computing arc lengths of ellipses. These generating functions facilitate series expansions in analysis, such as Taylor series for algebraic and transcendental functions involving square roots and elliptic forms, and appear in integral representations for evaluating definite integrals over special functions. For instance, the ordinary generating function enables closed-form sums in probabilistic models like random walks, while the elliptic form for squares aids in deriving identities for lattice path counts and modular forms.

Asymptotic Approximations

The leading asymptotic approximation for the central binomial coefficient \binom{2n}{n} is obtained by applying Stirling's formula, which states that n! \sim \sqrt{2\pi n} (n/e)^n as n \to \infty. To derive this, substitute the approximation into the factorial expression for the binomial coefficient: \binom{2n}{n} = \frac{(2n)!}{(n!)^2}. Using Stirling's formula yields \frac{\sqrt{4\pi n} (2n/e)^{2n}}{2\pi n (n/e)^{2n}} = \frac{4^n}{\sqrt{\pi n}}, so \binom{2n}{n} \sim \frac{4^n}{\sqrt{\pi n}} as n \to \infty. A refined asymptotic expansion provides higher-order terms beyond the leading approximation. The full series is \binom{2n}{n} = \frac{4^n}{\sqrt{\pi n}} \left(1 - \frac{1}{8n} + \frac{1}{128n^2} + O\left(\frac{1}{n^3}\right)\right), where the error term O(1/n^3) ensures the remainder is bounded by a constant multiple of $1/n^3 for sufficiently large n. This expansion arises from the asymptotic series for the logarithm of the gamma function, \ln \Gamma(z+1) = (z + 1/2)\ln z - z + \frac{1}{2}\ln(2\pi) + \sum_{k=1}^m \frac{B_{2k}}{2k(2k-1)z^{2k-1}} + O(1/z^{2m+1}), applied to the ratio of gamma functions expressing the binomial coefficient, with Bernoulli numbers B_{2k} generating the coefficients. Strict bounds sharpen the approximation with explicit inequalities. For all positive integers n, \frac{4^n}{\sqrt{\pi(n + 1/2)}} < \binom{2n}{n} < \frac{4^n}{\sqrt{\pi n}}. These are proved using the integral representation \binom{2n}{n} = \frac{2 \cdot 4^n}{\pi} \int_0^{\pi/2} \cos^{2n} \theta \, d\theta and properties of convex functions or continued fraction expansions for the error in Stirling's formula. The asymptotic behavior also connects to the local central limit theorem via the de Moivre-Laplace theorem, which approximates the binomial distribution \mathrm{Bin}(2n, 1/2) near its mean n. Specifically, P(X = n) = \binom{2n}{n}/4^n \sim 1/\sqrt{\pi n} as n \to \infty, implying the same leading approximation for the central binomial coefficient through normalization by the variance \sigma^2 = n/2.

Connections and Applications

Relation to Catalan Numbers

The Catalan numbers C_n are defined for nonnegative integers n by the explicit formula C_n = \frac{1}{n+1} \binom{2n}{n}, which expresses them as a normalized version of the central binomial coefficient \binom{2n}{n}. This relation highlights how Catalan numbers refine the unrestricted counting provided by the central binomial coefficient. A combinatorial justification for the formula arises from the reflection principle applied to Dyck paths, which are lattice paths from (0,0) to (2n,0) using up steps (1,1) and down steps (1,-1) that never descend below the x-axis; the total number of unrestricted paths of this form is \binom{2n}{n}, and the reflection principle subtracts the invalid paths that do go below to yield exactly C_n valid Dyck paths. Equivalently, C_n counts Dyck words of length $2n, which correspond to properly balanced parentheses strings with n pairs. The ordinary generating function for the Catalan numbers, C(x) = \sum_{n=0}^\infty C_n x^n = \frac{1 - \sqrt{1 - 4x}}{2x} (with the understanding that C(0) = 1 and the constant term is handled appropriately), can be derived algebraically from the generating function for the , \sum_{n=0}^\infty \binom{2n}{n} x^n = \frac{1}{\sqrt{1 - 4x}}. Using the integral representation or direct manipulation of the coefficient formula C_n = \frac{1}{n+1} \binom{2n}{n}, the generating function C(x) emerges as the solution to the quadratic equation C(x) = 1 + x C(x)^2 arising from the Catalan recurrence, with the square root form linking back to the binomial expansion. This connection underscores unique combinatorial properties of the Catalan numbers, which count restricted structures in contrast to the unrestricted paths tallied by \binom{2n}{n}. For instance, C_n enumerates the number of full binary trees with n+1 leaves (or n internal nodes), the number of non-crossing partitions of a set of n elements, and the number of triangulations of a convex (n+2)-gon using non-intersecting diagonals. These interpretations emphasize "non-crossing" or "balanced" configurations, refining the total \binom{2n}{n} lattice paths from (0,0) to (n,n) by excluding those that cross above the main diagonal. Asymptotically, the Catalan numbers satisfy C_n \sim \frac{4^n}{n^{3/2} \sqrt{\pi}}, which follows from applying Stirling's approximation to the central binomial coefficient \binom{2n}{n} \sim \frac{4^n}{\sqrt{\pi n}} and incorporating the \frac{1}{n+1} factor, yielding the extra n^{-1/2} decay. This adjustment reflects the stricter constraints in Catalan structures compared to the binomial case. The sequence was first expressed in its modern form by the Belgian mathematician in 1838, who derived C_n while solving the problem of dissecting convex polygons into triangles, building on earlier work with central binomial coefficients dating back to Euler in the 18th century. The name "Catalan numbers" was later formalized in the mid-20th century. In number theory, central binomial coefficients play a pivotal role in Paul Erdős's 1932 elementary proof of Bertrand's postulate, which states that there is always at least one prime number between n and $2n for any integer n > 1. The argument proceeds by assuming, for contradiction, that no such prime exists for some n \geq 1, and then bounding the central binomial coefficient \binom{2n}{n} from above using the prime factorization theorem and properties of binomial coefficients. Specifically, \binom{2n}{n} can be expressed as a product involving primes up to $2n, but under the assumption, all prime factors are at most n, leading to an upper bound of \prod_{p \leq n} p^{\log_ p (2n) + O(1)}, which grows slower than the known lower bound for \binom{2n}{n} derived from its position as the largest term in (1+1)^{2n} = 2^{2n}. This contradiction implies the existence of a prime in (n, 2n). In , central binomial coefficients appear in s that connect to the , particularly through Apéry-like identities expressing values of \zeta(2n+2) as accelerated series involving these coefficients and hypergeometric terms. For instance, experimental and theoretical work has identified s where \sum \frac{\binom{2n}{n}^2 x^n}{ (n+1)^{2k+1} } or similar forms yield closed evaluations for even values, aiding in and modular form connections. These relations stem from the ordinary \sum_{n=0}^\infty \binom{2n}{n} x^n = \frac{1}{\sqrt{1-4x}}, which integrates into broader hypergeometric structures evaluated at points. Central binomial coefficients find significant applications in probability and statistics, notably in modeling simple symmetric random walks on the integers, where \binom{2n}{n} counts the number of paths of length $2n that return to the origin, yielding the return probability \binom{2n}{n} / 4^n. This discrete structure underlies the convergence of rescaled random walks to , with the central term providing the dominant contribution to the local limit theorem for the position distribution. In the continuous limit, hitting times for —such as the first passage time to a level—inherit approximations from binomial tail sums, linking discrete to processes. Beyond direct applications, central binomial coefficients relate to broader combinatorial sequences such as Motzkin numbers and Schröder numbers, which generalize the path-counting interpretations underlying (themselves \frac{1}{[n+1](/page/N+1)} \binom{2n}{n}). Motzkin numbers count paths from (0,0) to (n,0) with steps (1,1), (1,-1), and (1,0) not going below the x-axis, incorporating the central as a base case when level steps are forbidden; Schröder numbers extend this to allow additional diagonal steps, yielding generating functions that refine the \frac{1}{\sqrt{1-4x}} form. Identities involving squares of central binomial coefficients further link to enumerative problems in these sequences.

References

  1. [1]
    A000984 - OEIS
    ### Summary of A000984 (Central Binomial Coefficients)
  2. [2]
    [PDF] On Some Series Involving the Central Binomial Coefficients - arXiv
    May 16, 2025 · The central binomial coefficients have long been a cornerstone in combinatorics and mathematical analysis due to their rich combinatorial ...
  3. [3]
    [PDF] Asymptotic Expansions of Central Binomial Coefficients and Catalan ...
    The central binomial coefficient has a well-known asymptotic approximation; see e.g., [3, p. 35]:. 2nn ∼ 22n. √nπ 1 −. 1. 8n. +. 1. 128n2. +. 5. 1024n3 −. 21.
  4. [4]
    Central Binomial Coefficient -- from Wolfram MathWorld
    The nth central binomial coefficient is defined as (2n; n) = ((2n)!)/((n!)^2) (1) = (2^n(2n-1)!!)/(n!), (2) where (n; k) is a binomial coefficient, ...
  5. [5]
    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.
  6. [6]
    Euler - Introduction to Analysis of the Infinite
    Many kinds of functions whose characteristic qualities are discovered by higher analysis are classified. First I have distinguished between algebraic and.
  7. [7]
  8. [8]
    [PDF] 1957-feller-anintroductiontoprobabilitytheoryanditsapplications-1.pdf
    ... Binomial Coefficients .............. 9, Stirling's Formula ... ballot theorem. The following amusing proposition was proved in 1878 by ...
  9. [9]
    [PDF] Applications of the q-Binomial Coefficients to Counting Problems
    The binomial coefficients are one of the essential building blocks of enumer- ative combinatorics. A great deal of research has gone into understanding.Missing: central | Show results with:central
  10. [10]
    Binomial Distribution - Online Statistics Book
    For the coin flip example, N = 2 and π = 0.5. The formula for the binomial distribution is shown below: where P(x) is the probability of x successes out of N ...
  11. [11]
    [PDF] Computing the Continuous Discretely - matthias beck
    Jun 19, 2020 · The mathematical interplay between polytopes and lattices comes to life when we study the relationships between the continuous volume of a ...
  12. [12]
    BinomialCoefficients
    Pascal's identity: algebraic proof. Vandermonde's identity. Combinatorial proof; Algebraic proof. Sums of binomial coefficients; Application: the inclusion ...
  13. [13]
    [PDF] Math 432 Lec 03 Binomial Formula, Combinatorial Identities, and ...
    A sum in an identity suggests that we count something in cases. So the first method is to count the n-subsets of [2n] in cases. The key is to come.
  14. [14]
    The Combinatorics of Kummer's Theorem on Binomial Coefficients
    We present novel combinatorial proofs for some cases of a classic result of Kummer, that the highest power of p p dividing the binomial coefficient n n ...
  15. [15]
    [0811.2028] The p-adic valuation of k-central binomial coefficients
    Nov 13, 2008 · The coefficients c(n,k) defined by (1-k^2x)^(-1/k) = sum c(n,k) x^n reduce to the central binomial coefficients for k=2.
  16. [16]
    [PDF] An Introduction to Ordinary Generating Functions - garsia at york
    The expression C(x,y) is a generating function for the binomial coefficients which are not just indexed by a single integer, but by a pair of integers. To see ...
  17. [17]
    [PDF] Generating functions, UMN Math 4707, Spr. 2020
    Feb 9, 2020 · xk = (1 + x)n (by the binomial theorem). Generalizing this, for ... xk is the central binomial coefficient generating function. 7. Use ...
  18. [18]
    [PDF] arXiv:2402.17654v1 [math.CO] 27 Feb 2024
    Feb 27, 2024 · We note that the connection between central binomial coef- ficients and Bessel functions was previously known [Slo] and can be deduced directly.
  19. [19]
    [PDF] Apéry-Type Series and Colored Multiple Zeta Values - arXiv
    Apr 12, 2022 · of the complete elliptic integral ... some Apéry-type series involving squares of central binomial coefficients in terms of colored MZVs.
  20. [20]
    [PDF] generatingfunctionology - Penn Math
    May 21, 1992 · By the exponential formula (3.4.4), the exponential generating function of ... Series[(1-Sqrt[1-4x])/(2x),{x,0,9}] and you will see. 1 + x + 2 x2 ...
  21. [21]
    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.
  22. [22]
    A Remark on Stirling's Formula - jstor
    Ficken, University of Tennessee,. Knoxville 16, Tenn. A REMARK ON STIRLING'S FORMULA. HERBERT ROBBINS, Columbia University. We shall prove Stirling's formula ...
  23. [23]
    [PDF] Catalan Numbers - MIT Mathematics
    Cn of triangulations of a convex (n + 2)-gon. (definition of Catalan numbers). ... Eugène Charles Catalan (1838): wrote Cn in the form. (2n)! n!(n+1)! and ...
  24. [24]
    [PDF] Lecture XXIII: Catalan numbers, Dyck paths & Catalan recurrences
    Thm: The number of Dyck paths of length. Proof : [ Anch's Reflection Method] length 2n is Cn. nE. • Pick all paths in grid with nN & NE steps. = do not cross.
  25. [25]
    [PDF] Topics in trees and Catalan numbers - UCSD Math
    On homework, you computed the binomial series for. √. 1 + x: (1 + x). 1 ... Generating function for Catalan numbers. Second derivation of generating function.
  26. [26]
    Catalan Number -- from Wolfram MathWorld
    (15). (OEIS A000108), exponential generating function. e^(2x)[I_0(2x)-I_1(2x)] ... Central Binomial Coefficient, Delannoy Number, Dyck Path, Euler's Polygon ...
  27. [27]
    [PDF] History of Catalan numbers - UCLA Mathematics
    Euler defines Catalan numbers Cn as the number of triangulations of (n + 2)-gon, and gives the values of Cn for n ≤ 8 (evidently, computed by hand). All these ...
  28. [28]
    [PDF] Erd˝os's proof of Bertrand's postulate - University of Notre Dame
    May 1, 2015 · Abstract. In 1845 Bertrand postulated that there is always a prime between n and 2n, and he verified this for n < 3 × 106.Missing: central | Show results with:central
  29. [29]
    [PDF] Random Walk and the Theory of Brownian Motion - Mark Kac
    Feb 6, 2006 · and the right member can be expressed in terms of binomial coefficients. This formula can also be derived in a much simpler way using, for ...Missing: central | Show results with:central
  30. [30]
    [PDF] arXiv:2503.20852v1 [stat.OT] 26 Mar 2025
    Mar 26, 2025 · The central limit theorem specialised to binomial random variables Xn,p with the indicated parameters says that the distribution functions ...
  31. [31]
    (PDF) A primality test for $Kp^n+1$ numbers - ResearchGate
    Aug 9, 2025 · In this paper we generalize the classical Proth's theorem for integers of the form N = K p n + 1 N=Kp^n+1 . For these families, we present a ...Missing: central post-
  32. [32]
    Optimizing Quantum Search with a Binomial Version of Grover's ...
    Jul 21, 2020 · The partitioning process is based on the binomial distribution. If the classes to which the search target states belong are known in advance, ...<|control11|><|separator|>
  33. [33]
    [PDF] Inverses of Motzkin and Schr\" oder Paths
    The Grand Catalan numbers are the Central Binomial coefficients, 2n n , with generating function 1//1 - 4t2 = Pn≥0. 2n n t2n. The wheighted Grand Motzkin.
  34. [34]
    [PDF] Identities for squared central binomial coefficients - arXiv
    It is interesting that similar identities exist for the squared central binomial coefficients. Here we prove several such identities, where (3) especially ...