Fact-checked by Grok 2 weeks ago

Gamma function

The gamma function, denoted by the Greek letter \Gamma (gamma), is a special function in mathematics that extends the factorial z \mapsto z! to non-integer values of z with positive real part, defined by the \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt for \operatorname{Re}(z) > 0. For positive integers z, it satisfies \Gamma(z) = (z-1)!, providing a continuous of the , but with a shift of 1. Introduced by Leonhard Euler in the early as a means to generalize factorial-like products, the function was later formalized with its current notation by in 1809. The gamma function arises naturally in various mathematical contexts due to its fundamental properties. It obeys the \Gamma(z+1) = z \Gamma(z), which allows recursive computation and underpins its role as a analogue. Notable values include \Gamma(1) = 1 and \Gamma\left(\frac{1}{2}\right) = \sqrt{\pi}, the latter linking it to Gaussian integrals. The reflection formula, \Gamma(z) \Gamma(1 - z) = \frac{\pi}{\sin(\pi z)}, extends its definition meromorphically to the entire except at non-positive integers, where it has simple poles. Beyond pure mathematics, the gamma function is essential in applied fields such as probability and statistics, where it normalizes the gamma distribution—a flexible model for positive continuous random variables—and relates to the beta function via \mathrm{B}(m, n) = \int_0^1 t^{m-1} (1-t)^{n-1} \, dt = \frac{\Gamma(m) \Gamma(n)}{\Gamma(m+n)}. It also appears in physics, including quantum mechanics and statistical mechanics, for solving integral equations and computing volumes of certain spaces.

Motivation and Definition

Motivation as Factorial Extension

The factorial function for positive integers n, denoted n!, is defined as the product n! = 1 \times 2 \times \cdots \times n, representing the number of permutations of n distinct objects. This discrete definition works seamlessly for integer arguments but encounters limitations when attempting to extend it to non-integer values, such as fractions or irrational numbers, where no direct multiplicative interpretation exists. Such an extension is essential in various mathematical contexts, including the analysis of series, integrals, and differential equations involving non-integer orders, prompting the need for a continuous interpolating function that preserves the factorial's core properties. In 1729, Leonhard Euler addressed this challenge by proposing a function that analytically interpolates the , laying the groundwork for what would become known as the gamma function, denoted \Gamma(z), satisfying \Gamma(n+1) = n! for all positive integers n. Euler's motivation stemmed from his efforts to generalize expressions involving factorials, particularly to evaluate infinite series and products that arise in , such as those related to exponential expansions. This interpolation allowed for a smooth, of factorial behavior across the real numbers, enabling computations and derivations that were previously intractable with the integer-restricted definition. Euler detailed his approach in correspondence with , marking the inception of this pivotal extension in . A compelling illustration of the gamma function's utility beyond integers is the value \Gamma(1/2) = \sqrt{\pi}, which Euler computed using his interpolating product formula and which links the generalized factorial directly to the \int_{-\infty}^{\infty} e^{-x^2} \, dx = \sqrt{\pi}. This result not only validates the extension's consistency but also underscores its role in bridging combinatorial concepts with continuous probability distributions and , motivating further exploration in fields like statistics and physics where arguments frequently appear.

Integral Definition

The gamma function is fundamentally defined by Euler's integral representation for numbers z with positive real part: \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt, \quad \Re(z) > 0. This converges absolutely in the specified half-plane, ensuring the function is well-defined and analytic there. Near the origin, the integrand behaves like t^{\Re(z)-1}, which is integrable provided \Re(z) > 0; at infinity, the e^{-t} dominates the growth of t^{z-1}, guaranteeing convergence for all such z. A direct consequence of this definition is the normalization \Gamma(1) = 1, obtained by substituting z = 1 to yield \int_0^\infty e^{-t} \, dt = [-e^{-t}]_0^\infty = 1. This aligns with the function's role in extending the , where \Gamma(n+1) = n! for positive integers n, verifiable by repeated on the . The form originates from Leonhard Euler's work in the , motivated by interpolating the beyond integers. One sketch of its derivation proceeds via a limiting process: consider the auxiliary \int_0^n (1 - t/n)^n t^{z-1} \, dt, which approaches \Gamma(z) as n \to \infty since (1 - t/n)^n \to e^{-t}, and for integer z recovers the through known limits. Alternatively, it can be derived from the B(x,y) = \int_0^1 t^{x-1} (1-t)^{y-1} \, dt = \Gamma(x)\Gamma(y)/\Gamma(x+y) via the substitution t = u/(1+u), yielding the after transformation and limits.

Alternative Product Definitions

One alternative definition of the gamma function arises from Euler's representation, which extends the for positive integers to arguments. For z \in \mathbb{C} \setminus \{0, -1, -2, \dots \}, \Gamma(z) = \lim_{n \to \infty} \frac{n! \, n^z}{z(z+1) \cdots (z+n)}. This form converges uniformly on compact subsets avoiding the poles and provides a means to compute \Gamma(z) via finite approximations that improve with larger n. A canonical infinite product form, due to Weierstrass, expresses the reciprocal of the gamma function and incorporates the Euler-Mascheroni constant \gamma \approx 0.57721. Specifically, \frac{1}{\Gamma(z)} = z e^{\gamma z} \prod_{n=1}^\infty \left(1 + \frac{z}{n}\right) e^{-z/n}, valid for all complex z except the non-positive integers where \Gamma(z) has simple poles. This representation highlights the nature of $1/\Gamma(z) and facilitates derivations of other properties, such as the reflection formula. The equivalence between these product definitions and the integral form \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt (valid for \operatorname{Re}(z) > 0) follows from . For \operatorname{Re}(z) > 0, repeated on the integral yields \Gamma(z) = \frac{\Gamma(z + n + 1)}{z(z+1) \cdots (z+n)} for positive integers n, and taking the limit as n \to \infty while using for \Gamma(n+1) \approx \sqrt{2\pi n} (n/e)^n matches the Euler product. The products then define \Gamma(z) meromorphically on the entire , with simple poles at non-positive integers, extending beyond the integral's domain. These product representations are essential for the meromorphic continuation of the gamma function to \mathbb{C} \setminus \{0, -1, -2, \dots \}, as they converge everywhere except at the poles and allow uniform approximation on compact sets, enabling applications in and .

Fundamental Properties

Recurrence and Functional Equation

The of the gamma function is given by \Gamma(z+1) = z \Gamma(z) for all complex numbers z not equal to $0, -1, -2, \dots. This relation, also known as the , allows the function to be evaluated recursively by shifting the argument. It holds initially for \Re(z) > 0 where the integral definition applies and extends by to the rest of the except at the poles. To derive this equation from the integral representation \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt for \Re(z) > 0, apply . Let u = t^z and dv = t^{-1} e^{-t} \, dt, so du = z t^{z-1} \, dt and v = -e^{-t}. Then, \Gamma(z+1) = \int_0^\infty t^z e^{-t} \, dt = \left[ -t^z e^{-t} \right]_0^\infty + z \int_0^\infty t^{z-1} e^{-t} \, dt. The boundary term vanishes because t^z e^{-t} \to 0 as t \to \infty for \Re(z) > 0 and is zero at t = 0, yielding \Gamma(z+1) = z \Gamma(z). Iterating the functional equation n times for positive integer n gives the generalized recurrence \Gamma(z + n) = (z + n - 1)(z + n - 2) \cdots z \, \Gamma(z) for \Re(z) > 0. This product form expresses the gamma function at shifted arguments in terms of its value at z. For positive integers, setting z = 1 and iterating yields \Gamma(n+1) = n!, confirming the gamma function's role as an extension of the to non-integer arguments.

Particular Values

For positive integers n \geq 1, the gamma function evaluates to the of the preceding integer, \Gamma(n) = (n-1)!, with \Gamma(1) = 1. This relation follows directly from the definition and the recurrence property, extending the continuously. For example, \Gamma(2) = 1!\ = 1, \Gamma(3) = 2!\ = 2, and \Gamma(4) = 3!\ = 6. At s, the gamma function yields values involving \sqrt{\pi}. Specifically, \Gamma\left(\frac{1}{2}\right) = \sqrt{\pi}, which can be derived from the \int_0^\infty e^{-t} t^{-1/2} \, dt = \sqrt{\pi}. Using the \Gamma(z+1) = z \Gamma(z) to compute further half-integer values, \Gamma\left(\frac{3}{2}\right) = \frac{1}{2} \Gamma\left(\frac{1}{2}\right) = \frac{1}{2} \sqrt{\pi}. In general, for positive integer n, \Gamma\left(n + \frac{1}{2}\right) = \frac{(2n-1)!!}{2^n} \sqrt{\pi}, where (2n-1)!! = 1 \cdot 3 \cdot 5 \cdots (2n-1) is the double factorial of the odd number $2n-1, equivalent to \frac{(2n)!}{4^n n!} \sqrt{\pi}. This formula is obtained by iterated application of the recurrence starting from \Gamma\left(\frac{1}{2}\right). For instance, \Gamma\left(\frac{5}{2}\right) = \frac{3}{2} \Gamma\left(\frac{3}{2}\right) = \frac{3}{4} \sqrt{\pi}. For other rational arguments such as thirds and quarters, closed forms involve more advanced special functions like complete elliptic integrals. The value \Gamma\left(\frac{1}{3}\right) can be expressed as $2^{7/9} 3^{-1/12} \pi^{1/3} [K(k_3)]^{1/3}, where K(k) is the complete elliptic integral of the first kind with modulus k_3 = \frac{\sqrt{3}-1}{2\sqrt{2}}. Similarly, \Gamma\left(\frac{1}{4}\right) = 2 \pi^{1/4} [K(k_1)]^{1/2}, where k_1 = \sqrt{2} - 1. These expressions arise from connections between the beta function (related to gamma via B(x,y) = \Gamma(x) \Gamma(y) / \Gamma(x+y)) and elliptic integrals, enabling efficient numerical computation via the arithmetic-geometric mean iteration. Approximate numerical values are \Gamma\left(\frac{1}{3}\right) \approx 2.6789385342 and \Gamma\left(\frac{1}{4}\right) \approx 3.6256099082. Values at other rationals can be computed using the recurrence relation from these base points, such as \Gamma\left(\frac{1}{3} + 1\right) = \frac{1}{3} \Gamma\left(\frac{1}{3}\right), though explicit closed forms are rare beyond integers and half-integers.

Relation to Pi Function

The pi function, denoted \pi(z), provides an alternative notation for extending the factorial to non-integer values and is defined as \pi(z) = \Gamma(z+1) for \Re(z) > 0. This definition aligns directly with the factorial such that \pi(n) = n! for positive integers n, differing from the gamma function's convention \Gamma(n+1) = n! by a shift in argument. Gauss introduced this notation in his 1812 work on hypergeometric functions, where it facilitated cleaner expressions in product formulas and interpolation problems. Historically, Leonhard Euler employed concepts akin to the in his 1729–1730 investigations into infinite products, particularly when deriving the product representation for the sine function from its expansion. Euler's approach involved expressing \sin(\pi z) as \pi z \prod_{k=1}^\infty \left(1 - \frac{z^2}{k^2}\right), a formula that bridges trigonometric identities with generalizations and underpins the for \pi/2 = \prod_{k=1}^\infty \frac{4k^2}{4k^2 - 1}. This product arises from evaluating the at half-integers using gamma values, linking the pi function to classical approximations of \pi. A key identity connecting the pi function to is \sin(\pi z) = \frac{\pi z}{\Gamma(1+z) \Gamma(1-z)}, which follows from the reflection property of the gamma function and expresses the sine in terms of gamma reciprocals. This relation highlights the pi function's role in expansions, as substituting the Weierstrass form of the gamma function yields Euler's sine product directly. Applications include deriving convergence criteria for series in and evaluating definite integrals involving , such as those in . For example, the value \Gamma(1/2) = \sqrt{\pi} emerges from this framework when z = 1/2, illustrating how the encapsulates connections between factorials, trigonometric products, and the constant \pi.

Analytic Continuation and Extensions

Extension to Complex Numbers

The gamma function, defined initially by the representation for complex arguments z with \operatorname{Re}(z) > 0, can be extended by to a on the entire \mathbb{C}. This extension is achieved through representations such as the Weierstrass , which provides an explicit formula valid everywhere except at the poles. The Weierstrass product form is given by \Gamma(z) = \frac{e^{-\gamma z}}{z} \prod_{n=1}^{\infty} \left(1 + \frac{z}{n}\right)^{-1} e^{z/n}, where \gamma \approx 0.57721 is the Euler-Mascheroni constant. This infinite product converges uniformly on compact subsets of \mathbb{C} avoiding the non-positive integers, thereby defining \Gamma(z) as a with simple poles precisely at z = 0, -1, -2, \dots. The residues at these poles are \operatorname{Res}(\Gamma, -k) = (-1)^k / k! for nonnegative integers k. Within the half-plane \operatorname{Re}(z) > 0, the continued function remains holomorphic and coincides with the original integral definition \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt. Outside this region, the meromorphic nature arises from the poles, but \Gamma(z) has no zeros anywhere in \mathbb{C}. The uniqueness of this meromorphic continuation is guaranteed by the functional equation \Gamma(z+1) = z \Gamma(z), valid for all z \in \mathbb{C} except the poles, and the identity theorem for analytic functions, which ensures agreement in overlapping domains of analyticity. This characterization, together with the specified simple poles at the non-positive integers, distinguishes the gamma function as the unique such extension, leveraging its order-one growth and the specified singularities.

Poles, Residues, and Reflection Formula

The Gamma function \Gamma(z) is meromorphic in the complex plane, exhibiting simple poles at the non-positive integers z = 0, -1, -2, \dots. These poles arise from the functional equation \Gamma(z+1) = z \Gamma(z), which, when iterated, reveals singularities at these points where the denominator vanishes while the numerator remains finite. The residue of \Gamma(z) at each simple pole z = -n, where n = 0, 1, 2, \dots, is given by \operatorname{Res}(\Gamma, -n) = \frac{(-1)^n}{n!}. This result follows from the Laurent series expansion around z = -n, derived using the recurrence relation and the known values of \Gamma at positive integers. For example, at z = 0, the residue is $1; at z = -1, it is -1. A key identity connecting values of the Gamma function across the real axis is the reflection formula: \Gamma(z) \Gamma(1 - z) = \frac{\pi}{\sin(\pi z)}, valid for all z except non-positive integers. This formula, discovered by Euler, links \Gamma(z) for \Re z > 0 to its values for \Re z < 1, facilitating evaluation in regions away from the poles. One standard derivation of the reflection formula proceeds via the beta function B(z, 1 - z), defined as B(z, 1 - z) = \int_0^1 t^{z-1} (1 - t)^{-z} \, dt = \frac{\Gamma(z) \Gamma(1 - z)}{\Gamma(1)} = \Gamma(z) \Gamma(1 - z), for $0 < \Re z < 1. Substituting t = u / (1 + u) transforms the integral to \int_0^\infty \frac{u^{z-1}}{(1 + u)^{z+1 - z}} \, du = \int_0^\infty \frac{u^{z-1}}{1 + u} \, du, which evaluates to \pi / \sin(\pi z) using contour integration or known results for the beta function in the plane. An alternative derivation uses the Weierstrass infinite product representation of $1/\Gamma(z), leading to the sine product via analytic continuation. The reflection formula proves particularly useful for evaluating \Gamma(z) at negative non-integer points, where direct integral definitions fail due to divergence; by relating such values to \Gamma(1 - z) in the right half-plane, where the function is well-behaved, numerical and analytical computations become feasible.

Behavior for Negative Non-Integers

For negative non-integer real arguments z < 0, the Gamma function \Gamma(z) is defined by analytic continuation and possesses simple at each non-positive integer z = 0, -1, -2, \dots, with residues (-1)^k / k! at z = -k for nonnegative integers k. In each open interval (-n-1, -n) for n = 0, 1, 2, \dots, \Gamma(z) is real-valued, analytic, and maintains a constant throughout the interval, alternating between negative and positive across successive intervals: negative in (-1, 0), positive in (-2, -1), negative in (-3, -2), and so on. This alternating arises from the reflection formula \Gamma(z) \Gamma(1 - z) = \pi / \sin(\pi z), where \sin(\pi z) changes according to (-1)^n in the nth interval, while \Gamma(1 - z) remains positive for $1 - z > 1. Near each , \Gamma(z) diverges to +\infty or -\infty depending on the approach and the residue , creating unbounded oscillatory excursions, but \Gamma(z) approaches 0 as z \to -\infty within the intervals, with the envelope of these oscillations decreasing in amplitude. The local extrema of \Gamma(z) in these intervals occur where its derivative vanishes, \Gamma'(z) = 0. Since \Gamma'(z) = \psi(z) \Gamma(z) and \Gamma(z) \neq 0 anywhere, these points coincide with the zeros of the digamma function \psi(z) = \Gamma'(z)/\Gamma(z), the of \Gamma(z). The has exactly one simple real zero in each (-n-1, -n), leading to one extremum per : a local maximum in intervals where both pole approaches yield -\infty (negative-sign intervals) and a local minimum where both yield +\infty (positive-sign intervals). The locations x_n of these extrema satisfy \psi(x_n) = 0 and become more negative as n increases, with the values \Gamma(x_n) alternating in sign and decreasing in magnitude toward 0. Representative numerical values illustrate this behavior. For instance, \Gamma(-0.5) = -2 \sqrt{\pi} \approx -3.54491, computed via the recurrence \Gamma(z+1) = z \Gamma(z) from the known \Gamma(1/2) = \sqrt{\pi}. Similarly, \Gamma(-1.5) = (4/3) \sqrt{\pi} \approx 2.36327 > 0, obtained by applying the recurrence again: \Gamma(-0.5) = (-1.5) \Gamma(-1.5). These points lie near the first two extrema at approximately x_1 \approx -0.50408 with \Gamma(x_1) \approx -3.54464 (local maximum) and x_2 \approx -1.57349 with \Gamma(x_2) \approx 2.30240 (local minimum). The poles, alternating signs, and decaying oscillatory behavior have significant implications for the convergence of series involving \Gamma(z) at negative non-integer points. Direct inclusion of \Gamma(z) terms can lead to divergence due to the unbounded poles, necessitating or alternative representations like the $1/\Gamma(z), which is an with zeros precisely at the poles and admits globally expansions such as the Weierstrass product. This structure ensures better properties in applications like infinite products or extensions incorporating Gamma factors.

Generalizations

Incomplete Gamma Functions

The incomplete gamma functions generalize the gamma function by truncating the integral at a finite point x, either from below (lower incomplete) or from above (upper incomplete). The lower incomplete gamma function is defined as \gamma(s, x) = \int_0^x t^{s-1} e^{-t} \, dt, for \operatorname{Re}(s) > 0 and x > 0. The upper incomplete gamma function is \Gamma(s, x) = \int_x^\infty t^{s-1} e^{-t} \, dt, and they satisfy the relation \Gamma(s) = \gamma(s, x) + \Gamma(s, x). These functions are meromorphic in the complex plane with poles at non-positive integers and are essential in probability, statistics, and physics for cumulative distribution functions and error functions.

Multivariate Gamma Function

The multivariate gamma function \Gamma_p(z) is a generalization of the gamma function to multiple dimensions, particularly useful in multivariate statistics, such as the normalizing constant for the Wishart distribution. It is defined as \Gamma_p(z) = \pi^{p(p-1)/4} \prod_{i=1}^p \Gamma\left(z - \frac{i-1}{2}\right), for positive integer p and \operatorname{Re}(z) > (p-1)/2. This function extends properties of the univariate gamma to higher dimensions, facilitating computations in matrix-variate distributions.

Double Gamma Function

The double gamma function, also known as Barnes' G-function, is a higher-order generalization satisfying the functional equation G(z+1) = \Gamma(z) G(z) with G(1) = 1. It is defined via the Weierstrass infinite product or as the reciprocal of the Barnes multiple zeta function. The G-function is entire and appears in the evaluation of multiple gamma values and in number theory. Its logarithm provides a super-logarithm for the gamma function.

Multiple Gamma Function

The multiple gamma function \Gamma_n(z), introduced by Ernest William Barnes in 1904, generalizes the gamma function to n dimensions through a recurrence relation \Gamma_n(z+1) = \Gamma_{n-1}(z) \Gamma_n(z), with \Gamma_1(z) = \Gamma(z). It is defined as an infinite product regularized by multiple Hurwitz zeta functions and is used in analytic continuation of multiple zeta functions and in quantum field theory. For n=2, it corresponds to the double gamma function.

q-Gamma Function

The q-gamma function is a q-analogue of the gamma function, defined for $0 < q < 1 as \Gamma_q(z) = (1-q)^{1-z} \prod_{k=0}^\infty \frac{1 - q^{k+1}}{1 - q^{k+z}}, satisfying \Gamma_q(z+1) = \frac{1 - q^z}{1 - q} \Gamma_q(z). It generalizes factorial to q-series contexts, with applications in basic hypergeometric functions and quantum groups. Properties include q-analogues of the reflection and duplication formulas.

Other Generalizations

Other notable generalizations include the generalized gamma function in the sense of Vardi (1989), which extends to parameters for distributions, and the Fox-Wright function incorporating gamma structures in hypergeometric series. These analogues maintain functional equations similar to the gamma function but adapt to specific analytical needs in special functions theory.

Inequalities and Bounds

Key Inequalities

The logarithm of the gamma function, \log \Gamma(x), is convex for x > 0. This means that for all x, y > 0 and \lambda \in (0,1), \log \Gamma(\lambda x + (1 - \lambda) y) \leq \lambda \log \Gamma(x) + (1 - \lambda) \log \Gamma(y), with equality if and only if x = y. This log-convexity property characterizes the gamma function uniquely among functions satisfying the recurrence \Gamma(x+1) = x \Gamma(x) and \Gamma(1) = 1, as established by the . Log-convexity implies several important inequalities for ratios of gamma values. One such consequence is Gautschi's inequality, which bounds the ratio \Gamma(x+1)/\Gamma(x+s) for x > 0 and $0 < s < 1: x^{1-s} < \frac{\Gamma(x+1)}{\Gamma(x+s)} < (x+1)^{1-s}. Proved by in 1959 using properties of the and monotonicity, this inequality is particularly effective for estimating gamma ratios where the arguments differ by less than 1. offers another fundamental bound for ratios, specifically for \Gamma(x+s)/(x^s \Gamma(x)) with x > 0 and $0 < s < 1: \left( \frac{x}{x+s} \right)^{1-s} \leq \frac{\Gamma(x+s)}{x^s \Gamma(x)} \leq 1. Derived by J. G. Wendel in 1948 via Hölder's inequality applied to the integral representation of the gamma function, this provides sharp limits that approach equality as x \to \infty. Rearranging yields bounds on \Gamma(x)/\Gamma(x+a) for $0 < a < 1, such as x^{-a} \leq \Gamma(x)/\Gamma(x+a) \leq (x+a)^{1-a} / x, though the original form is often preferred for precision. These inequalities are especially valuable near integers, where x = n + f with integer n \geq 0 and $0 < f < 1. For instance, applying with s = f and x = n gives n^{1-f} < \frac{\Gamma(n+1)}{\Gamma(n+f)} < (n+1)^{1-f}, bounding \Gamma(n+f) relative to the factorial \Gamma(n+1) = n!. Such estimates facilitate numerical computations and approximations close to integer arguments without evaluating the full function.

Local Minima and Maxima

The gamma function \Gamma(x) for real x > 0 is monotonically decreasing on the (0, \alpha) and monotonically increasing on (\alpha, \infty), where \alpha \approx 1.46163214496836234126 is the unique positive at which \Gamma(x) attains its global minimum value of \Gamma(\alpha) \approx 0.8856031944108887. This local (and global) minimum occurs at the sole positive critical point of \Gamma(x), determined by the condition \Gamma'(\alpha) = 0. The critical points of \Gamma(x) are the zeros of the digamma function \psi(x) = \Gamma'(x)/\Gamma(x), which is the logarithmic derivative of the gamma function. On the positive real line, \psi(x) has exactly one zero at x = \alpha, confirming the single minimum and the associated monotonicity behavior. For large negative real x, excluding the poles at non-positive integers, \Gamma(x) exhibits oscillatory behavior with infinitely many local maxima and minima between consecutive poles. These extrema correspond to the negative zeros of \psi(x), which are all real and simple. The k-th negative zero \alpha_k (indexed from the right) satisfies the asymptotic relation \alpha_k \approx -k + O(1/\log k) for large k, implying an asymptotic density of one zero (and thus one extremum) per unit interval along the negative real axis as x \to -\infty.

Asymptotic and Series Approximations

Stirling's Approximation

Stirling's approximation provides an asymptotic expansion for the Gamma function as its argument grows large in magnitude within certain sectors of the complex plane. The leading term, originally derived by Abraham de Moivre in 1730 and refined by James Stirling in the same year, states that \Gamma(z+1) \sim \sqrt{2\pi z} \left(\frac{z}{e}\right)^z as |z| \to \infty with |\arg z| < \pi, where the notation \sim indicates that the ratio of the left side to the right side approaches 1. This formula extends the approximation for large positive integers n! to the more general Gamma function via the functional equation \Gamma(z+1) = z \Gamma(z), allowing brief shifts for computation. The full asymptotic series, developed using the and involving B_{2k}, is most conveniently expressed in logarithmic form for numerical stability: \ln \Gamma(z) = \left(z - \frac{1}{2}\right) \ln z - z + \frac{1}{2} \ln (2\pi) + \sum_{k=1}^{m} \frac{B_{2k}}{2k(2k-1) z^{2k-1}} + R_m(z), where the remainder R_m(z) satisfies R_m(z) = O(1/|z|^{2m+1}) as |z| \to \infty. Exponentiating yields the series for \Gamma(z+1) itself: \Gamma(z+1) \sim \sqrt{2\pi z} \left(\frac{z}{e}\right)^z \exp\left( \sum_{k=1}^{\infty} \frac{B_{2k}}{2k(2k-1) z^{2k-1}} \right), valid in the sector |\arg z| \leq \pi - \delta for any fixed \delta > 0. The first few terms of the sum are \frac{1}{12z} - \frac{1}{360 z^3} + \frac{1}{1260 z^5} - \cdots, corresponding to B_2 = \frac{1}{6}, B_4 = -\frac{1}{30}, and B_6 = \frac{1}{42}. One standard derivation employs Laplace's method on the integral representation \Gamma(z+1) = \int_0^\infty t^z e^{-t} \, dt for \operatorname{Re}(z) > 0. Substituting t = z s yields \Gamma(z+1) = z^{z+1} e^{-z} \int_0^\infty s^z e^{-z(s-1)} \, ds = z^{z+1} e^{-z} \int_0^\infty e^{z [\ln s - (s-1)]} \, ds. For large |z|, the integrand peaks sharply at s=1, where the phase function f(s) = \ln s - (s-1) has f(1) = 0 and f''(1) = -1. Approximating f(s) \approx -\frac{(s-1)^2}{2} near s=1 transforms the integral into a Gaussian form \int_{-\infty}^\infty e^{-z u^2 / 2} \, du \approx \sqrt{2\pi / z}, leading to the leading Stirling term after simplification. Higher-order terms in the Taylor expansion of f(s) produce the full series via successive refinements. An alternative derivation uses the \frac{\pi}{2} = \lim_{n \to \infty} \prod_{k=1}^n \frac{(2k)^2}{(2k-1)(2k+1)}, which relates to ratios of s. Expressing the product in terms of \Gamma functions or bounds shows that in the n! / [n^n e^{-n} \sqrt{n}] \to \sqrt{2\pi} as n \to \infty, confirming the leading term and providing a route to error estimates like \sqrt{2\pi n} (n/e)^n < n! < \sqrt{2\pi n} (n/e)^n e^{1/(12n)} for positive integers n. For complex z, the series converges asymptotically in sectors excluding the negative real axis, with error bounds such as |R_K(z)| \leq \frac{(1 + \zeta(K)) \Gamma(K)}{2 (2\pi)^{K+1} |z|^K} \left(1 + \min(\sec(\arg z), 2\sqrt{K})\right) for |\arg z| \leq \pi/2 and sufficiently large |z|.

Other Asymptotic Expansions

The Lanczos approximation provides a practical method for evaluating the Gamma function, particularly for complex arguments with positive real part, by combining a modified Stirling-like form with a truncated series expansion. Specifically, for \operatorname{Re}(z) > 0, \Gamma(z+1) \approx \sqrt{2\pi} \left(z + g + \frac{1}{2}\right)^{z + 1/2} e^{-(z + g + 1/2)} \sum_{k=0}^{N} \frac{c_k}{z + k + 1/2}, where g is a positive (typically chosen as g \approx 5 for double-precision accuracy), the coefficients c_k are precomputed constants depending on g and N, and the approximation achieves relative error on the order of for moderate N. This form extends effectively to the away from the negative real axis and poles, offering uniform relative accuracy better than $10^{-15} for |z| \gtrsim 1 in suitable sectors. Spouge's approximation offers another finite-sum method suitable for high-precision computation across the , excluding a small neighborhood around the poles, with explicitly bounded . The formula is given by \Gamma(z+1) = (z + a)^{z + 1/2} e^{-(z + a)} \left[ \sqrt{2\pi} + \sum_{k=1}^{a-1} \frac{(-1)^{k-1} (a - k)^{k - 1/2} e^{a - k}}{(k-1)! (z + k)} + R_a(z) \right], where a is a positive controlling (e.g., a = [14](/page/14) yields about 40 digits), and the relative error satisfies |R_a(z)/\Gamma(z+1)| < a^{-1/2} (2\pi)^{-(a + 1/2)} for \operatorname{Re}(z) > 0. This approximation is particularly advantageous for , as the error decreases exponentially with a, and it remains stable in the right half-plane. Near the poles at non-positive integers, the Gamma function exhibits simple pole behavior, with the leading asymptotic term derived from the reflection formula \Gamma(z) \Gamma(1 - z) = \pi / \sin(\pi z). For small z near 0 (the simplest pole), \Gamma(z) \sim 1 / (z \Gamma(1 - z)), since \sin(\pi z) \approx \pi z and \Gamma(1 - z) \to \Gamma(1) = 1. More generally, at z = -n + \epsilon for non-negative integer n and small \epsilon, \Gamma(z) \sim (-1)^n / (n! \epsilon), reflecting the residue (-1)^n / n! at each pole. These expansions facilitate across branch cuts and provide uniform approximations in sectors avoiding the poles. Uniform approximations in the complex plane often leverage the reflection formula to cover regions where Stirling's series diverges, such as the left half-plane. For instance, combining the reflection with a principal branch definition yields \Gamma(z) \approx \pi / (\sin(\pi z) \Gamma(1 - z)) valid uniformly for |\arg(z)| < \pi - \delta and away from poles, with higher-order terms obtainable via Laurent series expansions around each singularity. Such methods ensure relative errors bounded independently of |z| in compact subsets of the cut plane, complementing large-|z| asymptotics.

Multiple Representations

Additional Integral Forms

The Hankel contour integral provides an important representation of the Gamma function that facilitates its analytic continuation to the complex plane, particularly for regions where the standard Euler integral does not converge. One form of this representation is given by \Gamma(z) = \frac{1}{e^{2\pi i z} - 1} \int_L e^{-t} t^{z-1} \, dt, where the contour L starts at +\infty along the positive real axis, proceeds to a small circle of radius \epsilon around the origin, encircles the origin counterclockwise, and returns to +\infty along the positive real axis just below it. This contour avoids the branch point at t=0 and is valid for all complex z except non-positive integers, enabling evaluation in the left half-plane by deforming the path appropriately. An equivalent form, often used for numerical computation and continuation, expresses the reciprocal as \frac{1}{\Gamma(z)} = \frac{1}{2\pi i} \int_H (-t)^{-z} e^{-t} \, dt, with the Hankel contour H running from -\infty just below the negative real axis, circling the origin positively, and returning to -\infty just above the axis; the branch of (-t)^{-z} is defined with argument from -\pi to \pi. This integral converges for all z \in \mathbb{C} \setminus \{0, -1, -2, \dots \} and is particularly useful for asymptotic analysis and high-precision calculations. Another fundamental integral form arises from the Beta function, which relates the Gamma function to a definite integral over [0,1]. Specifically, B(x,y) = \int_0^1 t^{x-1} (1-t)^{y-1} \, dt = \frac{\Gamma(x) \Gamma(y)}{\Gamma(x+y)}, valid for \Re x > 0 and \Re y > 0. This representation is essential for expressing products of Gamma functions in terms of a single integral, with applications in probability distributions (e.g., ) and the evaluation of definite integrals involving powers. The relation allows computation of \Gamma(x+y) from known values of \Gamma(x) and \Gamma(y), or vice versa, and extends to arguments under the conditions. Mellin-Barnes integrals offer a powerful framework for representing ratios or products of Gamma functions, commonly used in the theory of hypergeometric and . A prototypical example is the Barnes lemma, which states that \frac{1}{2\pi i} \int_{-i\infty}^{i\infty} \Gamma(a+s) \Gamma(b+s) \Gamma(c-s) \Gamma(d-s) \, ds = \frac{\Gamma(a+c) \Gamma(a+d) \Gamma(b+c) \Gamma(b+d)}{\Gamma(a+b+c+d)}, where the vertical contour separates the poles of the Gamma functions in the integrand, with appropriate conditions on a,b,c,d to ensure convergence (typically \Re(a+b+c+d) > 0 and the contour chosen such that poles of \Gamma(a+s), \Gamma(b+s) lie to the left and those of \Gamma(c-s), \Gamma(d-s) to the right). This identity equates a contour integral of a product of four Gammas to a product of four Gammas in the numerator over one in the denominator, serving as a cornerstone for deriving summation formulas and asymptotic expansions in multiple Gamma products. More general Mellin-Barnes forms involve products of multiple Gammas multiplied by a power z^{-s}, enabling representations of functions like the hypergeometric _pF_q. The Gamma function also appears prominently in transforms such as the Laplace and Fourier integrals, providing additional real-line representations. The Laplace transform connection is evident in the form \int_0^\infty e^{-zt} t^{w-1} \, dt = z^{-w} \Gamma(w), for \Re z > 0 and \Re w > 0, which generalizes the standard Euler integral and is used in solving differential equations and asymptotic analysis. For Fourier transforms, cosine and sine variants yield \Gamma(z) \cos\left(\frac{\pi z}{2}\right) = \int_0^\infty t^{z-1} \cos t \, dt, \quad 0 < \Re z < 1, and \Gamma(z) \sin\left(\frac{\pi z}{2}\right) = \int_0^\infty t^{z-1} \sin t \, dt, \quad -1 < \Re z < 1. These integrals link the Gamma function to , with applications in diffraction theory and signal processing, where the oscillatory nature allows extraction of Gamma values from transform inverses.

Series and Continued Fraction Representations

The Gamma function admits various series expansions that facilitate numerical computation and analytic study, particularly around integer points or through connections to hypergeometric functions. Near a positive integer n, the Gamma function is analytic, and its Taylor series expansion around z = n can be derived using the functional equation and the known Maclaurin series for \log \Gamma(1 + z), which is \log \Gamma(1 + z) = -\gamma z + \sum_{m=2}^\infty (-1)^m \frac{\zeta(m)}{m} z^m. Additionally, ratios involving the Gamma function can be expressed using generalized hypergeometric functions; for instance, \frac{\Gamma(z+a)}{\Gamma(z+b)} = z^{a-b} \, _2F_1(a-b,1;z+1;1/z) for large |z|, with the power series expansion of the hypergeometric term providing a local representation around integer values of z. A notable Fourier series representation arises for the logarithm of the , particularly useful for x \in (-1,1). The expansion for \log \Gamma(1+x) on this interval leverages the reflection formula and periodic properties, yielding \log \Gamma(1+x) = -\gamma x + \sum_{k=1}^{\infty} \left( \frac{x}{k} - \log\left(1 + \frac{x}{k}\right) \right). This series converges uniformly on compact subsets of (-1,1) and aids in evaluating integrals involving log-gamma. Continued fraction representations provide efficient approximations for ratios of Gamma functions, often derived from integral forms like Raabe's formula, which expresses \log \Gamma(z+1) = z \log z - z + \frac{1}{2} \log(2\pi z) + \int_0^\infty \left( \frac{1}{2} - \frac{1}{t} + \frac{1}{e^t - 1} \right) \frac{e^{-z t}}{t} dt. For the specific ratio \frac{\Gamma(z+1)}{\Gamma(z+a)}, a continued fraction expansion is \frac{\Gamma(z+1)}{\Gamma(z+a)} = \frac{1}{z+a-1 - \frac{(a-1)(1-a)}{2z + a - \frac{(a-1)(2-a)}{3z + a + \ddots}}}, obtainable via transformation of Raabe's integral into a Stieltjes-type fraction, converging rapidly for \Re(z) > 0 and a > 0. These fractions are particularly valuable for and high-precision computation. The Barnes G-function extends the Gamma function to a multiple analogue, generalizing the to G(z+1) = \Gamma(z) G(z) with G(1) = 1, representing a double Gamma function G(z) = 1 / \Gamma_2(z) where \Gamma_2 is the Barnes double Gamma. This extension incorporates higher-order zeta regularization in its Weierstrass product form, \log G(z) = (z-1) \log \Gamma(z) - z(z-1)/2 + \sum_{k=1}^{\infty} \left[ \log \Gamma(k+1) - \log(k+z) \right], and finds applications in multiple gamma settings for Barnes integrals and zeta function multiple extensions.

Log-Gamma and Derivative Functions

Log-Gamma Definition and Properties

The logarithm of the gamma function, denoted \ln \Gamma(z) or \operatorname{Ln} \Gamma(z), is defined as the branch of the applied to \Gamma(z), given by \operatorname{Ln} \Gamma(z) = \ln|\Gamma(z)| + i \operatorname{ph} \Gamma(z), where \operatorname{ph} \Gamma(z) is the principal phase with -\pi < \operatorname{ph} \Gamma(z) \leq \pi. This definition ensures analytic continuation across the complex plane excluding the non-positive integers, where \Gamma(z) has poles, and provides numerical stability for computations involving large values of \Gamma(z) by avoiding overflow. A key functional property follows directly from the recurrence relation of the gamma function: \operatorname{Ln} \Gamma(z+1) = \operatorname{Ln} z + \operatorname{Ln} \Gamma(z) for \Re z > 0, with to other regions. For real arguments x > 0, \ln \Gamma(x) is a , as established by the Bohr–Mollerup theorem, which uniquely characterizes \Gamma(x) among positive functions satisfying the and normalization at x=1. This convexity implies that \ln \Gamma(x) lies above its tangents, aiding in bounds and approximations for positive reals. The derivative of the log-gamma function introduces the digamma function: \psi(z) = \frac{d}{dz} \operatorname{Ln} \Gamma(z) = \frac{\Gamma'(z)}{\Gamma(z)}, which captures the relative rate of change of \Gamma(z). This relation is fundamental for studying the local behavior and zeros of the gamma function. For large |z| in the sector |\operatorname{ph} z| \leq \pi - \delta with \delta > 0, the log-gamma function admits Stirling's asymptotic approximation: \operatorname{Ln} \Gamma(z) \sim \left(z - \frac{1}{2}\right) \operatorname{Ln} z - z + \frac{1}{2} \operatorname{Ln}(2\pi). This leading-order expansion provides essential insight into the growth of \Gamma(z) and underpins higher-order refinements using numbers.

Digamma and Polygamma Functions

The , denoted \psi(z), arises as the first derivative of the logarithm of the Gamma function, providing a key tool for analyzing its local behavior and generalizations of series. It is defined for complex z not equal to zero or negative integers by the series representation \psi(z+1) = -\gamma + \sum_{n=1}^{\infty} \left[ \frac{1}{n} - \frac{1}{n+z} \right], where \gamma \approx 0.57721 is the Euler-Mascheroni constant, defined as the limit \gamma = \lim_{m \to \infty} \left( \sum_{k=1}^m \frac{1}{k} - \ln m \right). This expression converges for \operatorname{Re}(z) > -1 and extends analytically to the rest of the except at the poles. For positive integers n, the generalizes the numbers H_n = \sum_{k=1}^n \frac{1}{k}, satisfying \psi(n+1) = -\gamma + H_n. This connection positions \psi(z) as a continuous extension of sums, enabling the of infinite series and differences that mimic partial sums, such as \psi(a) - \psi(b) for non-integer arguments. The polygamma functions extend this framework as higher-order derivatives of the , defined for m \geq 1 by \psi^{(m)}(z) = \frac{d^m}{dz^m} \psi(z). These functions capture successive differentiations of \ln \Gamma(z), with \psi^{(1)}(z) known as the and higher orders as tetragamma, pentagamma, and so on. They are meromorphic, with poles at non-positive integers, and play roles in representing higher-order generalizations and in the analysis of series expansions involving zeta functions. Key properties of the digamma function include the recurrence relation \psi(z+1) = \psi(z) + \frac{1}{z}, which allows shifting arguments and facilitates computations for larger z. Additionally, the reflection formula \psi(1-z) - \psi(z) = \pi \cot(\pi z) holds for z not an integer, linking values across the complex plane and aiding in symmetry-based evaluations. Similar recurrences extend to polygamma functions, such as \psi^{(m)}(z+1) = \psi^{(m)}(z) + (-1)^m m! z^{-m-1}. In applications, the and polygamma functions are essential for summing series in , such as expressing generalized harmonic numbers H_n^{(s)} = \sum_{k=1}^n \frac{1}{k^s} via limits involving \psi^{(s-1)}(z), and evaluating finite differences or integrals that reduce to polygamma terms. For instance, they appear in the closed-form of alternating or weighted harmonic-like series, providing exact representations where direct is intractable.

Computation and Implementation

Numerical Evaluation Methods

Numerical evaluation of the Gamma function requires robust algorithms to handle its behavior across real and arguments, ensuring accuracy and efficiency while managing singularities and large magnitudes. A key preliminary step in these computations is argument reduction, which leverages the functional \Gamma(z+1) = z \Gamma(z) to map the input argument z into a principal strip, typically where $1 \leq \operatorname{Re}(z) \leq 2 or $0.5 \leq \operatorname{Re}(z) \leq 1.5, where approximation methods converge rapidly. This reduction minimizes the number of iterations needed for high accuracy and is essential for both real and inputs, as repeated application of the recurrence shifts the argument toward the origin while multiplying by accumulated factors. Within the principal range, the provides an effective method for computing \Gamma(z) for positive real parts, extending naturally to complex arguments with real coefficients. Introduced by , this approach uses a truncated that achieves high relative accuracy, often better than other global methods for the same number of terms, though rigorous error bounds remain empirical rather than analytically proven. Similarly, Spouge's approximation offers a global method suitable for real and complex arguments, employing a series with precomputed coefficients that yields a relative error bounded by \sqrt{r} (2\pi)^{r+1/2} / \operatorname{Re}(z+r) for parameter r \geq 3 and \operatorname{Re}(z+r) > 0, where choosing r \approx 1.1 d ensures relative error less than $10^{-d} for d decimal digits. Both methods require approximately $0.377d terms for d digits of and are particularly efficient when coefficients are cached, outperforming direct integral evaluations for moderate to high precision. The Gamma function has simple poles at non-positive integers z = 0, -1, -2, \dots, where numerical evaluations typically return to indicate divergence, while nearby points are computed using the reflection formula \Gamma(z) \Gamma(1-z) = \pi / \sin(\pi z) to avoid direct pole encounters. In specialized contexts, such as residue calculus, the residue at z = -k (for non-negative k) is (-1)^k / k!, but standard algorithms prioritize detecting exact poles via integer checks on \operatorname{Re}(z) and returning or signaling an error. For high-precision requirements exceeding machine precision, libraries implement these methods using ball arithmetic to provide rigorous error bounds, adjusting working precision to account for effects like \log_2(|\log |z|| \cdot |z|). Such systems, often employing Stirling's series for very large |z| after reduction, enable computations to thousands of digits with controlled , ensuring the final result lies within a certified .

Software Libraries and Reference Tables

Several software libraries provide robust implementations for computing the Gamma function and its variants, supporting various programming languages and precision requirements. In , the library offers the scipy.special.gamma function for evaluating Γ(z) for real and complex arguments, along with gammaln for the natural logarithm of the absolute value of the Gamma function, which helps avoid overflow in computations. The Scientific Library (GSL) in C includes gsl_sf_gamma, which computes Γ(x) for real x > 0 using the method. For , the Arb , built on the GMP multiple-precision , supports high-precision evaluation of the Gamma function via arb_gamma, enabling computations with rigorous error bounds for both real and arguments. In Mathematica, the built-in Gamma[z] function provides arbitrary-precision evaluation of the Euler Gamma function for z, with extensive support for related properties and transformations. These libraries rely on established numerical evaluation methods to ensure accuracy and efficiency. Historical reference tables for the Gamma function include Egon S. Pearson's 1922 compilation of the logarithms of the complete Gamma function for arguments from 2 to 1200, providing values to 10 decimal places, which extended beyond earlier tables like Legendre's and facilitated statistical computations. Modern reference tables offer quick lookup values for the Gamma function across a range of positive real arguments. The following table lists approximate values of Γ(z) for z from 0.1 to 10.0 in increments of 0.5, rounded to 6 decimal places for :
zΓ(z)
0.19.513508
0.51.772454
1.01.000000
1.50.886227
2.01.000000
2.51.329340
3.02.000000
3.53.323351
4.06.000000
4.511.631728
5.024.000000
5.553.597396
6.0120.000000
6.5287.885281
7.0720.000000
7.52016.024767
8.05040.000000
8.514392.756640
9.040320.000000
9.5102960.772422
10.0362880.000000
These values are derived from standard computations and can be verified using the aforementioned libraries.

Applications

In Calculus and Integration

The Beta function serves as a fundamental tool in evaluating definite integrals over finite intervals, particularly those of the form B(m,n) = \int_0^1 t^{m-1} (1-t)^{n-1} \, dt for \operatorname{Re}(m) > 0 and \operatorname{Re}(n) > 0. This integral equals \frac{\Gamma(m) \Gamma(n)}{\Gamma(m+n)}, providing a direct connection between the Beta function and the that facilitates the computation of such integrals when Gamma values are known or approximable. The relation originates from Euler's integral representations and is derived by expressing the Beta integral in terms of double integrals over the positive reals, which factor into products of . For parameters, this reduces to expressions involving factorials, but the Gamma formulation extends its utility to non- cases in problems like probability density normalizations or computations. A broader class of improper integrals over [0, \infty) can also be evaluated using the Gamma function through substitutions that transform them into the standard Gamma form. Specifically, the integral \int_0^\infty x^{a-1} e^{-x^b} \, dx = \frac{1}{b} \Gamma\left( \frac{a}{b} \right) holds for b > 0, a > 0, by setting u = x^b, which yields dx = \frac{1}{b} u^{\frac{1}{b} - 1} \, du and simplifies the exponent to e^{-u} u^{\frac{a}{b} - 1}. This result is pivotal in solving integrals arising in areas such as heat conduction or , where the exponent involves powers, and it generalizes the basic Gamma integral \int_0^\infty x^{a-1} e^{-x} \, dx = \Gamma(a). The Gamma function appears centrally in integral transforms that aid in solving differential equations and analyzing functions in calculus. In the Mellin transform, defined as \mathcal{M}\{f\}(s) = \int_0^\infty x^{s-1} f(x) \, dx, the transform of the exponential f(x) = e^{-x} is precisely \Gamma(s) for \operatorname{Re}(s) > 0, serving as the kernel that links multiplicative structures to additive ones. Similarly, in Fourier transforms, Gamma functions emerge in the kernels for certain representations, such as the Fourier transform of the Gamma density or in Parseval-type identities involving Gamma ratios, though the Mellin form is more directly tied to the Gamma's integral definition. These transforms enable the inversion and decomposition of functions, with applications in solving convolution integrals or boundary value problems. In multidimensional calculus, the Gamma function quantifies volumes of high-dimensional regions, notably the volume of the unit ball in \mathbb{R}^n, given by V_n(1) = \frac{\pi^{n/2}}{\Gamma\left( \frac{n}{2} + 1 \right)}. This formula arises from integrating the uniform density over the ball using hyperspherical coordinates, where the radial component integrates to a Beta function that reduces via the Gamma relation. For even dimensions, it yields rational multiples of \pi^{n/2}, while for odd dimensions, it involves square roots; the Gamma denominator ensures analytic continuation across all positive integers n, providing a unified expression essential for asymptotic analysis of volumes as n grows large.

In Number Theory and Special Functions

The \zeta(s), central to , admits an representation that explicitly involves the Gamma function for \Re(s) > 1: \zeta(s) = \frac{1}{\Gamma(s)} \int_0^\infty \frac{t^{s-1}}{e^t - 1} \, dt. This form arises from term-by-term integration of the \zeta(s) = \sum_{n=1}^\infty n^{-s} after expressing each term via the Gamma integral \frac{1}{n^s} = \frac{1}{\Gamma(s)} \int_0^\infty t^{s-1} e^{-n t} \, dt, and interchanging sum and under suitable convergence conditions. The representation facilitates and , linking the distribution of primes—via the Euler product \zeta(s) = \prod_p (1 - p^{-s})^{-1}—to properties of the Gamma function, such as its poles and formula, which appear in the \zeta(s) = 2^s \pi^{s-1} \sin(\pi s / 2) \Gamma(1-s) \zeta(1-s). In , generalizations of the , such as the Barnes multiple zeta function \zeta_m(s, \mathbf{a}) = \sum_{k_1=1}^\infty \cdots \sum_{k_m=1}^\infty (k_1 + \cdots + k_m + a_1 + \cdots + a_m)^{-s} for \Re(s) > m, play a role in the study of multiple L-functions and periods associated with prehomogeneous vector spaces. The meromorphic continuation of these functions relies on the multiple Gamma functions \Gamma_n(z), introduced by Barnes as higher-dimensional analogs of the Gamma function, satisfying the \Gamma_n(z+1) = z \Gamma_n(z) with additional properties derived from Weierstrass products regularized by multiple Hurwitz zeta functions. These multiple Gamma functions enable the of Barnes zeta functions, mirroring the role of the single Gamma in the Riemann case, and have applications in evaluating sums over lattices and in the theory of Shintani zeta functions for quadratic forms. The Gamma function also connects to other special functions in number theory through expressions involving Bessel and hypergeometric functions. For instance, the modified Bessel function of the first kind I_\nu(z) expresses as I_\nu(z) = \frac{(z/2)^\nu}{\Gamma(\nu+1)} \, _1F_1\left(\nu + \frac{1}{2}; 2\nu + 1; 2z\right), valid for |\arg z| < \pi/2 and \nu \neq 0, -1, -2, \dots, where _1F_1 is the confluent hypergeometric function; the Gamma factor normalizes the leading behavior near z=0. Similarly, the ordinary Bessel function J_\nu(z) relates via J_\nu(z) = (z/2)^\nu / \Gamma(\nu+1) \times {}_0F_1( ; \nu+1 ; -z^2/4), a generalized hypergeometric function, with the Gamma ensuring the correct asymptotic scaling. These relations appear in number-theoretic contexts, such as integral representations of zeta values or sums over arithmetic progressions involving oscillatory terms. The incomplete Gamma function \gamma(a,z) further ties to the confluent hypergeometric via \gamma(a,z) = a^{-1} z^a e^{-z} \, _1F_1(1; a+1; z), facilitating evaluations of hypergeometric series at integer parameters relevant to partition functions and modular forms. In number theory, the Gamma function generalizes primorials—the products of the first n primes, denoted p_n\#—analogous to its interpolation of factorials, through expressions like \Gamma(p\# + 1) = (p\#)!, which encodes the primorial as the argument of a factorial incorporating all primes up to p. This construction highlights the exponential growth of primorials under the , where \log(p\#) \sim p, leading to super-factorial scales in analytic estimates of prime distributions, though no unique log-convex interpolation akin to Gamma exists for primorials due to their discrete jumps.

In Physics and Statistics

The Gamma function plays a central role in the probability density function (PDF) of the Gamma distribution, a continuous probability distribution widely used in statistics to model waiting times for events in a . The PDF is given by f(x; k, \theta) = \frac{x^{k-1} e^{-x/\theta}}{\theta^k \Gamma(k)}, \quad x > 0, where k > 0 is the and \theta > 0 is the ; here, \Gamma(k) serves as the normalizing constant, generalizing the for non-integer k. For integer k = n, this reduces to the , which specifically describes the time until the nth event in a with rate $1/\theta. In statistical applications, the Gamma function appears in the moments of distributions related to the Gamma family. For instance, the with n degrees of freedom is a special case of the with n/2 and 2, so its PDF involves \Gamma(n/2) in the ; this connection is crucial for hypothesis testing and variance in samples. Additionally, in , the is the for the rate parameter of a or the of an , ensuring that the posterior remains Gamma-distributed after updating with or likelihoods. This conjugacy simplifies analytical computations for parameter in models of count data or lifetimes. In physics, particularly quantum mechanics, the Gamma function arises in the normalization of radial wave functions for the hydrogen atom. The radial part R_{n\ell}(r) of the wave function includes a normalization factor involving \Gamma(n + \ell + 1) (equivalent to (n + \ell)! for integer quantum numbers) and implicitly \Gamma(\ell + 1) = \ell! through the associated Laguerre polynomials and factorial terms in the coefficient: R_{n\ell}(r) = -\left[\frac{(n - \ell - 1)!}{2n[(n + \ell)!]^3}\right]^{1/2} \left(\frac{2}{n a_0}\right)^{\ell + 3/2} r^\ell e^{-r / (n a_0)} L_{n-\ell-1}^{2\ell+1}\left(\frac{2r}{n a_0}\right), where a_0 is the Bohr radius, ensuring the wave function is properly normalized over radial space. This structure reflects the generalization of discrete quantum states via continuous integral representations provided by the Gamma function.

Historical Development

18th Century Foundations

In the early 18th century, mathematicians including Leonhard Euler and James Stirling sought to generalize the factorial function, motivated by the need to interpolate the sequence of integer factorials to non-integer values for broader analytical applications. Euler's pioneering efforts began with his correspondence on sequence interpolation, laying the groundwork for what would become the Gamma function. In 1729, Euler developed an approach to using an representation derived from the infinite product expansion for the sine function, allowing extension to non-integer arguments. This method provided a means to define the analog for fractional values, marking the initial beyond positive integers. Euler detailed this in a letter to dated October 13, 1729, where he outlined the product form as a expression suitable for . Independently, in 1730, James Stirling formulated an approximation for n! based on infinite series expansions, offering insights into the asymptotic behavior of factorials for large n. Stirling's work appeared in his treatise Methodus differentialis: sive tractus de summatione et interpolatione serierum infinitarum, where he derived the formula through methods of series summation and differential equations, providing a practical tool for estimating large factorials that later aligned with Gamma function properties. During the 1730s, Euler advanced his further with preliminary representations for the extension, introduced in a letter to Goldbach on , 1730, which tied the product form to an over logarithmic terms. This approach offered a new perspective on the generalization, emphasizing convergence for positive real arguments. Euler and exchanged correspondence on mathematical generalizations, including series and extensions, fostering shared insights into these emerging concepts amid the broader 18th-century discourse on transcendental functions.

19th Century Formalizations

In 1811, introduced the notation \Gamma(z) for the gamma function in his three-volume treatise Exercices de calcul intégral, where he refined the integral definition originally proposed by Euler as \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt for \Re(z) > 0, emphasizing its role in evaluating definite integrals and its connection to the . This notation, with \Gamma(n) = (n-1)! for positive integers n, became the standard symbol and facilitated clearer expressions of properties like the reflection formula. Legendre's work also included tables of values and applications to elliptic integrals, solidifying the function's utility in . Two years later, in 1813, Carl Friedrich Gauss advanced the formalization by defining the pi function \pi(z) = \frac{\Gamma(z+1)}{z}, which satisfies the functional equation \pi(z+1) = z \pi(z) and extends the shifted factorial such that \pi(n) = (n-1)! for positive integers n. In his paper "Disquisitiones generales circa seriem infinitam," Gauss derived an infinite product representation for \pi(z) as \pi(z) = \lim_{n \to \infty} \frac{n! \, n^z}{z(z+1) \cdots (z+n)} and proved the multiplication theorem: \pi(nz) = n^{nz - 1/2} (2\pi)^{(n-1)/2} \prod_{k=0}^{n-1} \pi\left(z + \frac{k}{n}\right) for positive integer n, demonstrating the function's multiplicative structure over rational arguments. These developments provided rigorous proofs of uniqueness under the functional equation combined with the limit form, ensuring \pi(z) (and thus \Gamma(z)) is the sole analytic continuation satisfying the recurrence with \pi(1) = 1. In 1856, Karl Weierstrass offered a foundational representation using his theory of entire functions, expressing the reciprocal gamma as the infinite product \frac{1}{\Gamma(z)} = z e^{\gamma z} \prod_{n=1}^\infty \left(1 + \frac{z}{n}\right) e^{-z/n}, where \gamma is the Euler-Mascheroni constant, portraying $1/\Gamma(z) as a Weierstrass canonical product with zeros at the non-positive integers. This form underscored the gamma function's meromorphic nature, with simple poles at z = 0, -1, -2, \dots, and enabled proofs of uniqueness by showing that any function satisfying \Gamma(z+1) = z \Gamma(z) and analytic in \Re(z) > 0 must coincide with this product via analytic continuation and growth estimates. Weierstrass's approach integrated the gamma function into the broader framework of complex analysis, confirming its determination solely by the functional equation and normalization at a point like \Gamma(1) = 1.

20th Century Extensions and Characterizations

In 1904, Ernest William Barnes introduced the multiple gamma functions G_n(z), which extend the classical Gamma function to higher dimensions by satisfying a of order n. These functions are defined through an representation regularized using multiple Hurwitz zeta functions and provide a framework for generalizing factorial-like interpolations in several variables. Barnes' construction has proven influential in , particularly for studying multiple zeta values and residues of higher-dimensional meromorphic functions. The Bohr-Mollerup theorem, established in 1922 by and Johannes Mollerup, provides a uniqueness characterization of the Gamma function among positive functions on the positive reals that satisfy the f(x+1) = x f(x) with f(1) = 1 and exhibit log-convexity, meaning \log f is a . This theorem demonstrates that the Gamma function is the only such interpolant of the that possesses these properties, resolving questions about the of different representations of the Gamma function. The proof relies on the convexity condition to bound the function and ensure uniqueness via the functional equation. In 1931, offered an alternative characterization of the Gamma function that avoids explicit reliance on integral or representations, instead using the and a growth condition related to the difference \log f(x+1) - \log f(x). Artin's approach simplifies the Bohr-Mollerup proof and emphasizes the Gamma function's role as the unique solution satisfying f(x+1) = x f(x), f(1) = 1, and a specific asymptotic behavior for the . This characterization highlights the Gamma function's intrinsic properties without reference to its analytic definitions. Post-1950 developments extended the Gamma function into non-archimedean settings within , notably through q-analogs and p-adic variants. The p-adic Gamma function, explicitly defined by Yasuo Morita in 1975, interpolates the p-adic factorial and satisfies a p-adic analog of the , enabling applications to p-adic L-functions and measures. Building on this, Benedict Gross and Neal Koblitz in 1979 introduced a p-adic q-analog via the Gross-Koblitz formula, which relates Gauss sums to values of the p-adic Gamma at rational arguments and has been pivotal in p-adic interpolation of special values in . These extensions facilitate the study of and L-functions over p-adic fields.

References

  1. [1]
    [PDF] Introduction to the Gamma Function
    Feb 4, 2002 · The gamma function, introduced by Euler, generalizes the factorial to non-integer values and is a special transcendental function.
  2. [2]
    15.5 - The Gamma Function | STAT 414
    An Aside. The gamma function, denoted , is defined, for , by: Γ ( t ) = ∫ 0 ∞ y t − 1 e − y d y. We'll primarily use the definition in order to help us prove ...
  3. [3]
    [PDF] The Gamma Function - IITB Math
    Sep 12, 2018 · The gamma function - introduced into analysis in 1729: by Euler in a letter to Goldbach. Issue: Analytic Interpolation of the factorial.
  4. [4]
    [PDF] Gamma the function - How Euler Did It
    A nice history of the gamma function is found in a 1959 article by Philip Davis, ... "Leonhard Euler's Integral: A Historical Profile of the Gamma Function", The ...
  5. [5]
    Gamma Function -- from Wolfram MathWorld
    The (complete) gamma function Gamma(n) is defined to be an extension of the factorial to complex and real number arguments.Missing: original | Show results with:original
  6. [6]
    DLMF: §5.2 Definitions ‣ Properties ‣ Chapter 5 Gamma Function
    §5.2(iii) Pochhammer's Symbol ; 5.2.5, ( a ) n ; a ≠ 0 , − 1 , − 2 , … . ; ⓘ. Symbols: Γ ⁡ ( z ) : gamma function, ( a ) n : Pochhammer's symbol (or shifted ...
  7. [7]
    [PDF] 05. The Gamma function Γ(s) [draft]
    Nov 3, 2024 · Euler's integral for Γ(s). The Gamma function Γ(s) can be defined by Euler's integral. Γ(s) = ∫ ∞. 0 ts e−t dt t. (absolute convergence for Re(s) ...
  8. [8]
    Proof of convergence of integral representation of the Gamma ...
    Sep 10, 2020 · For Re(z)>0, we can represent the Gamma Function by the integral. Γ(z)=∫∞0e−xxz−1dx. For any integer n and x≥0, we have the bound.Euler's Gamma function and convergence of an integrallast step in gamma function derivation using dominated converge ...More results from math.stackexchange.com
  9. [9]
    None
    ### Summary of Definite Integral (Euler) Definition from Gamma Function PDF
  10. [10]
    Leonhard Euler's Integral: A Historical Profile of the Gamma Function
    Mar 12, 2018 · (1959). Leonhard Euler's Integral: A Historical Profile of the Gamma Function. The American Mathematical Monthly: Vol. 66, No. 10, pp.
  11. [11]
    Introduction to the Gamma Function
    It was introduced by the famous mathematician L. Euler (1729) as a natural extension of the factorial operation from positive integers to real and even complex ...
  12. [12]
    5.8 Infinite Products ‣ Properties ‣ Chapter 5 Gamma Function
    §5.8 Infinite Products ; z ≠ 0 , − 1 , − 2 , … , ; ⓘ. Symbols: Γ ⁡ ( z ) : gamma function, ! : factorial (as in n ! ), k : nonnegative integer and z : complex ...
  13. [13]
    [PDF] Chapter 8 Euler's Gamma function
    Then the infinite product defining h is pointwise absolutely convergent, which im- plies that h(z) 6= 0 whenever any of the factors in the product is 6= 0; that ...
  14. [14]
    [PDF] The Gamma function 1. Basics involving Γ(s), B(α, β), etc.
    Sep 15, 2019 · The Gamma function Γ(s) is defined by Euler's integral, and it relates to factorials for 0 < s ∈ Z. Γ(s) · Γ(1 - s) = π/sin πs.<|control11|><|separator|>
  15. [15]
    [PDF] 17 The functional equation
    Nov 5, 2018 · Integration by parts yields. Γ(s) = tse−t s. ∞. 0. +. 1 s ... thus the gamma function can be viewed as an extension of the factorial function.
  16. [16]
    DLMF: §5.4 Special Values and Extrema ‣ Properties ‣ Chapter 5 ...
    Contents · §5.4(i) Gamma Function · §5.4(ii) Psi Function · §5.4(iii) Extrema ...
  17. [17]
    DLMF: §5.5 Functional Relations ‣ Properties ‣ Chapter 5 Gamma ...
    §5.5(ii) Reflection ; z ≠ 0 , ± 1 , … , ; ⓘ. Symbols: Γ ⁡ ( z ) : gamma function, π : the ratio of the circumference of a circle to its diameter, sin ⁡ z : sine ...
  18. [18]
    Fast evaluation of the gamma function for small rational fractions ...
    The reverse procedure of expressing gamma functions in terms of complete elliptic integrals allows us to use the arithmetic-geometric mean iteration to ...Missing: primary source
  19. [19]
    [PDF] Euler and the multiplication formula for the Gamma Function
    The interpolation of the factorial by the r-function was found nearly simultaneously by Daniel Bernoulli (1700-1782) [Be29] and Euler in 1729 (see also his ...
  20. [20]
    [PDF] A Historical Profile of the Gamma Function - OSU Math
    The gamma function, a fundamental concept, was born in 1729 through a correspondence between Euler and Goldbach, arising from interpolation and integral ...
  21. [21]
    [PDF] METHODS OF MATHEMATICAL PHYSICS
    Derivation of Euler's reflection formula [r] Review the derivation of. Γ(z)Γ(1 − z) = Γ(1)B(z,1 − z) = ∫ 1. 0 dt tz−1(1 − t)−z = ∫ ∞. 0 dx xz−1. 1 + x.
  22. [22]
    DLMF: §5.6 Inequalities ‣ Properties ‣ Chapter 5 Gamma Function
    Γ ⁡ ( z ) : gamma function, π : the ratio of the circumference of a circle ... Γ ⁡ ( z ) : gamma function, π : the ratio of the circumference of a ...
  23. [23]
    [0904.1048] Bounds for the ratio of two gamma functions - arXiv
    Apr 7, 2009 · Title:Bounds for the ratio of two gamma functions--From Wendel's and related inequalities to logarithmically completely monotonic functions.
  24. [24]
    the principal inverse of the gamma function
    Aug 3, 2011 · THE PRINCIPAL INVERSE OF THE GAMMA FUNCTION. 1345. 2. Proof of Theorem 1. We begin with a simple fact which follows from Theorem A because log ...
  25. [25]
    Digamma Function -- from Wolfram MathWorld
    The digamma function is the logarithmic derivative of the gamma function, or the logarithmic derivative of the factorial, often denoted as psi_0(z).<|control11|><|separator|>
  26. [26]
    [PDF] A note on the zeros and local extrema of Digamma related functions
    Feb 9, 2016 · It is known that the Digamma function has only real and simple zeros, and all of them except only one are negative. with the normalization G(1) ...
  27. [27]
    [PDF] Stirling's Formula and Laplace's Method
    Dec 4, 2001 · Stirling's formula was found by Abraham de. Moivre and published in “Miscellenea Analyt- ica” 1730. It was later refined, but published.
  28. [28]
    [PDF] stirling's formula - keith conrad
    Any proof of Stirling's formula needs to bring in a formula that involves π. One such formula, which Stirling knew, is the Wallis product π. 2. = 2. 1. ·. 2. 3.
  29. [29]
    [PDF] 1.3 Stirling's Series
    Stirling's series is an approximation for the Gamma function, used when computing it for large arguments, and is a precise form when z=1.
  30. [30]
    5.11 Asymptotic Expansions ‣ Properties ‣ Chapter 5 Gamma ...
    The scaled gamma function Γ ∗ ⁡ ( z ) is defined in (5.11.3) and its main property is Γ ∗ ⁡ ( z ) ∼ 1 as z → ∞ in the sector | ph ⁡ z | ≤ π − δ.
  31. [31]
    [PDF] LAPLACE'S METHOD, FOURIER ANALYSIS, AND RANDOM ...
    Laplace's approach to Stirling's formula is noteworthy first, because it makes a direct connection with the Gaussian (normal) distribution (whereas in other ...
  32. [32]
    [PDF] Wallis' Formula and Stirling's Formula In class we used Stirling's ...
    Here, “∼” means that the ratio of the left and right hand sides will go to 1 as n → ∞. We will prove Stirling's Formula via the Wallis Product Formula.Missing: approximation | Show results with:approximation
  33. [33]
    A Precision Approximation of the Gamma Function - SIAM.org
    Jul 14, 2006 · A Precision Approximation of the Gamma Function. Author: C. LanczosAuthors Info & Affiliations. https://doi.org/10.1137/0701008.
  34. [34]
    [2005.10449] The Lanczos Approximation for the $Γ - arXiv
    May 21, 2020 · Abstract:We examined the properties of the coefficients of the \cite{lanczos1964} approximation of the \Gamma-function with complex values ...
  35. [35]
    [PDF] Gamma.pdf - Wolfram Functions
    (Hankel's contour integral.) The path of integration L starts at ¥+д 0 on the real axis, goes to Ε+д 0, circles the origin in the counterclockwise direction ...
  36. [36]
    5.9 Integral Representations ‣ Properties ‣ Chapter 5 Gamma ...
    Addition (effective with 1.0.10):: To increase the region of validity of this equation, the logarithm of the gamma function that appears on its left-hand side ...
  37. [37]
    [PDF] Computing the gamma function using contour integrals and rational ...
    Abstract. Some of the best methods for computing the gamma function are based on numerical evaluation of Hankel's contour integral.
  38. [38]
    DLMF: §5.12 Beta Function ‣ Properties ‣ Chapter 5 Gamma Function
    ### Summary of Gamma Function and Beta Function from https://dlmf.nist.gov/5.12
  39. [39]
    DLMF: §5.19 Mathematical Applications ‣ Applications ‣ Chapter 5 ...
    Many special functions f ⁡ ( z ) can be represented as a Mellin–Barnes integral, that is, an integral of a product of gamma functions, reciprocals of gamma ...
  40. [40]
    [PDF] Series expansion of the Gamma function and its reciprocal
    Dec 3, 2021 · This paper gives representations for the coefficients of the Maclaurin series for Γ(z + 1) and its reciprocal, using a differential operator D.
  41. [41]
    Fourier series representations of the logarithms of the Euler gamma ...
    Mar 25, 2009 · Kummer's Fourier series for the log gamma function is well known, having been discovered in 1847. In this paper we develop a corresponding ...
  42. [42]
    [PDF] Raabe's formula for p-adic gamma and zeta functions - Numdam
    In this paper we compute a Volkenborn integral of Log ΓD in terms of its derivative (Log ΓD)0, and show that this integral formula and the difference equation ...
  43. [43]
    5.17 Barnes' G -Function (Double Gamma Function)
    In this equation (and in (5.17.5) below), the Ln 's have their principal values on the positive real axis and are continued via continuity.Missing: multiple | Show results with:multiple
  44. [44]
    DLMF: §4.2 Definitions ‣ Logarithm, Exponential, Powers ‣ Chapter ...
    The general logarithm function Ln ⁡ z is defined by where the integration path does not intersect the origin. This is a multivalued function of z with branch ...
  45. [45]
  46. [46]
  47. [47]
    5.7 Series Expansions ‣ Properties ‣ Chapter 5 Gamma Function
    DLMF · Index · Notations; Search; Help? Citing · Customize; Annotate; UnAnnotate. About the Project · 5 Gamma FunctionProperties5.6 Inequalities5.8 Infinite ...
  48. [48]
    DLMF: §5.15 Polygamma Functions ‣ Properties ‣ Chapter 5 ...
    The functions ψ ( n ) ⁡ ( z ) , n = 1 , 2 , ... , are called the polygamma functions. In particular, ψ ′ ⁡ ( z ) is the trigamma function.
  49. [49]
    [2109.08392] Arbitrary-precision computation of the gamma function
    Sep 17, 2021 · We discuss the best methods available for computing the gamma function \Gamma(z) in arbitrary-precision arithmetic with rigorous error bounds.
  50. [50]
    Computation of the Gamma, Digamma, and Trigamma Functions
    This paper gives an approximation for the gamma function that, while different, has the same form as one by Lanczos.<|separator|>
  51. [51]
    scipy.special.gamma — SciPy v1.16.2 Manual
    The gamma function is often referred to as the generalized factorial since for natural numbers. More generally it satisfies the recurrence relation for complex.1.15.2 · 1.14.0 · 1.15.1 · 1.15.0
  52. [52]
    Arb - a C library for arbitrary-precision ball arithmetic — Arb 2.23.0 ...
    Arb is a C library for rigorous real and complex arithmetic with arbitrary precision. Arb tracks numerical errors automatically using ball arithmetic.
  53. [53]
    Euler gamma function—Wolfram Documentation
    Gamma[z] is the Euler gamma function Gamma[z]. Gamma[a, z] is the incomplete gamma function a. Gamma[a, z0, z1] is the generalized incomplete gamma function ...PolyGamma · LogGamma · QGamma
  54. [54]
    Table of the Logarithms of the Complete -function (for Arguments 2 ...
    Author, Egon Sharpe Pearson ; Publisher, Cambridge University Press, 1922 ; Original from, Pennsylvania State University ; Digitized, May 4, 2009 ; Length, 16 pages.<|separator|>
  55. [55]
    values of gamma function for small positive real values - PlanetMath
    Mar 22, 2013 · The following table lists values of Γ(x) Γ ⁢ ( x ) for real 0≤x<10 0 ≤ x < 10 in steps of 110 1 10 . If the table is correct, it should confirm ...<|control11|><|separator|>
  56. [56]
    The Gamma Function (XII) - A Course of Modern Analysis
    The Gamma Function · E. T. Whittaker, G. N. Watson; Book: A Course of Modern Analysis; Online publication: 05 September 2013; Chapter DOI: https://doi.org ...
  57. [57]
    DLMF: §2.5 Mellin Transform Methods ‣ Areas ‣ Chapter 2 ...
    The Mellin transforms converge absolutely and define analytic functions in the half-planes shown in Table 2.5.1.<|control11|><|separator|>
  58. [58]
    [PDF] The Volume of n-balls - Rose-Hulman Scholar
    As we will see, the volume of n-balls is closely related to the gamma function. In this section we compute various quantities related to the gamma function ...
  59. [59]
    DLMF: §25.2 Definition and Expansions ‣ Riemann Zeta Function ‣ Chapter 25 Zeta and Related Functions
    ### Integral Representation of the Riemann Zeta Function Involving the Gamma Function
  60. [60]
    DLMF: §25.4 Reflection Formulas ‣ Riemann Zeta Function ...
    §25.4 Reflection Formulas ; Γ · ( z ) : gamma function, ; ζ · ( s ) : Riemann zeta function, π : the ratio of the circumference of a circle to its diameter, ; cos · z ...Missing: functional | Show results with:functional
  61. [61]
    On Barnes' Multiple Zeta and Gamma Functions - ScienceDirect.com
    We show how various known results concerning the Barnes multiple zeta and gamma functions can be obtained as specializations of simple features.
  62. [62]
    [PDF] On Barnes' Multiple Zeta and Gamma Functions
    We also demonstrate how Barnes' multiplt: zeta and gamma functions tit into a recently developed theory of minimal solutions to first order analytic dit1erence ...
  63. [63]
    DLMF: §10.39 Relations to Other Functions ‣ Modified Bessel ...
    Confluent Hypergeometric Functions ; Γ · : gamma function ; I ν · : modified Bessel function of the first kind ...
  64. [64]
    Factorial : Gamma :: Primorial :? - MathOverflow
    May 15, 2017 · The function n# has long flat sections, which messes with convexity. For example, if f(n)=n# then f(2)=2, f(3)=6, and f(4)=6.Studying primes via the gamma function alone: $(x+1)\prod_n ...Primes $p$ for which $p-1$ has a large prime factor - MathOverflowMore results from mathoverflow.netMissing: Primorials | Show results with:Primorials
  65. [65]
    1.3.6.6.11. Gamma Distribution - Information Technology Laboratory
    The following is the plot of the gamma hazard function with the same values of γ as the pdf plots above.
  66. [66]
    15.4 - Gamma Distributions | STAT 414
    In an approximate Poisson process with mean, the waiting time until the first event occurs follows an exponential distribution with mean.
  67. [67]
    [PDF] Theorem The chi-square distribution is a special case of the gamma ...
    Theorem The chi-square distribution is a special case of the gamma distribution when n = 2β and α = 2. f(x) = 1 αβΓ(β) xβ-1e-x/α x > 0. When n = 2β and α = 2, ...
  68. [68]
    [PDF] Chapter 9 The exponential family: Conjugate priors - People @EECS
    The corresponding conjugate prior retains the shape of the Poisson likelihood: p(θ |α) ∝ θα1−1e−α2θ,. (9.32) which we recognize as the gamma distribution:.
  69. [69]
    [PDF] Hydrogen Radial Wavefunctions - MIT OpenCourseWare
    Hydrogen Radial Wavefunctions. The Hydrogen atom is special because it has electronic states and properties that scale with the principal, n, and orbital ...Missing: Gamma | Show results with:Gamma
  70. [70]
    [PDF] on the definition of euler gamma function - IMJ-PRG
    For a brief historical accounts on the subject of the Gamma function and Eulerian. Integrals (as the subject was known in the XIXth century), we refer to the ...
  71. [71]
    [PDF] 1035-W1-54 - Joint Mathematics Meetings
    The Euler-Stirling Correspondence. The real analysis course is often ... For many years, I have used three letters between Leonhard Euler and James Stirling for.
  72. [72]
    Exercices de calcul intégral sur divers ordres de transcendantes et ...
    Mar 30, 2006 · Legendre, A. M. (Adrien Marie), 1752-1833. Publication date: 1811-17. Topics: Calculus, Integral, Elliptic functions, Gamma functions. Publisher ...
  73. [73]
  74. [74]
    [PDF] ·The Gamma Function
    integration by parts gives ... We have aquainted ourselves with three different functional equations for the gamma function: the functional equation (2.2), the ...
  75. [75]
    [PDF] adic $\Gamma$-function - Benedict H. Gross; Neal Koblitz
    Mar 7, 2004 · 1979 by Princeton University Mathematics Department. For copying information, see inside back cover. Page 3. 570. BENEDICT H. GROSS AND NEAL ...Missing: original | Show results with:original
  76. [76]
    DLMF: Chapter 8 Incomplete Gamma and Related Functions
    Definitions and properties of incomplete gamma functions from NIST Digital Library of Mathematical Functions.
  77. [77]
    Multivariate gamma function - Wikipedia
    Definition and applications of the multivariate gamma function.
  78. [78]
    DLMF: §5.17 Barnes' G-Function (Double Gamma Function)
    Properties of the Barnes G-function from NIST DLMF.
  79. [79]
    Barnes G-Function -- from Wolfram MathWorld
    Overview of the Barnes G-function.
  80. [80]
    Multiple Gamma Function and Its Application to Computation of Series
    arXiv paper on multiple gamma function by Barnes.
  81. [81]
    DLMF: §5.18 q-Gamma and q-Beta Functions
    Definition and properties of q-gamma function from NIST DLMF.
  82. [82]
    DLMF: §8.16 Generalizations
    Generalizations of incomplete gamma functions.
  83. [83]
    The generalized gamma functions
    arXiv paper on generalized gamma functions.