Fact-checked by Grok 2 weeks ago

Mellin transform

The Mellin transform is an in that maps a f(t) defined on the positive real line to a \phi(z) in the , defined by the formula \phi(z) = \int_0^\infty t^{z-1} f(t) \, dt for appropriate complex values z where the integral converges. Its inverse is given by the contour integral f(t) = \frac{1}{2\pi i} \int_{c-i\infty}^{c+i\infty} t^{-z} \phi(z) \, dz, where c is chosen in the strip of analyticity. This transform is particularly suited for analyzing with multiplicative or scaling properties, as it diagonalizes dilations unlike the shift-invariant . Named after the Finnish mathematician Hjalmar Mellin (1854–1933), who systematically developed it in the late , the transform builds on earlier work in integral representations and has since become a cornerstone in . It is closely related to the , obtained by the substitution t = e^{-u}, which converts the Mellin transform of f(t) into the of f(e^{-u}). Similarly, it connects to the through exponential changes of variables, making it a "multiplicative version" of these classical transforms. Key properties include , (M\{f(at); z\} = a^{-z} \phi(z)), and differentiation rules involving the , such as M\{f'(t); z\} = -\frac{\Gamma(z)}{\Gamma(z-1)} \phi(z-1). The transform's typically requires f(t) to decay appropriately at both t \to 0^+ and t \to \infty, often analyzed via the strip of holomorphy in the . Notable applications span multiple fields: in , it facilitates proofs of the by relating arithmetic functions like the to its Mellin transform. In statistics, it simplifies the study of distributions for products and quotients of independent positive random variables, yielding closed forms for densities like the Student's t- and F-distributions. For partial differential equations, such as in wedge domains, the Mellin transform reduces problems to ordinary differential equations via in logarithmic coordinates. In , its makes it essential for of algorithms, including divide-and-conquer recurrences and harmonic sums in probabilistic models.

Definition and Convergence

Definition

The Mellin transform of a function f(x) is defined as \mathcal{M}\{f\}(s) = \int_{0}^{\infty} f(x) \, x^{s-1} \, dx, where s \in \mathbb{C} is a complex variable. This generalizes certain aspects of the Laplace and transforms by incorporating a multiplicative x^{s-1}, making it particularly suited for problems involving scaling or homogeneity. The inverse Mellin transform recovers the original function via the contour integral f(x) = \frac{1}{2\pi i} \int_{c - i\infty}^{c + i\infty} \mathcal{M}\{f\}(s) \, x^{-s} \, ds, where the path of integration is a vertical line \operatorname{Re}(s) = c in the complex plane. This inversion formula relies on the residue theorem and analytic properties of the transform, ensuring uniqueness under appropriate conditions. Common notation for the Mellin transform includes \mathcal{M}\{f\}(s) or M_f(s), with the inverse often denoted similarly. In related contexts, such as the theory of , the transform connects directly to the \Gamma(s); for instance, the Mellin transform of e^{-x} yields \Gamma(s) for \operatorname{Re}(s) > 0, highlighting its role in representing factorial-like generalizations. The Mellin transform was introduced by the Finnish mathematician Robert Hjalmar Mellin in 1897, initially as a tool for analyzing integral equations and the theories of the gamma and hypergeometric functions.

Fundamental Strip and Analytic Continuation

The convergence of the Mellin transform M\{f\}(s) = \int_0^\infty f(x) x^{s-1} \, dx is analyzed in the complex s-plane, where s = \sigma + i \tau with \sigma = \operatorname{Re}(s) and \tau = \operatorname{Im}(s). The integral converges absolutely in a vertical strip defined by a < \sigma < b, known as the fundamental strip \langle a, b \rangle, which is the maximal such open strip. This strip is determined by the asymptotic behavior of f(x) at the endpoints of the integration interval. Specifically, near x \to 0^+, if f(x) = O(x^{-\alpha}) for some \alpha \in \mathbb{R}, the condition for convergence of the integral over (0,1] requires \sigma > \alpha, setting the left boundary a = \alpha. Similarly, as x \to \infty, if f(x) = O(x^{-\beta}) for \beta > 0, convergence over [1, \infty) holds for \sigma < \beta, establishing the right boundary b = \beta. Provided a < b, the Mellin transform exists and is analytic throughout this fundamental strip. The boundaries of the fundamental strip arise directly from the integrability conditions split across the near-origin and far-origin behaviors of f. For the portion near zero, the integral \int_0^1 |f(x)| x^{\sigma - 1} \, dx converges when the exponent ensures the power-law decay dominates, yielding the lower limit on \sigma. At infinity, \int_1^\infty |f(x)| x^{\sigma - 1} \, dx imposes an upper limit on \sigma to counteract the growth or decay of f(x). If f is piecewise continuous and of bounded variation in finite intervals, or satisfies milder local integrability conditions, the strip remains well-defined as long as the global asymptotics permit a < b. Outside this strip, the integral may diverge, but the transform can still be meaningfully extended. Within the fundamental strip, M\{f\}(s) is holomorphic, reflecting the analytic dependence of the integral on the parameter s. Beyond the strip, analytic continuation extends the domain of definition, typically resulting in a meromorphic function in larger regions of the complex plane. This continuation is achieved through methods such as functional equations relating values at different points or contour deformations via the residue theorem, allowing evaluation in strips where the original integral diverges. The meromorphic nature implies that singularities are isolated poles, with the transform analytic elsewhere except at these points. Singularities in the continued Mellin transform often manifest as poles whose locations and orders correspond to the asymptotic expansions of f(x) at $0^+ or \infty. For instance, a term x^c (\log x)^k in the expansion of f(x) near zero induces a pole of order k+1 at s = -c, with the residue capturing the coefficient of that term. Such poles frequently arise from factors involving the gamma function in explicit representations, where simple poles occur at non-positive integers due to \Gamma(s). These residues play a key role in inversion formulas and asymptotic approximations, enabling the recovery of f(x) via contour integrals that encircle the poles.

Relations to Other Transforms

Connection to Laplace and Fourier Transforms

The Mellin transform is closely related to the two-sided Laplace transform through a logarithmic change of variables. Specifically, substituting t = -\log x (so x = e^{-t} and dx = -e^{-t} dt) into the Mellin transform integral M\{f\}(s) = \int_0^\infty f(x) x^{s-1} \, dx yields M\{f\}(s) = \int_{-\infty}^\infty f(e^{-t}) e^{-s t} \, dt, which is the two-sided Laplace transform \mathcal{L}\{g\}(s) of the function g(t) = f(e^{-t}). This equivalence holds in the fundamental strip where both transforms converge, determined by the asymptotic behavior of f(x) as x \to 0^+ and x \to \infty. The relation extends to the Fourier transform by viewing the Mellin transform as the Fourier transform on the multiplicative group of positive real numbers under the Haar measure dx/x. With the substitution x = e^u (so u = \log x and dx = e^u du), the Mellin transform becomes M\{f\}(s) = \int_{-\infty}^\infty f(e^u) e^{s u} \, du, which corresponds to the Fourier transform of h(u) = f(e^u) evaluated at frequency -s/i (depending on the Fourier convention). This perspective highlights the Mellin transform's role in analyzing scale-invariant problems, analogous to the Fourier transform's treatment of translation-invariant ones. Extensions to two-sided transforms further link these: the bilateral Mellin transform for functions supported on (0, \infty) maps directly to the bilateral via the same logarithmic substitution, facilitating analysis on the full real line.

Mellin-Barnes Integral Representation

The Mellin-Barnes integral representation expresses certain special functions as contour integrals in the complex plane, where the integrand consists of a product or ratio of multiplied by a power of the variable. This form arises naturally as the inverse Mellin transform of the Mellin transform of a function, providing a powerful tool for analytic continuation and representation of functions beyond their power series domains. Specifically, for functions like the generalized hypergeometric series, the representation takes the form f(z) = \frac{1}{2\pi i} \int_{\gamma - i\infty}^{\gamma + i\infty} M(s) z^{-s} \, ds, where M(s) is a meromorphic function typically expressed as a ratio of products of Gamma functions, such as \prod \Gamma(a_j + A_j s) / \prod \Gamma(b_k + B_k s), with positive coefficients A_j, B_k ensuring convergence properties. The contour of integration is chosen as a vertical line \operatorname{Re}(s) = \gamma in the fundamental strip of analyticity, where \gamma is selected to lie between the leftmost and rightmost poles of M(s), ensuring the integral converges absolutely for | \arg z | < \pi/2. In cases where poles lie on the line, the contour may be indented with small semicircles to avoid them, forming a Hankel-type path that separates poles of the numerator Gammas (typically to the left) from those of the denominator (to the right). This setup allows the integral to enclose residues corresponding to series expansions when deformed appropriately. These representations are particularly valuable for asymptotic analysis, as the Mellin-Barnes form facilitates the application of the saddle-point method to evaluate the integral for large |z|. By deforming the contour to pass through saddle points of the phase function \log M(s) - s \log z, one obtains uniform asymptotic expansions that capture the leading behavior and higher-order terms, often superior to direct methods for functions with branch points or Stokes lines. The Mellin-Barnes integral is synonymous with the Barnes integral, introduced by E. W. Barnes in his work on multiple Gamma functions, and extends to the general Barnes integral with multiple complex parameters. In its most general form, involving products of multiple Gamma factors, it defines the Meijer G-function, a unifying framework for many transcendental functions, where the integral G_{p,q}^{m,n} \left( z \,\middle|\, \begin{matrix} a_1, \dots, a_p \\ b_1, \dots, b_q \end{matrix} \right) = \frac{1}{2\pi i} \int_L \frac{\prod_{j=1}^m \Gamma(b_j + s) \prod_{j=1}^n \Gamma(1 - a_j - s)}{\prod_{j=m+1}^q \Gamma(1 - b_j - s) \prod_{j=n+1}^p \Gamma(a_j + s)} z^{-s} \, ds encodes the structure with a suitable contour L separating pole clusters.

Examples of Mellin Transforms

Transforms of Basic Functions

The Mellin transforms of elementary functions often yield closed-form expressions involving the or other special functions, valid within specific vertical strips in the complex plane where the defining integral converges absolutely. These transforms are fundamental for understanding the analytic continuation and properties of the Mellin transform, as the location and nature of singularities in the transform reveal asymptotic behaviors of the original function near zero and infinity. For the power function f(x) = x^{\alpha}, the integral \int_0^\infty x^{s-1} x^{\alpha} \, dx diverges in the classical sense, but in the distributional framework, its Mellin transform is given by \delta(s + \alpha), where \delta denotes the Dirac delta distribution. This holds over the entire complex plane, reflecting the homogeneous nature of power functions under scaling. The exponential function f(x) = e^{-x} has the Mellin transform \Gamma(s), where \Gamma is the , converging for \operatorname{Re}(s) > 0. This is a direct consequence of the integral representation of the gamma function itself. For the generalized Gaussian f(x) = e^{-x^{\beta}} with \beta > 0, the Mellin transform is \frac{1}{\beta} \Gamma\left(\frac{s}{\beta}\right), valid for \operatorname{Re}(s) > 0. The case \beta = 2 simplifies to \frac{1}{2} \Gamma\left(\frac{s}{2}\right), obtained via the substitution u = x^{\beta}. Polynomials, being finite linear combinations of power functions, have Mellin transforms that are sums of Dirac delta distributions; for the monomial f(x) = x^{k} with nonnegative k, it is \delta(s + k) in the distributional sense. Alternatively, for regularized versions like moments of exponentials, such as M\{x^{k} e^{-x}\}(s) = \frac{\Gamma(s + k)}{\Gamma(s)} for \operatorname{Re}(s) > 0, these can be derived by repeated of the transform, since differentiation under the integral corresponds to multiplication by powers in the transform domain. A representative rational function, f(x) = \frac{1}{1 + x}, has Mellin transform \pi \csc(\pi s), converging in the strip $0 < \operatorname{Re}(s) < 1. This result follows from the beta function representation after substitution t = x/(1+x).
Function f(x)Mellin Transform M\{f\}(s)Convergence Strip
x^{\alpha}\delta(s + \alpha)Entire (distributional)
e^{-x}\Gamma(s)\operatorname{Re}(s) > 0
e^{-x^{\beta}} (\beta > 0)\frac{1}{\beta} \Gamma\left(\frac{s}{\beta}\right)\operatorname{Re}(s) > 0
x^{k} (k \in \mathbb{N}_0)\delta(s + k)Entire (distributional)
\frac{1}{1 + x}\pi \csc(\pi s)$0 < \operatorname{Re}(s) < 1

Transforms Involving Special Functions

The Mellin transform provides integral representations for several special functions, particularly through connections to , Dirichlet series, and generating functions for hypergeometric series. One prominent example is the relation to the Riemann zeta function via the Jacobi , defined as \theta(x) = \sum_{n=-\infty}^{\infty} e^{-\pi n^2 x} for x > 0. The Mellin transform of \theta(x) - 1 yields \int_0^\infty (\theta(x) - 1) x^{s/2 - 1} \, dx = 2 \pi^{-s/2} \Gamma(s/2) \zeta(s), valid for \Re(s) > 1, where the subtraction of 1 ensures at the lower limit. This formula links the oscillatory behavior of the to the analytic properties of the zeta function, facilitating its meromorphic continuation via the derived from the applied to \theta(x). The of \Re(s) > 1 corresponds to the of the zeta function's representation. Dirichlet series arise naturally as Mellin transforms of power series generating functions. Consider f(x) = \sum_{n=0}^\infty a_n x^n for $0 < x < 1, where the coefficients a_n grow sufficiently slowly to ensure convergence. The Mellin transform is M\{f\}(s) = \int_0^1 f(x) x^{s-1} \, dx = \sum_{n=0}^\infty a_n \int_0^1 x^{n + s - 1} \, dx = \sum_{n=0}^\infty \frac{a_n}{n + s}, converging in half-planes determined by the asymptotic growth of a_n, typically \Re(s) > -\sigma_0 for some \sigma_0 \geq 0 depending on the series. This representation interchanges summation and integration under suitable conditions, such as uniform convergence on compact subsets of the unit disk, and extends to more general \sum a_n n^{-s} via logarithmic substitution x = e^{-t}. Power series generating functions , such as those appearing , have Mellin transforms expressible . , the binomial expansion (1 - x)^{-a} = \sum_{n=0}^\infty \binom{n + a - 1}{n} x^n for $0 < x < 1 and \Re(a) > 0 has Mellin transform \int_0^1 (1 - x)^{-a} x^{s - 1} \, dx = B(s, a) = \frac{\Gamma(s) \Gamma(a)}{\Gamma(s + a)}, valid for \Re(s) > 0 and \Re(a) > 0. This beta function integral serves as a building block for hypergeometric functions, as the hypergeometric {}_2F_1(a, b; c; x) admits a Mellin-Barnes contour integral representation involving products of gamma functions, linking power series expansions to meromorphic continuations in the complex plane. A specific application of the Mellin transform to alternating series is the Cahen-Mellin integral, which provides representations for sums like \sum_{n=0}^\infty \frac{(-1)^n}{n + s}. This sum equals \int_0^1 \frac{x^{s-1}}{1 + x} \, dx for \Re(s) > 0, obtained by term-by-term integration of the geometric series $1/(1 + x) = \sum_{n=0}^\infty (-1)^n x^n. Extending to the full positive real line gives the key identity \int_0^\infty \frac{x^{s-1}}{1 + x} \, dx = \frac{\pi}{\sin(\pi s)}, for $0 < \Re(s) < 1, derived via contour integration or residue theorem. The original alternating sum relates to this via the decomposition of the integral over (0,1) and (1,\infty), with convergence in half-planes \Re(s) > 0 adjusted for the pole structure. For zeta-related contexts, such representations hold in strips like \Re(s) > 1.

Properties

Linearity and Differentiation Under the Integral

The Mellin transform is a linear operator. For complex constants \alpha and \beta, and functions f(x) and g(x) for which the respective transforms exist in a common vertical strip of the complex plane, the following holds: \mathcal{M}\{\alpha f + \beta g\}(s) = \alpha \mathcal{M}\{f\}(s) + \beta \mathcal{M}\{g\}(s). This property arises directly from the linearity of the defining integral \mathcal{M}\{f\}(s) = \int_0^\infty f(x) x^{s-1} \, dx. Under suitable regularity conditions on f(x) that permit differentiation under the integral sign—such as absolute integrability of f(x) x^{s-1} \log x in the fundamental strip—the derivative of the Mellin transform with respect to the complex parameter s satisfies \frac{d}{ds} \mathcal{M}\{f\}(s) = \int_0^\infty f(x) x^{s-1} \log x \, dx = \mathcal{M}\{f(x) \log x\}(s). This relation links the Mellin transform to functions modified by a logarithmic factor and is often applied in and the evaluation of integrals involving parameters. The Mellin transform also accommodates differentiation of the original function via . Assuming boundary conditions where the relevant boundary terms vanish (e.g., f(x) \to 0 sufficiently rapidly as x \to 0^+ and x \to \infty), the transform of the is \mathcal{M}\left\{\frac{d}{dx} f(x)\right\}(s) = -(s-1) \mathcal{M}\{f\}(s-1), provided the transforms exist in the appropriate shifted strips. This property, derived from partial integration on the defining integral, facilitates the solution of equations in Mellin space. A key feature related to is the transform's response to argument : for a > 0, \mathcal{M}\{f(ax)\}(s) = a^{-s} \mathcal{M}\{f\}(s). This follows from the u = ax in the , highlighting the transform's utility in problems with multiplicative structure and .

Multiplication and Convolution Theorems

The Mellin transform exhibits a convolution theorem analogous to those in Fourier and Laplace analysis, but adapted to the multiplicative structure of the positive real line. Specifically, the Mellin convolution of two functions f and g is defined as (f \ast g)(x) = \int_0^\infty f(t) \, g\left(\frac{x}{t}\right) \frac{dt}{t}, provided the integral converges appropriately in the fundamental strips of f and g. The Mellin transform of this convolution is the pointwise product of the individual transforms: \mathcal{M}\{f \ast g\}(s) = \mathcal{M}\{f\}(s) \, \mathcal{M}\{g\}(s), where the convergence holds for \operatorname{Re}(s) in the intersection of the respective fundamental strips. This property follows from the integral representation and Fubini's theorem under suitable analyticity conditions. The dual property, known as the multiplication theorem, addresses the transform of a product of functions. For functions f and g, the Mellin transform of their product f(x) g(x) (adjusted for the kernel's homogeneity) yields a convolution in the transform domain: \mathcal{M}\{f \cdot g\}(s) = \frac{1}{2\pi i} \int_{c - i\infty}^{c + i\infty} \mathcal{M}\{f\}(s - w) \, \mathcal{M}\{g\}(w) \, dw, where the vertical contour \operatorname{Re}(w) = c is chosen within the intersection of the strips where both transforms are analytic, ensuring absolute convergence. This integral arises from substituting the inverse Mellin formula for one function into the direct transform of the product and interchanging orders via the residue theorem or analytic continuation. These theorems reflect the Mellin transform's in diagonalizing operations on the (\mathbb{R}_+, \times). The transform effectively Fourier-analyzes dilations on this group, converting the Mellin —invariant under group multiplication—into multiplication in the s-, thereby simplifying problems involving scale-invariant structures. In practice, the facilitates solving Mellin-type integral equations of the form (f \ast k)(x) = h(x), where k is a known . Taking Mellin transforms yields \mathcal{M}\{f\}(s) \, \mathcal{M}\{k\}(s) = \mathcal{M}\{h\}(s), so \mathcal{M}\{f\}(s) = \mathcal{M}\{h\}(s) / \mathcal{M}\{k\}(s) (assuming \mathcal{M}\{k\}(s) \neq 0 and where needed), followed by inversion to recover f. This approach is particularly effective for equations arising in scaling problems, such as those in asymptotic analysis or singular integrals.

Advanced Theorems

Parseval's and Plancherel's Theorems

Parseval's theorem for the Mellin transform relates the integral of the product of two functions over the positive real line to an integral involving their Mellin transforms along a vertical in the . For suitable functions f and g analytic in a common strip of , the theorem asserts that \int_0^\infty f(x) g(x) \, dx = \frac{1}{2\pi i} \int_{c-i\infty}^{c+i\infty} \hat{f}(s) \hat{g}(1-s) \, ds, where \hat{f}(s) and \hat{g}(s) denote the Mellin transforms of f and g, respectively (assuming real-valued functions), and c lies in the intersection of the fundamental strips. This relation holds under conditions ensuring absolute integrability or square-integrability in the respective domains, often requiring f, g \in L^2(\mathbb{R}^+, dx) with appropriate decay at the boundaries. Plancherel's theorem extends this to the preservation of norms for square-integrable functions, establishing that the Mellin transform acts as an up to a constant factor on suitable L^2 spaces. Specifically, for f \in L^2(\mathbb{R}^+, dx) with Mellin transform analytic in a vertical strip containing \sigma, \|f\|_2^2 = \int_0^\infty |f(x)|^2 \, dx = \frac{1}{2\pi} \int_{-\infty}^\infty |\hat{f}(\sigma + i\tau)|^2 \, d\tau, where the line \operatorname{Re}(s) = \sigma is fixed within the strip of convergence. This formula confirms the unitarity of the Mellin transform (up to the $1/(2\pi) factor) when mapping between L^2(\mathbb{R}^+, dx) and the corresponding L^2 space along the imaginary axis shifted by \sigma, via the logarithmic substitution x = e^u with normalization h(u) = f(e^u) e^{u/2}. For the weighted space L^2(\mathbb{R}^+, dx/x), see the unitary structure below. The inner product preservation implied by these theorems underscores the Mellin transform's role as a (modulo scaling) on appropriate Hilbert spaces, facilitating energy conservation in applications like and . A sketch of the proof relies on the logarithmic x = e^u, which maps the Mellin transform to a of the rescaled function h(u) = f(e^u) e^{u/2}; applying the standard for the on L^2(\mathbb{R}, du) then yields the result after substitution back to the original variables.

Unitary Structure on L2 Spaces

The Hilbert space L^2((0, \infty), dx/x) consists of measurable functions f: (0, \infty) \to \mathbb{C} such that \int_0^\infty |f(x)|^2 \, dx/x < \infty, equipped with the inner product \langle f, g \rangle = \int_0^\infty f(x) \overline{g(x)} \, dx/x. This space is invariant under multiplication by positive constants, reflecting the structure of the multiplicative group (0, \infty) under which the measure dx/x is the unique (up to scalar) Haar measure, ensuring left-invariance: for a > 0, \int_0^\infty f(ax) \, dx/x = \int_0^\infty f(x) \, dx/x. The weight dx/x thus endows the space with a group-theoretic foundation, analogous to dx for the additive group \mathbb{R}. The Mellin transform \mathcal{M} f(s) = \int_0^\infty f(x) x^{s-1} \, dx (often normalized with the measure incorporated as \int_0^\infty f(x) x^{s-1} \, dx/x) acts as an from L^2((0, \infty), dx/x) to L^2(\mathbb{R}, d\tau/(2\pi)), where the is identified with functions on the vertical line \operatorname{Re}(s) = 1/2 via s = 1/2 + i\tau with \tau \in \mathbb{R}. Specifically, for f \in L^2((0, \infty), dx/x), \int_0^\infty |f(x)|^2 \, \frac{dx}{x} = \frac{1}{2\pi} \int_{-\infty}^\infty |\mathcal{M} f(1/2 + i\tau)|^2 \, d\tau, establishing the L^2-preservation. This mapping is unitary, meaning it is a bijective with a bounded given by the inverse Mellin transform f(x) = \frac{1}{2\pi i} \int_{1/2 - i\infty}^{1/2 + i\infty} \mathcal{M} f(s) x^{-s} \, ds, which recovers f . The unitarity follows from the substitution x = e^u, which conjugates the Mellin transform to the on L^2(\mathbb{R}, du), a known . The role of the measure dx/x is pivotal in preserving the multiplicative group structure, where the characters x^{it} (for t \in \mathbb{R}) diagonalize the dilation operators D_a f(x) = a^{-1/2} f(ax) (unitarily normalized), leading to the via the . This framework highlights the as the tool for scale-invariant problems on the positive reals. Extensions of this unitary structure include mappings to spaces H^2 in vertical strips of the , where the of functions in weighted L^2 subspaces on (0, \infty) yields analytic functions bounded in the half-plane \operatorname{Re}(s) > 1/2, with boundary values in L^2 on the critical line. Additionally, the explicitly links the structure to the standard via the logarithmic change of variables, facilitating extensions to broader function classes like tempered distributions on the .

Applications

In Number Theory and Zeta Functions

The Mellin transform plays a pivotal role in deriving the of the through the inversion formula for the . The Jacobi \theta(t) = \sum_{n \in \mathbb{Z}} e^{-\pi n^2 t} for \operatorname{Re}(t) > 0 satisfies the transformation law \theta(1/t) = \sqrt{t} \, \theta(t), obtained via Poisson summation. Applying the Mellin transform to \theta(t), specifically \phi(s) = \int_1^\infty (\theta(t) - 1) t^{s/2 - 1} \, dt + \int_0^1 (\theta(t) - t^{-1/2}) t^{s/2 - 1} \, dt, yields an that satisfies \phi(s) = \phi(1 - s). This , combined with the relation \pi^{-s/2} \Gamma(s/2) \zeta(s) = \frac{1}{s-1} + \frac{1}{1-s} + \phi(s)/2, provides the of \zeta(s) to the (except for a simple pole at s=1) and establishes the \Lambda(s) = \pi^{-s/2} \Gamma(s/2) \zeta(s) = \Lambda(1 - s). In analytic number theory, the Mellin transform connects the pole of \zeta(s) at s=1 to asymptotic estimates in the prime number theorem. The logarithmic derivative \zeta'(s)/\zeta(s) admits a Mellin transform representation as \zeta'(s)/\zeta(s) = -\sum_p \log p \sum_{k=1}^\infty p^{-k s} / (1 - p^{-k s}), which relates to the Chebyshev function \psi(x) = \sum_{p^k \leq x} \log p. By expressing \log \zeta(s) as the Mellin transform s \int_1^\infty J(x) x^{-s-1} \, dx where J(x) = \sum_{k=1}^\infty \pi(x^{1/k})/k, the residue at the pole s=1 (of order 1 with residue 1) implies \psi(x) \sim x via Tauberian theorems like Ikehara's. This yields the prime number theorem \pi(x) \sim x / \log x, with error terms refined by zero-free regions near \operatorname{Re}(s)=1. The approach extends to Dirichlet L-functions L(s, \chi) = \sum_{n=1}^\infty \chi(n) n^{-s} for a primitive Dirichlet character \chi modulo q, using Mellin transforms of associated theta functions. Define \theta_\chi(x) = \sum_{n=1}^\infty \chi(n) e^{-\pi n^2 x} (adjusted for parity), whose Mellin transform gives \pi^{-(s+a)/2} \Gamma((s+a)/2) L(s, \chi) = \int_0^\infty \theta_\chi(x) x^{(s+a)/2 - 1} \, dx, where a = 0 or 1 depending on \chi. Splitting the integral and applying Poisson summation to the part from 0 to 1, leveraging the character's Gauss sum, produces the functional equation q^{s/2} (2\pi)^{-s} \Gamma(s) L(s, \chi) = \epsilon(\chi) q^{(1-s)/2} (2\pi)^{s-1} \Gamma(1-s) L(1-s, \overline{\chi}) (up to parity adjustments), enabling meromorphic continuation and revealing non-vanishing properties at s=1 for non-principal \chi. This generalizes the zeta case and underpins class number formulas and Artin reciprocity. The Selberg trace formula further illustrates the Mellin transform's utility in linking spectral theory to number-theoretic sums over arithmetic groups like \mathrm{PSL}_2(\mathbb{Z}). On the hyperbolic plane, the trace formula equates the spectral side \sum_j h(1/2 + i r_j) + \frac{1}{4\pi} \int_{-\infty}^\infty h(1/2 + i t) \frac{t \tanh(\pi t)}{2} \, dt (summing over Laplacian eigenvalues \lambda_j = 1/4 + r_j^2) to the geometric side involving lengths of closed geodesics. Here, h(s) is the Mellin transform of a test function g(y) = \int K(x,y) \, dx on the upper half-plane, facilitating the inversion and Plancherel measure \frac{t \tanh(\pi t)}{2\pi} dt for the continuous spectrum. This framework connects eigenvalue distributions to prime-like sums in the fundamental domain, with applications to subconvexity bounds for L-functions via spectral gaps.

In Probability and Moment-Generating Functions

In , the Mellin transform serves as a moment-generating tool for positive s, capturing their power moments through evaluation at integer points. For a non-negative X with f(x) supported on (0, \infty), the Mellin transform is given by M(s) = \int_0^\infty x^{s-1} f(x) \, dx = \mathbb{E}[X^{s-1}], where the integral converges for s in a suitable strip of the depending on the distribution's tail behavior. Specifically, for positive integers k, M(k+1) = \mathbb{E}[X^k] yields the raw moments, providing a multiplicative analog to the classical M_X(t) = \mathbb{E}[e^{tX}] but adapted to the positive reals and power laws rather than exponentials. This property facilitates the study of distributions with heavy tails, where ordinary moments may diverge, by analytically continuing M(s) beyond integers. The Mellin transform also connects deeply with log-stable distributions, where \log X follows a stable law. In such cases, the transform links directly to the characteristic function of \log X via the substitution u = \log x, recasting the Mellin as a over the logarithmic scale. Precisely, M(1 + it) = \mathbb{E}[X^{it}] = \phi_{\log X}(t), where \phi_{\log X}(t) denotes the of \log X. This relationship is particularly valuable for deriving densities and stability parameters in multiplicative processes, such as those in or physics, where log-stable laws model phenomena like or income distributions. Regarding cumulants, the logarithm of the Mellin transform generates the cumulants of \log X. The function \log M(1 + z) admits a whose coefficients are the cumulants \kappa_n of \log X, via \log M(1 + z) = \sum_{n=1}^\infty \kappa_n \frac{z^n}{n!}, offering insights into , additivity, and asymptotic behavior under . This logarithmic cumulant approach, rooted in the Mellin framework, enhances robustness against outliers compared to direct methods, as it emphasizes relative scales and is widely applied in robust for skewed or heavy-tailed . Representative examples illustrate these concepts for common distributions. The , with shape \alpha > 0 and scale 1 (density f(x) = \alpha x^{-\alpha-1} for x \geq 1), has Mellin transform M(s) = \frac{\alpha}{\alpha - s + 1}, \quad \Re(s) < \alpha + 1, directly implying moments \mathbb{E}[X^k] = \alpha / (\alpha - k) for k < \alpha, which highlights the distribution's power-law tails and finite moments only below the shape parameter. Similarly, the lognormal distribution, where \log X \sim \mathcal{N}(\mu, \sigma^2), yields M(s) = \exp\left( (s-1)\mu + \frac{(s-1)^2 \sigma^2}{2} \right), mirroring the moment-generating function of the underlying normal and enabling exact computation of all moments while underscoring the distribution's log-convexity and applications in modeling multiplicative noise.

In Solving Differential Equations

The Mellin transform is particularly effective for solving partial differential equations (PDEs) and ordinary differential equations (ODEs) that exhibit scale invariance or multiplicative structure, such as those arising in radial or cylindrical coordinates. By transforming the radial variable r into the complex parameter s, the transform converts differentiation with respect to r into algebraic operations involving s, simplifying equations like the Euler-Cauchy type that commonly appear in separated variables solutions. In cylindrical coordinates, the Mellin transform facilitates the solution of the radial Laplace equation, \nabla^2 u = 0, especially for problems with azimuthal dependence handled via . Assuming separation of variables u(r, \theta) = R(r) \Theta(\theta), the angular part yields \Theta'' + m^2 \Theta = 0 for integer m, leading to the radial r^2 R'' + r R' - m^2 R = 0. Applying the Mellin transform \mathcal{M}\{R(r)\}(s) = \int_0^\infty r^{s-1} R(r) \, dr reduces this to the algebraic relation s^2 \hat{R}(s) - m^2 \hat{R}(s) = 0, with solutions \hat{R}(s) \propto s^2 - m^2 = 0, corresponding to power-law behaviors R(r) \propto r^{\pm m} upon inversion. For full cylindrical problems without axial dependence (effectively 2D polar), the transform in r combined with Fourier in \theta yields through the inverse, as the Mellin contour integral evaluates to modified for certain boundary geometries. For Sturm-Liouville problems, particularly those with , the maps the differential operator L = \sum_{k=0}^n a_k (r \frac{d}{dr})^k to a polynomial in s: \hat{L}(s) = \sum_{k=0}^n a_k (-s)^k \hat{u}(s) = \hat{g}(s). This algebraic form allows solving for \hat{u}(s) = \hat{g}(s) / \hat{L}(s), with the inverse providing the solution in the spatial domain, provided the poles of $1/\hat{L}(s) are analyzed within the fundamental strip of analyticity. Such reductions are standard for self-adjoint problems on (0, \infty) with weight r^{\alpha}, where boundary conditions at r=0 and r=\infty determine the contour shift. A representative example is the radial heat equation in cylindrical coordinates, u_t = k (u_{rr} + \frac{1}{r} u_r), assuming azimuthal symmetry for simplicity. Applying the Mellin transform in r, \hat{u}(s, t) = \int_0^\infty r^{s-1} u(r, t) \, dr, transforms the spatial operator to k s^2 \hat{u}(s, t), yielding the ODE \partial_t \hat{u} = -k s^2 \hat{u}, whose solution is \hat{u}(s, t) = \hat{u}(s, 0) e^{-k s^2 t}. Inverting via the Mellin contour gives u(r, t), effectively solving a diffusion equation in the logarithmic variable \log r, where the transform acts as a Fourier representation on the multiplicative group. For wedge domains, this approach extends to Robin boundary conditions by parameterizing angular modes and using weighted Sobolev spaces defined via Mellin symbols for regularity analysis. In boundary value problems, the Mellin transform enables matching conditions across domains by exploiting the strip of convergence in the complex s-plane, where initial or boundary data are incorporated via residue calculus or contour deformation. For instance, in a wedge with Dirichlet data on the rays, the transform of the boundary function determines the coefficients, and the inverse along a suitable Bromwich contour resolves discontinuities at interfaces like r = a, producing series solutions convergent in subdomains. This method avoids singularities at the origin inherent in radial coordinates and aligns with convolution theorems for composite problems.

Selected Mellin Transforms

Table of Common Transforms

The Mellin transform, defined as M\{f\}(s) = \int_0^\infty x^{s-1} f(x) \, dx, is typically unilateral for functions supported on (0, \infty). The table below summarizes common examples, with convergence strips indicated. These are drawn from standard integral representations in mathematical analysis.
f(x)M\{f\}(s)Convergence Strip
e^{-x}\Gamma(s)\Re(s) > 0
x^{a-1} (1 + x)^{-b}\frac{\Gamma(s + a - 1) \Gamma(b - s - a + 1)}{\Gamma(b)}$1 - a < \Re(s) < b - a + 1, \Re(a) > 0, \Re(b - a) > 0
e^{-x^2}\frac{1}{2} \Gamma\left(\frac{s}{2}\right)\Re(s) > 0
\frac{1}{1 + x}\frac{\pi}{\sin(\pi s)}$0 < \Re(s) < 1
x^{a-1} e^{-x}\Gamma(a + s)\Re(a + s) > 0
The bilateral Mellin transform extends the to (-\infty, \infty) via \int_{-\infty}^\infty |x|^{s-1} f(x) \, dx, requiring adjustments for and , often used for functions defined on the full real line. For distributions like the Dirac delta, formal expressions such as the transform of \delta(x - 1) yielding $1 hold in the distributional sense, while representations involving x^{-s} / \Gamma(1 - s) arise in contexts or limiting cases.

Inverse Mellin Transform Examples

The inverse Mellin transform is often computed using the residue theorem by evaluating the Bromwich contour integral along a vertical line in the complex plane and closing the contour in the appropriate half-plane depending on the value of x. For functions analytic in a vertical strip, the poles of the Mellin transform M(s) determine the residues that sum to the original function f(x). A classic example involves the product of gamma functions M(s) = \Gamma(s) \Gamma(1-s), whose inverse Mellin transform is f(x) = \frac{1}{1+x} for x > 0. This follows from the reflection formula \Gamma(s) \Gamma(1-s) = \frac{\pi}{\sin(\pi s)}, which is derived using the residue theorem on a keyhole contour for the beta function integral representation. To compute the inverse explicitly, consider the contour integral \frac{1}{2\pi i} \int_{c-i\infty}^{c+i\infty} \Gamma(s) \Gamma(1-s) x^{-s} ds with $0 < c < 1. For x > 1, close the contour to the right, enclosing poles of \Gamma(1-s) at s = 1, 2, 3, \dots; the residues at these poles yield a series expansion \sum_{n=0}^\infty (-1)^n x^{-n-1}, which sums to \frac{1}{1+x}. For $0 < x < 1, close to the left, enclosing poles of \Gamma(s) at s = 0, -1, -2, \dots, leading to the equivalent series \sum_{n=0}^\infty (-1)^n x^{n}, again summing to \frac{1}{1+x}. This demonstrates how the residue theorem resolves the inversion across the unit interval. The choice of closing the Bromwich contour to the left or right is determined by the growth of x^{-s} as \operatorname{Im}(s) \to \pm \infty, with the exponential decay of the gamma functions ensuring the arc contribution vanishes. For x > 1, |x^{-s}| = x^{-\operatorname{Re}(s)} decays in the right half-plane (\operatorname{Re}(s) \to +\infty), so the arc contribution vanishes when closing rightward; residues at right poles give f(x). For $0 < x < 1, decay occurs in the left half-plane, closing leftward to capture left poles. This procedure ensures convergence within the fundamental strip $0 < \operatorname{Re}(s) < 1 for this example and generalizes to other meromorphic M(s). In number theory, the inverse Mellin transform of M(s) = \Gamma(s) \zeta(s) is f(x) = \sum_{n=1}^\infty e^{-n x}, the generating function for the geometric series of exponentials. This follows from the Mellin transform of each e^{-n x} being n^{-s} \Gamma(s), summing to \Gamma(s) \zeta(s). Perron's formula, a related inversion, approximates partial sums of Dirichlet series like \sum_{n \leq x} 1 \approx x using \frac{1}{2\pi i} \int \zeta(s) \frac{x^{s}}{s} ds, with residues contributing main terms and error estimates from contour shifts. Hypergeometric functions often arise as inverse Mellin transforms of products of gamma functions through the Mellin-Barnes representation. For instance, the Gauss {}_2F_1(a, b; c; x) for |x| < 1 is given by {}_2F_1(a, b; c; x) = \frac{\Gamma(c)}{ \Gamma(a) \Gamma(b) } \frac{1}{2\pi i} \int_{-i\infty}^{i\infty} \frac{ \Gamma(a + s) \Gamma(b + s) \Gamma(-s) }{ \Gamma(c + s) } (-x)^s ds, where the contour separates poles of \Gamma(-s) from those of \Gamma(a + s) and \Gamma(b + s). Closing the contour and summing residues at the poles of \Gamma(-s) at s = 0, -1, -2, \dots recovers the hypergeometric series \sum_{n=0}^\infty \frac{(a)_n (b)_n}{(c)_n} \frac{x^n}{n!}. This representation facilitates and of {}_2F_1. Similar forms hold for generalized hypergeometric functions, with products of more gamma factors corresponding to Meijer G-functions expressible in terms of {}_2F_1.

References

  1. [1]
    Mellin Transform -- from Wolfram MathWorld
    The Mellin transform is the integral transform defined by phi(z) = int_0^inftyt^(z-1)f(t)dt (1) f(t) = 1/(2pii)int_(c-iinfty)^(c+iinfty)t^(-z)phi(z)dz.
  2. [2]
    (PDF) The Mellin Transform - A Basic Introduction - ResearchGate
    Nov 23, 2019 · The Mellin transform is an integral transform named after the finnish mathematician Hjalmar Mellin (1854-1933).
  3. [3]
    [PDF] Mellin Transforms and Asymptotics - Inria
    It relies on the Mellin transform, a close relative of the integral transforms of Laplace and Fourier.
  4. [4]
    Some Applications of the Mellin Transform in Statistics - Project Euclid
    The Mellin transform is a natural analytical tool to use in studying the distribution of products and quotients of independent random variables.Missing: definition | Show results with:definition
  5. [5]
    [PDF] The Mellin Transform - HAL
    Feb 25, 2021 · Maybe the most famous application is the computation of the solution of a potential problem in a wedge-shaped region where the unknown function.
  6. [6]
    [PDF] The Mellin Transform
    The largest open strip hα, βi in which the integral converges is called the funda- mental strip, where hα, βi is the open strip of complex numbers s = σ+it such ...
  7. [7]
    [PDF] Mellin Transforms and Asymptotics : Harmonic Sums - Inria
    As the integral defining f*(s) depends analytically on the complex parameter s, a Mellin transform is in addition analytic in its fundamental strip. For ...
  8. [8]
    [PDF] Asymptotic Analysis via Mellin Transforms for Small Deviations in L2 ...
    We use Mellin transforms to compute a full asymptotic expansion for the tail of the. Laplace transform of the squared L2-norm of any multiply-integrated ...<|control11|><|separator|>
  9. [9]
    The Fourier, Hilbert and Mellin transforms on a half-line - arXiv
    Feb 7, 2023 · Fourier comes into play in the sense that the Mellin transform is simpy the Fourier transform on the locally compact Abelian multiplicative ...
  10. [10]
    Mellin-Barnes Integral -- from Wolfram MathWorld
    A type of integral containing gamma functions in its integrand. A typical such integral is given by f(z)=1/(2pii)int_(gamma-iinfty)^(gamma+iinfty)(Gamma(a_1+A_ ...
  11. [11]
    DLMF: §16.17 Definition ‣ Meijer 𝐺-Function ‣ Chapter 16 ...
    1 ≤ k ≤ n and 1 ≤ j ≤ m . Then the Meijer G -function is defined via the Mellin–Barnes integral representation: 16.17.1, G p , q m , n ⁡ ( z ; 𝐚 ; 𝐛 ) = G ...
  12. [12]
    [1712.05601] Asymptotics of the contour of the stationary phase and ...
    Dec 15, 2017 · A new approximation is proposed for the contour of the stationary phase of the Mellin-Barnes integrals in the case of its finite asymptotic ...
  13. [13]
    DLMF: §2.5 Mellin Transform Methods ‣ Areas ‣ Chapter 2 ...
    The inversion formula is given by. 2.5.2, f ⁡ ( t ) = 1 2 ⁢ π ⁢ i ⁢ ∫ c − i ... 1: Domains of convergence for Mellin transforms. Transform, Domain of ...
  14. [14]
    [PDF] 10. Functional Equation of the Zeta Function
    Apr 14, 2003 · θ(x) = O. 1. √x as x ց 0. 10.3. Proposition. For all s ∈ C with Re(s) > 1 one has. Γ s2 ζ(s) = πs ...
  15. [15]
    [PDF] appendix. the mellin transform
    The Mellin transformation is a basic tool for analyzing the behavior of many important functions in mathematics and mathematical physics, such as the zeta ...
  16. [16]
    [PDF] Tables of Integral Transforms
    Some of these integrals cannot be written as transforms, others were not included in the transform tables and are given here. Generally speaking, an integral ...
  17. [17]
    [PDF] Mellin transforms and asymptotics: Harmonic sums
    There, the integration line 9i (s) = c should be taken in the fundamental strip of the. Mellin transform. ... [ f (x) (log x); s] = d—ds f * (s). For ...
  18. [18]
    [PDF] Solution of a Certain Nonlocal ODE by Reduction to a Riemann ...
    (Mellin Convolution Theorem). Let M denote the Mellin transform. Given two functions ϕ, ψ for which the Mellin transforms Mϕ, Mψ are well-defined, we have.<|control11|><|separator|>
  19. [19]
    None
    Summary of each segment:
  20. [20]
    [PDF] New convolutions associated with the Mellin transform and their ...
    One of the most important properties of a convolution is to satisfy a fac- torization property which is typically associated with one or more than one integral.
  21. [21]
    [PDF] arXiv:1401.3143v2 [math.CA] 10 Mar 2014
    Mar 10, 2014 · The generalized Parseval equality for the Mellin transform is employed to prove the in- version theorem in L2 with the respective inverse ...
  22. [22]
    [PDF] General structure of Mellin transforms
    In number theory, the Mellin transform is used to study the behavior of certain arithmetic functions, such as the Riemann zeta func- tion. In probability theory ...
  23. [23]
    Mellin and Widder–Lambert Transforms with Applications in ... - MDPI
    This paper presents a comprehensive study of Plancherel's theorem and inversion formulae for the Widder–Lambert transform, extending its scope to Lebesgue ...
  24. [24]
    [PDF] The Mellin Transform in Signal Analysis. - DTIC
    Mar 11, 1994 · On L2(R) the K-transform is given by. A fOO. F+{t) = (M+f)(t) = f(x)K(x,t)dx, (3.3)(.) J—oo and. /•oo. F-(t) = (M.f)(t) = f(x)K{-x,t)dx; (3.3)(« ...
  25. [25]
    [PDF] Fourier Analysis Notes, Spring 2020 - Columbia Math Department
    Sep 3, 2020 · f(x)xs dx x. Note that the Mellin transform is the analog of the Fourier transform one gets when one replaces the additive group R with the ...
  26. [26]
    [PDF] Index Hypergeometric Integral Transform - Erwin Schrödinger Institute
    Aug 10, 2012 · The Mellin transform is a unitary operator from L2(R, dx/x) to L2 on ver- tical line Re s = 1/2. In particular,. Z ∞. 0 f(x)g(x) dx/x = 1. 2π Z.
  27. [27]
    [PDF] Unitary equivalence - DiVA portal
    Laplace transform on L2(0, ∞) are unitary equivalent to multiplication by the gamma functions Γ(1. 2 ± iw) on L2(R). We then give a new proof of the known.
  28. [28]
    [PDF] The Classical Theta Function and the Riemann Zeta Function
    Apr 3, 2019 · Crucially, using the Mellin transform we are able to use various properties of the theta function itself, in particular the modularity of the ...
  29. [29]
    [PDF] Lecture 2 : Functional equation of the Riemann ζ-function
    Note that the rapid decay of g implies that the integral defining the Mellin transform always converges at ∞. Example 5. Γ(s) = M(e−t)(s). In the proof of ...
  30. [30]
    [PDF] The Zeta Function and the Prime Number Theorem | Michael Taylor
    Thus there is a hope (vaguely realized) of obtaining J(x) via inversion of the Mellin transform. We make some comments on the close relationship of π(x) and J(x) ...
  31. [31]
    [PDF] Analytic continuation of Dirichlet L-functions & the Mellin transform
    e−πn2x =1+2θ(x). One derives 1.1 from the preliminary integral representation ζ(s) = πs/2. Γ(s/2).
  32. [32]
    [PDF] Spectral Theory and the Trace Formula (Expanded Text)
    We give an account of a portion of the spectral theory ?nSL2(R), particularly the. Selberg trace formula, emphasizing ideas from representation theory.
  33. [33]
    Mellin's Transform and Application to Some Time Series Models
    Feb 18, 2014 · Thus, the Mellin transform has significant applications in probability theory, Markov chains, renewal theory, and time series.
  34. [34]
    The Mellin transform for moment–generation and for the probability ...
    The Mellin transform for moment–generation and for the probability density of products and quotients of random variables. Published in: Proceedings of the ...
  35. [35]
    [PDF] Mellin-Barnes integrals for stable distributions and their convolutions
    In this expository paper we first survey the method of Mellin-Barnes integrals to represent the α stable Lévy distributions in probability theory.
  36. [36]
    [PDF] On the Method of Logarithmic Cumulants for Parametric Probability ...
    May 6, 2013 · Based on the Mellin transform, the MoLC approach can be considered an alternative to MoM that is both more robust to outliers and in several ...
  37. [37]
    On elementary functions of Pareto variables | Metrika
    The Mellin transform with its convolution and exponentiation properties is utilized to that end. ... Krishnaji, N.: Characterization of the Pareto Distribution ...
  38. [38]
    The moment generating function of a bivariate gamma-type distribution
    Aug 15, 2012 · The moment generating function of a bivariate gamma-type distribution ... Mellin transform and transformation of variable techniques. The ...
  39. [39]
    [PDF] Application of the Mellin Transforin to Boundary Value Problems
    The Mellin transform is investigated with special emphasis on its applications to the solution of boundary value problems. A technique is given for solution of ...
  40. [40]
    [PDF] Well-Posedness and Regularity of the Heat Equation with Robin ...
    In this section we provide known properties of the Mellin transform and Laplace transform, which consti- tute an important tool in our analysis. It will be ...
  41. [41]
    [PDF] On Mellin-Barnes integral representations for GKZ hypergeometric ...
    Feb 14, 2018 · In this short paper, we discuss Mellin-Barnes integral representations of GKZ hyperge- ometric functions as well as its relation to Laplace ...