Fact-checked by Grok 2 weeks ago

Harmonic analysis

Harmonic analysis is a branch of mathematics that studies the representation of functions and signals as superpositions of basic waves, particularly through tools like Fourier series and Fourier transforms, and examines the quantitative properties of these representations under operations such as convolution and differentiation. It originated from efforts to solve partial differential equations, such as the heat equation, by decomposing solutions into harmonic components like sines and cosines. Central to the field is the Fourier transform, defined for integrable functions f \in L^1(\mathbb{R}) as \hat{f}(\xi) = \int_{-\infty}^{\infty} f(x) e^{-2\pi i x \xi} \, dx, which maps spatial domain functions to their frequency domain counterparts and reveals properties like decay related to smoothness. The scope of harmonic analysis extends beyond classical Fourier theory to more general settings, including non-abelian groups, Hilbert spaces, and modern applications in , where it enables filtering and via . Key concepts include L^p norms for measuring function sizes, such as \|f\|_{L^p} = \left( \int |f(x)|^p \, dx \right)^{1/p} for $1 \leq p < \infty, and the Riemann-Lebesgue lemma, which states that the Fourier transform of an L^1 function vanishes at infinity. In partial differential equations, harmonic analysis provides methods for boundary value problems by leveraging Fourier methods to solve equations like the wave or Laplace equation. Broader applications appear in number theory, probability, and geometry, often using techniques like oscillatory integrals and covering lemmas to bound function behaviors. Notable developments include extensions to locally compact abelian groups, where the Fourier transform generalizes to Pontryagin duality, linking the group to its dual for representation theory. The field also intersects with functional analysis through spaces like L^2, where Plancherel's theorem equates L^2 norms in time and frequency domains, \|f\|_{L^2} = \|\hat{f}\|_{L^2}. In engineering and physics, harmonic analysis underpins sampling theorems, such as the Nyquist-Shannon theorem, ensuring accurate signal reconstruction from discrete samples. Overall, it provides a framework for decomposing complex objects into spectral components, facilitating analysis across diverse mathematical and applied domains.

Introduction and Fundamentals

Definition and Scope

Harmonic analysis is a branch of mathematics concerned with the representation of functions or signals as superpositions of basic waves, initially but generalized to eigenfunctions of linear operators defined on groups or spaces. This field examines how such decompositions allow for the breakdown of complex structures into simpler oscillatory components, facilitating deeper insights into their properties. The core objectives of harmonic analysis include the decomposition of functions into harmonic expansions, their reconstruction from these components, and the subsequent analysis of their behavior, with particular emphasis on issues of convergence, uniqueness, and inversion formulas. These goals enable the quantification of function properties under transformations, such as translations or convolutions, often through norms and inequalities that preserve or reveal structural features. A foundational example is the , which decomposes periodic functions on the circle into sums of sines and cosines. The scope of harmonic analysis encompasses classical settings on Euclidean spaces, where tools like Fourier transforms dominate, as well as abstract frameworks on topological groups, extending to non-commutative structures via representation theory. It also includes applied dimensions, such as signal processing and the solution of partial differential equations (), where harmonic methods model wave propagation and diffusion. Key connections arise with functional analysis through spaces of integrable functions, operator theory via spectral decompositions, and approximation theory in estimating error bounds for expansions. Basic prerequisites involve familiarity with inner product spaces, where orthogonality ensures the independence of basis elements in decompositions, without requiring detailed derivations here.

Historical Development

The origins of harmonic analysis can be traced to ancient civilizations, where early explorations of periodic phenomena in music and astronomy laid foundational concepts. Babylonian astronomers developed sophisticated mathematical tables to predict celestial events, such as lunar eclipses and planetary positions, using sexagesimal arithmetic to model periodic cycles. In ancient Greece, Pythagoras (c. 570–495 BCE) advanced the study of harmonics by discovering that consonant musical intervals arise from simple integer ratios of string lengths, such as 2:1 for the octave and 3:2 for the fifth, linking music theory to mathematical proportions. This work influenced later Greek scholars, including Ptolemy (c. 100–170 CE), who in his treatise Harmonics integrated musical theory with astronomy, creating tables that connected harmonic ratios to cosmic structures and emphasizing their shared mathematical basis. These ancient efforts were motivated by the desire to understand and predict natural rhythms, from musical tones to heavenly motions. In the 18th century, the foundations of modern harmonic analysis emerged through investigations into physical vibrations, driven by problems in acoustics and mechanics. Jean le Rond d'Alembert introduced the one-dimensional wave equation in 1747 to model the transverse vibrations of a string, providing an analytical framework for propagating waves. Daniel Bernoulli, building on earlier ideas, proposed series solutions in 1753, representing string vibrations as superpositions of harmonic modes with frequencies that are integer multiples of a fundamental tone. Leonhard Euler expanded this in his 1748 paper "Sur la vibration des cordes," applying d'Alembert's equation to derive solutions using trigonometric series and addressing initial controversies over the nature of wave propagation. These developments were spurred by the need to solve real-world problems, such as the behavior of musical instruments and elastic bodies. The 19th century saw a pivotal breakthrough with Joseph Fourier's 1822 publication Théorie Analytique de la Chaleur, where he introduced to represent arbitrary functions as sums of sines and cosines, motivated by modeling heat conduction in solids. Initially met with controversy—Lagrange and Laplace criticized the convergence of these series for non-smooth functions, and Biot disputed priority over heat transfer models—the approach gained acceptance by the 1880s as its utility in solving partial differential equations became evident. The 20th century brought rigorous mathematical expansions, addressing convergence issues and generalizing to abstract settings. Henri Lebesgue's development of the integral in 1902 enabled precise analysis of Fourier series convergence, proving term-by-term integration for bounded integrable functions and resolving earlier Riemann integral limitations for discontinuous cases. Norbert Wiener advanced the field in the 1930s with his tauberian theorems, published in 1932, which linked Fourier analysis to broader inversion problems in harmonic settings. Post-World War II, Lev Pontryagin's work in the 1930s–1940s on topological groups introduced Pontryagin duality for locally compact abelian groups, shifting harmonic analysis toward abstract structures like characters on non-Euclidean spaces. Throughout, these advancements were propelled by applications to physical phenomena, including heat flow, wave propagation, and celestial mechanics.

Classical Harmonic Analysis

Fourier Series

Fourier series provide a method to represent periodic functions as infinite sums of trigonometric functions, fundamental to classical harmonic analysis. For a periodic function f(x) with period $2\pi, the Fourier series expansion is given by f(x) = \frac{a_0}{2} + \sum_{n=1}^\infty \left( a_n \cos(nx) + b_n \sin(nx) \right), where the coefficients are determined by integrals over one period. This representation decomposes the function into its constituent frequencies, leveraging the periodicity to express complex waveforms through simpler harmonic components. The Fourier coefficients are computed as a_n = \frac{1}{\pi} \int_{-\pi}^\pi f(x) \cos(nx) \, dx, \quad n = 0, 1, 2, \dots, b_n = \frac{1}{\pi} \int_{-\pi}^\pi f(x) \sin(nx) \, dx, \quad n = 1, 2, \dots. In complex form, the series becomes f(x) = \sum_{n=-\infty}^\infty c_n e^{i n x}, with coefficients c_n = \frac{1}{2\pi} \int_{-\pi}^\pi f(x) e^{-i n x} \, dx. These formulas arise from the orthogonality of the trigonometric basis functions over [-\pi, \pi]. Orthogonality underpins the derivation of the coefficients and ensures the uniqueness of the expansion. The functions $1, \cos(nx), \sin(nx) for n \geq 1 form an orthogonal set with respect to the inner product \langle g, h \rangle = \int_{-\pi}^\pi g(x) h(x) \, dx. Specifically, \int_{-\pi}^\pi \cos(nx) \cos(mx) \, dx = \begin{cases} 0 & n \neq m, \\ 2\pi & n = m = 0, \\ \pi & n = m \geq 1, \end{cases} and similarly for sines, with cross terms vanishing. This follows from trigonometric product-to-sum identities, such as \cos(nx) \cos(mx) = \frac{1}{2} [\cos((n+m)x) + \cos((n-m)x)], whose integrals over [-\pi, \pi] are zero unless n = m. For the complex exponentials, \int_{-\pi}^\pi e^{i(n-m)x} \, dx = 2\pi \delta_{n m}. Convergence of the Fourier series depends on the regularity of f. Under Dirichlet conditions—f is periodic with period $2\pi, bounded, has finitely many maxima and minima, and finitely many discontinuities in each period—the series converges pointwise to f(x) at points of continuity and to the average of the left and right limits at jump discontinuities. For square-integrable functions (f \in L^2[-\pi, \pi]), the series converges in the L^2 norm to f, as established by Lebesgue's theorem, with the partial sums forming a complete orthogonal basis in the Hilbert space. Fejér's theorem further guarantees uniform convergence of the Cesàro means (averages of partial sums) to f for continuous periodic functions. A notable limitation is the Gibbs phenomenon, observed near discontinuities. For a jump of height d, the partial sums overshoot by approximately $0.089 d (or 9% of the jump), with the overshoot persisting and its location approaching the discontinuity as more terms are added. This arises from the slow decay of coefficients for discontinuous functions, like the $1/n decay for a square wave. Consider the square wave function defined as f(x) = -1 for -\pi \leq x < 0 and f(x) = 1 for $0 < x \leq \pi, extended periodically. Its Fourier series is f(x) \sim \frac{4}{\pi} \sum_{k=1}^\infty \frac{\sin((2k-1)x)}{2k-1}, which converges to f(x) except at discontinuities, where it equals zero, and exhibits the Gibbs overshoot near x = 0, \pm \pi. This example illustrates how odd harmonics capture the antisymmetric nature of the function. Extensions to arbitrary intervals [a, b] with period L = b - a scale the basis functions accordingly: f(x) = \frac{a_0}{2} + \sum_{n=1}^\infty \left( a_n \cos\left(\frac{2 n \pi x}{L}\right) + b_n \sin\left(\frac{2 n \pi x}{L}\right) \right), with a_n = \frac{2}{L} \int_a^b f(x) \cos\left(\frac{2 n \pi x}{L}\right) \, dx, \quad b_n = \frac{2}{L} \int_a^b f(x) \sin\left(\frac{2 n \pi x}{L}\right) \, dx. This adaptation preserves orthogonality and convergence properties for periodic extensions. For non-periodic functions, generalize to , handling aperiodic cases continuously.

Fourier Transforms

The Fourier transform provides a fundamental tool for analyzing aperiodic functions on Euclidean spaces by decomposing them into their frequency components via an integral representation. For a function f \in L^1(\mathbb{R}^n), the Fourier transform \hat{f} is defined as \hat{f}(\xi) = \int_{\mathbb{R}^n} f(x) e^{-2\pi i x \cdot \xi} \, dx, where x \cdot \xi denotes the dot product, and this convention employs the angular frequency variable \xi without a scaling factor in the exponent. This normalization, often called the unitary or angular frequency convention, ensures that the transform is dimensionless and facilitates certain symmetry properties in L^2 spaces. Alternative conventions exist, such as the non-unitary form \hat{f}(\omega) = \int f(x) e^{-i \omega \cdot x} \, dx using ordinary frequency \omega = 2\pi \xi, which appears in physics applications but alters the inversion formula by a factor of $2\pi. Under suitable regularity conditions, the Fourier transform is invertible, recovering the original function from its transform. Specifically, for functions in the Schwartz space \mathcal{S}(\mathbb{R}^n) of smooth, rapidly decaying functions and their derivatives, the inversion theorem states that f(x) = \int_{\mathbb{R}^n} \hat{f}(\xi) e^{2\pi i x \cdot \xi} \, d\xi. This holds pointwise and in the L^1 norm for Schwartz functions, with the integral converging absolutely due to the rapid decay of \hat{f}. The theorem extends to more general classes, such as L^1 functions whose transforms are also in L^1, via limits or principal value interpretations. The Fourier transform exhibits several key properties that underpin its utility in analysis. It is linear, meaning \widehat{\alpha f + \beta g} = \alpha \hat{f} + \beta \hat{g} for scalars \alpha, \beta and functions f, g. Translation in the spatial domain corresponds to modulation in the frequency domain: if g(x) = f(x - a), then \hat{g}(\xi) = e^{-2\pi i a \cdot \xi} \hat{f}(\xi). Similarly, modulation in space yields translation in frequency: if h(x) = e^{2\pi i b \cdot x} f(x), then \hat{h}(\xi) = \hat{f}(\xi - b). The convolution theorem states that the transform of a convolution (f * g)(x) = \int f(y) g(x - y) \, dy is the product of the transforms: \widehat{f * g} = \hat{f} \hat{g}, assuming the integrals converge appropriately. For square-integrable functions, the Plancherel theorem extends the Parseval identity from Fourier series to the continuous case, establishing that the Fourier transform is an isometry on L^2(\mathbb{R}^n). Precisely, for f \in L^2(\mathbb{R}^n), \|f\|_2 = \|\hat{f}\|_2, where the norms are the standard L^2 norms, and the transform is extended by density from the . This preserves energy and enables the inversion formula in the L^2 sense. The Fourier transform extends naturally to higher dimensions on \mathbb{R}^n, where the definition remains \hat{f}(\xi) = \int_{\mathbb{R}^n} f(x) e^{-2\pi i x \cdot \xi} \, dx, \quad \xi \in \mathbb{R}^n, and all listed properties hold analogously, with convolutions and dot products in \mathbb{R}^n. Due to the separability of the exponential, the multidimensional transform factors into a product of one-dimensional transforms along each coordinate axis for functions that are products themselves. Illustrative examples highlight these features. The Fourier transform of a Gaussian f(x) = e^{-\pi |x|^2} on \mathbb{R}^n is itself: \hat{f}(\xi) = e^{-\pi |\xi|^2}, demonstrating self-duality under this normalization and following from completing the square in the integral or using the one-dimensional case iteratively. In the sense of distributions, the Fourier transform of the Dirac delta \delta at the origin is the constant function 1, as \langle \hat{\delta}, \phi \rangle = \phi(0) for test functions \phi, reflecting that \delta contains all frequencies equally; more generally, \widehat{\delta_a}(\xi) = e^{-2\pi i a \cdot \xi}, obtained as a limit of narrow Gaussians.

Abstract Harmonic Analysis

Analysis on Locally Compact Groups

Harmonic analysis on locally compact groups extends the classical theory from Euclidean spaces to more general topological settings, providing a framework for studying convolutions, representations, and transforms on abstract groups. A locally compact Hausdorff abelian group G is a topological group where every point has a compact neighborhood, ensuring the existence of a unique (up to scalar multiple) left-invariant measure known as the \mu. This measure, introduced by in 1933, allows for the integration of functions over G and underpins the construction of function spaces central to the analysis. The spaces L^1(G) and L^2(G) are defined as the completions of continuous functions with compact support under the norms \|f\|_1 = \int_G |f(x)| \, d\mu(x) and \|f\|_2 = \left( \int_G |f(x)|^2 \, d\mu(x) \right)^{1/2}, respectively; these are Banach and Hilbert spaces, respectively, enabling the study of integrable and square-integrable functions on G. Convolution on L^1(G) is defined by (f * g)(x) = \int_G f(y) g(x - y) \, d\mu(y) for f, g \in L^1(G), which extends the Euclidean convolution and satisfies \|f * g\|_1 \leq \|f\|_1 \|g\|_1, making it a continuous operation. This operation generalizes classical convolutions, such as those on \mathbb{R}^n, to arbitrary locally compact abelian groups. Approximate identities play a crucial role in regularization and approximation within this framework; these are nets \{u_\alpha\} in L^1(G) with \|u_\alpha\|_1 = 1 such that \lim_\alpha (u_\alpha * f)(x) = f(x) for all f \in L^1(G) and continuous functions with compact support, approximating the . The algebra L^1(G) forms a commutative under convolution and the L^1-norm, with the involution \tilde{f}(x) = \overline{f(-x)} making it a . By for commutative Banach algebras, the spectrum of L^1(G) consists of its maximal ideals, which correspond to the continuous characters of G, homomorphisms from G to the multiplicative group of nonzero complex numbers. Prominent examples illustrate this setup: the real line \mathbb{R} with Lebesgue measure recovers classical Fourier analysis; the integers \mathbb{Z} under addition with counting measure yields discrete Fourier series; and finite abelian groups, equipped with discrete uniform measure (normalized to total mass 1), reduce to finite Fourier transforms on cyclic groups. These cases demonstrate how the general theory unifies and generalizes familiar harmonic tools.

Pontryagin Duality and Characters

In the context of abstract harmonic analysis on locally compact abelian groups, characters play a fundamental role as the building blocks for the Fourier transform. A character of a locally compact abelian group G is defined as a continuous group homomorphism \chi: G \to \mathbb{C}^*, where \mathbb{C}^* denotes the multiplicative group of nonzero complex numbers, but more precisely, the image lies in the unit circle S^1 = \{ z \in \mathbb{C} : |z| = 1 \} due to the unitarity requirement in harmonic analysis. These characters are continuous unitary representations of G into the circle group, and the set of all such characters, denoted \hat{G}, forms the Pontryagin dual group of G. The group operation on \hat{G} is pointwise multiplication: (\chi_1 \cdot \chi_2)(x) = \chi_1(x) \chi_2(x) for \chi_1, \chi_2 \in \hat{G} and x \in G, equipped with the compact-open topology, which ensures \hat{G} is a topological group. The Pontryagin duality theorem establishes a profound symmetry between a locally compact abelian group G and its dual \hat{G}. Specifically, if G is locally compact and abelian, then \hat{G} is also locally compact and abelian, and there exists a canonical topological isomorphism G \cong \hat{\hat{G}}, where \hat{\hat{G}} is the dual of the dual, known as the bidual. This isomorphism is given by the evaluation map \Phi: G \to \hat{\hat{G}} defined by \Phi(x)(\chi) = \chi(x) for x \in G and \chi \in \hat{G}, which is a continuous homomorphism that is both injective and surjective, hence an isomorphism of topological groups. The theorem, originally proved by Lev Pontryagin in his foundational work on topological groups, extends the classical duality between position and momentum spaces in physics to arbitrary locally compact abelian settings, enabling a unified framework for . This double dual identification \hat{\hat{G}} \cong G highlights the self-duality inherent in the structure of such groups under character theory. Building on Pontryagin duality, the Fourier transform on a locally compact abelian group G is defined using the dual group \hat{G} and a Haar measure \mu on G. For a function f \in L^1(G), the Fourier transform is given by \hat{f}(\chi) = \int_G f(x) \overline{\chi(x)} \, d\mu(x), \quad \chi \in \hat{G}, where \overline{\chi(x)} is the complex conjugate, ensuring the transform is well-defined and captures the projection onto characters. This generalizes the classical Fourier transform on \mathbb{R} or the Fourier series on the circle. A normalized Haar measure on \hat{G}, denoted \hat{\mu}, is chosen compatibly via duality to facilitate inversion. For functions f \in L^1(G) \cap L^2(G), the inversion formula recovers f almost everywhere: f(x) = \int_{\hat{G}} \hat{f}(\chi) \chi(x) \, d\hat{\mu}(\chi), provided the integral converges appropriately, often requiring additional regularity conditions like continuity of f. The abstract Plancherel theorem extends the L^2-isometry property to this setting: the Fourier transform defines a unitary isomorphism L^2(G, \mu) \to L^2(\hat{G}, \hat{\mu}), preserving norms via \|f\|_2 = \|\hat{f}\|_2, with the dual measure ensuring the Parseval relation \langle f, g \rangle = \langle \hat{f}, \hat{g} \rangle. This theorem, proved using the Riesz-Fischer representation and density arguments for compactly supported continuous functions, underpins the spectral decomposition in group algebras. Illustrative examples demonstrate the versatility of Pontryagin duality across different group structures. For the additive group \mathbb{R} of real numbers with Lebesgue measure, the dual \hat{\mathbb{R}} is isomorphic to \mathbb{R} itself, with characters given by \chi_\xi(x) = e^{2\pi i \xi x} for \xi \in \mathbb{R}, recovering the standard Fourier transform on \mathbb{R}. Similarly, the dual of the discrete group \mathbb{Z} of integers is the circle group \mathbb{T} = \mathbb{R}/\mathbb{Z}, with characters \chi_\theta(n) = e^{2\pi i \theta n} for \theta \in [0,1), corresponding to Fourier series expansions. For finite abelian groups, such as \mathbb{Z}/n\mathbb{Z}, the dual is isomorphic to the group itself, with characters forming a complete set of nth roots of unity, enabling discrete Fourier analysis on cyclic groups and their products. These examples underscore how duality interchanges compactness and discreteness: compact groups have discrete duals, and vice versa. Vanishing theorems in this framework reveal deep constraints on the supports of functions and their transforms, arising from the rigidity of . A fundamental result states that if both f \in L^1(G) and \hat{f} have compact support, then f = 0 almost everywhere, unless G is both compact and discrete (i.e., finite), as the duality map cannot preserve compactness in both directions otherwise. More generally, if \hat{f} vanishes on an open set in \hat{G}, the support of f cannot be too concentrated, leading to that bound the product of the measures of \operatorname{supp}(f) and \operatorname{supp}(\hat{f}). These theorems, often derived from the and properties of , have implications for the localization of harmonic components in group representations. Convolution on G, defined as (f * g)(x) = \int_G f(y) g(x - y) \, d\mu(y), interacts with the Fourier transform via \widehat{f * g} = \hat{f} \cdot \hat{g}, further highlighting support restrictions through multiplicative structure on \hat{G}.

Applications in Mathematics and Physics

Partial Differential Equations

Harmonic analysis provides powerful tools for solving partial differential equations (PDEs) by decomposing solutions into frequency components, particularly for elliptic and hyperbolic equations on bounded or unbounded domains. For boundary value problems on finite domains, such as the Laplace equation \Delta u = 0 on a rectangle [0, a] \times [0, b] with Dirichlet boundary conditions u(x,0) = f(x), u(x,b) = u(0,y) = u(a,y) = 0, the method of separation of variables assumes u(x,y) = X(x)Y(y), leading to eigenvalue problems whose solutions are sine and cosine eigenfunctions. The general solution is then expressed as a Fourier sine series expansion u(x,y) = \sum_{n=1}^\infty b_n \sin\left(\frac{n\pi x}{a}\right) \sinh\left(\frac{n\pi (b-y)}{a}\right), where coefficients b_n are determined by the initial boundary data via Fourier coefficients \int_0^a f(x) \sin\left(\frac{n\pi x}{a}\right) dx. This approach leverages the completeness of Fourier eigenfunctions to represent solutions efficiently. On unbounded domains, the Fourier transform extends this methodology to initial value problems, exemplified by the heat equation \partial_t u = \Delta u on \mathbb{R}^n \times (0,\infty) with initial condition u(0,x) = g(x). Applying the Fourier transform in space yields \hat{u}(t,\xi) = \hat{g}(\xi) e^{-|\xi|^2 t}, and inverting the transform gives the solution u(t,x) = (4\pi t)^{-n/2} \int_{\mathbb{R}^n} g(y) e^{-|x-y|^2/(4t)} dy, which diffuses the initial data according to the Gaussian kernel. This transform method simplifies the PDE to an ordinary differential equation in frequency space, highlighting the smoothing effect of the heat operator. For hyperbolic equations like the wave equation \partial_t^2 u = c^2 \Delta u on \mathbb{R}^n \times (0,\infty) with initial data u(0,x) = \phi(x), \partial_t u(0,x) = \psi(x), the Fourier transform leads to \hat{u}(t,\xi) = \hat{\phi}(\xi) \cos(c|\xi| t) + c^{-1} \hat{\psi}(\xi) \sin(c|\xi| t) / |\xi|, whose inversion recovers d'Alembert's formula in one dimension as u(t,x) = [\phi(x+ct) + \phi(x-ct)]/2 + (2c)^{-1} \int_{x-ct}^{x+ct} \psi(y) dy. Similarly, the Poisson equation -\Delta u = f on \mathbb{R}^n solves via \hat{u}(\xi) = \hat{f}(\xi) / |\xi|^2, with inversion yielding u(x) = c_n \int_{\mathbb{R}^n} f(y) |x-y|^{2-n} dy for n \geq 3, the fundamental solution for the Laplacian. Green's functions further integrate harmonic analysis into PDE solutions, representing responses to point sources via Fourier integrals. For the Helmholtz equation (\Delta + k^2) u = f on \mathbb{R}^n, the Green's function is G(x,y;k) = (4\pi)^{-n/2} \int_{\mathbb{R}^n} e^{i\xi \cdot (x-y)} (|\xi|^2 - k^2)^{-1} d\xi, which for outgoing waves takes the form of the Hankel function in three dimensions, G(x,y;k) = e^{ik|x-y|}/(4\pi |x-y|), enabling the solution u(x) = \int_{\mathbb{R}^n} G(x,y;k) f(y) dy. This Fourier representation facilitates numerical and asymptotic analysis for wave propagation problems. In the context of function spaces, harmonic analysis defines Sobolev spaces H^s(\mathbb{R}^n) via Fourier multipliers, where the norm is \|f\|_{H^s} = \left( \int_{\mathbb{R}^n} (1 + |\xi|^2)^s |\hat{f}(\xi)|^2 d\xi \right)^{1/2}, capturing the L^2 integrability of up to s derivatives. This framework underpins elliptic regularity theory, where solutions to Lu = f with elliptic operator L gain smoothness: if f \in H^s, then u \in H^{s+2} locally, proved using multiplier theorems like Mihlin's, which bound operators T_m f = \mathcal{F}^{-1}(m(\xi) \hat{f}(\xi)) for symbols m satisfying |\partial^\alpha m(\xi)| \lesssim |\xi|^{-|\alpha|} for |\alpha| \leq n+1. Such results ensure higher derivatives of solutions are bounded in appropriate norms, essential for proving existence and uniqueness in weak formulations.

Representation Theory and Quantum Mechanics

In harmonic analysis, unitary representations provide a framework for decomposing functions on groups into components that reflect the group's symmetry. A unitary representation of a topological group G is a continuous homomorphism \pi: G \to U(\mathcal{H}), where U(\mathcal{H}) is the group of unitary operators on a complex Hilbert space \mathcal{H}, such that the inner product on \mathcal{H} is preserved. These representations are central to extending Fourier analysis from abelian to non-abelian groups, as they allow the regular representation on L^2(G) to be decomposed into irreducible unitary representations (irreps), which are indecomposable under the group action. For compact groups G, the Peter-Weyl theorem establishes that the Hilbert space L^2(G) admits a direct sum decomposition into finite-dimensional irreps, with the matrix coefficients of these irreps forming an orthonormal basis for L^2(G). Specifically, if \{\pi_j\} are the irreps of G, then L^2(G) = \bigoplus_j (\pi_j \otimes \overline{\pi_j}), where the multiplicity of each irrep equals its dimension, enabling a non-abelian analogue of expansion. This theorem, proved by in the 1920s, underpins harmonic analysis on compact by providing an explicit basis for function spaces. A key tool in this decomposition is Schur orthogonality, which quantifies the independence of distinct irreps. For two irreps \pi and \sigma of a compact group G, the characters satisfy \int_G \overline{\chi_\pi(g)} \chi_\sigma(g) \, dg = \delta_{\pi\sigma} \frac{1}{\dim \pi}, where \chi_\pi(g) = \mathrm{tr}(\pi(g)) is the character and dg is the normalized . More generally, the matrix coefficients \langle \pi(g) v_i, v_j \rangle for orthonormal bases \{v_i\} in the representation spaces are orthogonal across different irreps, ensuring the uniqueness of the decomposition. These relations, derived from on intertwiners, facilitate the computation of integrals and projections in representation theory. The Plancherel formula extends this to non-abelian settings by providing an L^2-norm preservation via the Fourier transform on the dual. For a compact group G, the Plancherel measure on the unitary dual \widehat{G} (the set of irreps, equipped with a discrete counting measure weighted by dimensions) yields \|f\|_{L^2(G)}^2 = \sum_{\pi \in \widehat{G}} \dim(\pi) \|\widehat{f}(\pi)\|_{HS}^2, where \widehat{f}(\pi) is the Fourier coefficient operator in the irrep \pi, and \|\cdot\|_{HS} is the Hilbert-Schmidt norm. This formal sum over irreps replaces the integral over characters in the abelian case, confirming the completeness of the representation decomposition. In quantum mechanics, these concepts link directly to operator algebras and symmetry groups. The Stone-von Neumann theorem asserts that all irreducible unitary representations of the Heisenberg group H^n (the nilpotent group modeling phase space translations) on a Hilbert space are equivalent to the Schrödinger representation, where position Q and momentum P operators satisfy [Q_j, P_k] = i \hbar \delta_{jk}. This uniqueness ensures a canonical quantization of classical phase space, foundational for wave mechanics. Weyl quantization further connects Fourier analysis to quantum operators by associating a classical symbol a(x, \xi) on phase space to the operator Op^W(a) via the Fourier-Wigner transform, preserving the symplectic structure and enabling pseudodifferential analysis. For instance, representations of SU(2) classify angular momentum states in quantum mechanics, where the irreps labeled by half-integers j correspond to spin-j particles, with raising and lowering operators acting within finite-dimensional spaces of dimension $2j+1. Similarly, Fourier analysis on the Heisenberg group decomposes phase-space functions into twisted convolutions, modeling quantum evolution beyond abelian symmetries.

Modern Extensions and Branches

Wavelet Theory

Wavelet theory extends classical harmonic analysis by providing bases of functions that are localized in both time and scale, offering superior analysis for non-stationary signals compared to the global frequency localization of . This framework decomposes signals into components that capture transient features, with applications spanning , , and partial differential equations. Seminal developments trace back to the construction of orthogonal , enabling efficient representations that preserve energy while allowing sparse approximations. The core of wavelet theory lies in generating a family of functions from a single mother wavelet \psi \in L^2(\mathbb{R}) through dyadic dilations and translations, defined as \psi_{j,k}(x) = 2^{j/2} \psi(2^j x - k), \quad j,k \in \mathbb{Z}. These form an orthonormal basis for L^2(\mathbb{R}) under suitable conditions on \psi, such as integrability and zero mean. The structure is underpinned by multiresolution analysis (MRA), which organizes the space into nested subspaces V_j approximated by dilations of a scaling function \phi, satisfying \overline{\text{span}}\{\phi(2^j x - k) : k \in \mathbb{Z}\} = V_j with V_{j+1} = V_j \oplus W_j, where W_j is the wavelet subspace. The scaling function \phi obeys a two-scale equation \phi(x) = \sqrt{2} \sum_h h_n \phi(2x - n), ensuring hierarchical refinement. For continuous analysis, the continuous wavelet transform (CWT) of a function f \in L^2(\mathbb{R}) is given by W_f(a,b) = \frac{1}{\sqrt{|a|}} \int_{-\infty}^{\infty} f(x) \overline{\psi\left(\frac{x-b}{a}\right)} \, dx, \quad a \neq 0, \, b \in \mathbb{R}. Invertibility requires the mother wavelet to satisfy the admissibility condition \int_{-\infty}^{\infty} \frac{|\hat{\psi}(\xi)|^2}{|\xi|} \, d\xi < \infty, where \hat{\psi} is the Fourier transform of \psi, ensuring zero mean in frequency and allowing perfect reconstruction via an integral over the scale-position plane. This condition guarantees that the CWT acts as an isometry up to a constant, preserving the L^2-norm of f. The Mexican hat wavelet, derived from the second derivative of a Gaussian, exemplifies an admissible function suitable for detecting singularities. The discrete wavelet transform (DWT) discretizes this framework for computational efficiency, using dyadic scales a = 2^{-j} and integer translations b = k. Orthogonal bases are constructed via the , a pyramid scheme applying low-pass (scaling) and high-pass (wavelet) filters iteratively for decomposition and reconstruction. Perfect reconstruction is achieved with quadrature mirror filters satisfying |H(\omega)|^2 + |H(\omega + \pi)|^2 = 2, where H is the low-pass filter. The simplest example is the , defined as \psi(x) = 1 for $0 \leq x < 1/2 and -1 for $1/2 \leq x < 1, zero elsewhere, ideal for representing piecewise constant functions like step discontinuities. More sophisticated are , which balance compact support and smoothness through filters of length $2N yielding N-1 vanishing moments. Key properties of wavelets include compact support, which limits their influence to finite intervals, and vanishing moments, where \int x^m \psi(x) \, dx = 0 for m = 0, \dots, M-1, allowing annihilation of polynomials up to degree M-1 and efficient approximation of smooth signals. These enable signal compression by thresholding small coefficients, retaining significant features while discarding noise, with reconstruction error decaying rapidly for functions in . In image denoising, for instance, wavelet coefficients of noisy images are soft-thresholded, exploiting sparsity to recover clean signals with minimal mean squared error. Wavelets integrate deeply with harmonic analysis, serving as kernels for Calderón-Zygmund singular integral operators, where their dyadic structure approximates smooth symbols and bounds operator norms on L^p spaces. They also embody , partitioning the frequency domain into dyadic annuli via wavelet filters, facilitating proofs of square function estimates and boundedness of maximal operators. In the fine-scale limit, wavelet bases recover the frequency-localized behavior of .

Time-Frequency Analysis

Time-frequency analysis provides joint representations of signals in both time and frequency domains, enabling the study of non-stationary phenomena where frequency content varies over time. Unlike the global Fourier transform, which assumes stationarity, time-frequency methods localize the analysis, revealing how spectral components evolve. This approach is essential in signal processing for applications such as radar, speech analysis, and seismology, where traditional Fourier methods fail to capture transient behaviors. Seminal developments include linear and quadratic representations that balance resolution and artifacts. The short-time Fourier transform (STFT) is a foundational linear time-frequency representation, defined as V_g f(t, \omega) = \int_{-\infty}^{\infty} f(x) \overline{g(x - t)} e^{-i \omega x} \, dx, where g is a window function, often a Gaussian for optimal localization due to its minimal uncertainty properties. Introduced by in 1946, the STFT applies the Fourier transform to windowed segments of the signal, producing a two-dimensional time-frequency plane. The choice of window g trades off time and frequency resolution, as shorter windows enhance time localization at the cost of frequency smearing, governed by the . Gabor analysis extends the STFT to discrete frames generated by modulations and translations of a window function, forming overcomplete bases for signal representation in the time-frequency plane. These Gabor frames, consisting of elements g_{m,n}(x) = e^{2\pi i n b x} g(x - m a) for lattice parameters a, b, allow flexible analysis with redundancy that improves robustness over orthonormal bases. However, the Balian-Low theorem imposes fundamental limits on time-frequency concentration: for a Gabor system to form a Riesz basis of L^2(\mathbb{R}), the window g cannot be well-localized in both time and frequency, specifically \int |t g(t)|^2 \, dt \cdot \int |\hat{g}(\omega)|^2 \omega^2 \, d\omega = \infty. Originally formulated by Balian in 1981 and independently by Low in 1985, this theorem underscores the incompatibility of perfect localization and non-redundancy, influencing designs in communications and optics. The Wigner-Ville distribution offers a quadratic alternative, defined as W_f(t, \omega) = \int_{-\infty}^{\infty} f\left(t + \frac{\tau}{2}\right) \overline{f\left(t - \frac{\tau}{2}\right)} e^{-i \omega \tau} \, d\tau, providing high-resolution energy density in the time-frequency plane without a window parameter. Originating from in quantum mechanics and adapted by Ville in 1948 for signal analysis, it satisfies desirable properties like marginals matching the instantaneous energy and spectrum. However, its bilinear nature introduces cross-terms—oscillatory artifacts between signal components—that obscure interpretation for multicomponent signals. These interferences, located midway between auto-terms in the time-frequency plane, necessitate smoothing techniques. Pseudo-differential operators in the time-frequency context formalize operations on these representations via symbols \sigma(t, \xi), with the Weyl quantization \mathrm{Op}(\sigma) f(x) = \iint \sigma\left( \frac{x+y}{2}, \xi \right) f(y) e^{2\pi i (x-y) \xi} \, dy \, d\xi, though often simplified to the form \int \sigma(t, \xi) \hat{f}(\xi) e^{2\pi i t \xi} \, d\xi for analysis. Developed by in 1979, this Weyl form ensures covariance under time-frequency shifts, linking operator symbols to phase-space functions and facilitating microlocal analysis. Such operators model filtering and propagation in non-stationary media, preserving time-frequency structure. Extensions of the Heisenberg uncertainty principle to time-frequency representations quantify localization limits, such as |\mathrm{supp} V_g f| \cdot |\mathrm{supp} \hat{V}_g f| \geq c for some constant c > 0, reflecting the minimal area in the time-frequency plane occupied by any signal. The Balian-Low theorem provides a sharp version for Gabor systems, prohibiting compact support in both domains for basis-generating windows. These bounds, rooted in , ensure that perfect time-frequency atoms are impossible, guiding the design of approximate representations. For example, signals—frequency-modulated waves with instantaneous frequency \omega(t) = \alpha t + \beta—exemplify non-stationary behavior, where the STFT shows smeared ridges due to fixed window resolution, while the Wigner-Ville distribution yields sharp instantaneous frequency curves but with cross-terms if multicomponent. Cohen's class of distributions generalizes the Wigner-Ville via a \phi(t, \tau), as C_f(t, \omega) = \int W_f(t + u, \omega + v) \phi(u, v) \, du \, dv, allowing smoothing of cross-terms (e.g., via Gaussian kernels) at the expense of auto-term blurring. Introduced by in , this class unifies many distributions, enabling tailored analysis for in .

Key Theorems and Results

Plancherel and Parseval Theorems

provides a fundamental link between the L^2 of a and the \ell^2 of its coefficients for functions on . Specifically, for a f on [-\pi, \pi] with coefficients a_0 = \frac{1}{\pi} \int_{-\pi}^{\pi} f(x) \, dx and a_n = \frac{1}{\pi} \int_{-\pi}^{\pi} f(x) \cos(nx) \, dx, b_n = \frac{1}{\pi} \int_{-\pi}^{\pi} f(x) \sin(nx) \, dx for n \geq 1, \frac{1}{\pi} \int_{-\pi}^{\pi} |f(x)|^2 \, dx = \frac{a_0^2}{2} + \sum_{n=1}^\infty (a_n^2 + b_n^2). This identity, originally established in the context of orthogonal expansions, demonstrates that the trigonometric system forms a complete orthonormal basis in L^2([-\pi, \pi]), equating the energy of the function to that of its harmonic components. The Plancherel theorem extends this preservation of norms to the Fourier transform on \mathbb{R}. The Fourier transform \hat{f}(\xi) = \int_{-\infty}^\infty f(x) e^{-2\pi i x \xi} \, dx is initially defined for Schwartz functions, which are dense in L^2(\mathbb{R}). On this dense subspace, the transform satisfies \|\hat{f}\|_{L^2(\mathbb{R})} = \|f\|_{L^2(\mathbb{R})}. By the Riesz-Fischer theorem, which guarantees that an isometry from a dense subspace extends uniquely to a unitary operator on the full Hilbert space, the Fourier transform extends to an isometry on all of L^2(\mathbb{R}), yielding \|f\|_{L^2(\mathbb{R})} = \| \hat{f} \|_{L^2(\mathbb{R})}. This extension preserves the L^2 structure, confirming the completeness of the "harmonic" basis provided by the transform. In the abstract setting of locally compact abelian groups, the Plancherel theorem generalizes these results via . For a locally compact abelian group G with dual group \hat{G}, the maps L^1(G) \cap L^2(G) to continuous functions vanishing at infinity on \hat{G}, and extends to a unitary L^2(G) \cong L^2(\hat{G}), where the measure on \hat{G} is the induced by the dual. This is with respect to the on \hat{G}, ensuring that the L^2 norms are preserved: \|f\|_{L^2(G)}^2 = \int_{\hat{G}} |\hat{f}(\chi)|^2 \, d\mu(\chi), where \mu is the . The theorem relies on the theorem, which identifies G topologically with the double dual \hat{\hat{G}}. For non-abelian locally compact groups, the Plancherel theorem involves a decomposition into irreducible unitary representations rather than characters alone. The regular representation of G decomposes as a direct integral over the irreducible representations \pi in the dual \hat{G}, equipped with the Plancherel measure \mu: L^2(G) \cong \int^{\oplus}_{\hat{G}} \mathcal{H}_\pi \, d\mu(\pi), where \mathcal{H}_\pi is the Hilbert space of \pi. The Fourier transform \mathcal{F}f(\pi) is the operator \int_G f(g) \pi(g) \, dg, and the theorem states that \|f\|_{L^2(G)}^2 = \int_{\hat{G}} \|\mathcal{F}f(\pi)\|_{\mathrm{HS}}^2 \, d\mu(\pi), with the Hilbert-Schmidt norm on operators. This measure \mu is uniquely determined and supported on the square-integrable representations. These theorems interpret the as an energy-preserving operation, conserving the total "energy" \|f\|_2^2 of signals across domains, which underscores the completeness of harmonic decompositions in both classical and abstract settings. In applications to partial differential equations and , this preservation facilitates solving evolution equations by transforming to spectral domains. A variant for s highlights the full of characters. For a G, the irreducible characters \chi_i form an for the space of class functions under the inner product \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \overline{\chi(g)} \psi(g), with \langle \chi_i, \chi_j \rangle = \delta_{ij}. This , part of the Peter-Weyl theorem for s, implies that the characters diagonalize the , providing a complete harmonic analysis on G.

Sampling and Uncertainty Principles

In harmonic analysis, sampling theorems provide conditions under which a function can be uniquely reconstructed from its values at discrete points, particularly for bandlimited signals whose transforms are supported on a bounded interval. The Shannon-Nyquist sampling theorem states that if a f \in L^2(\mathbb{R}) is bandlimited to the frequency interval [-B, B], meaning its \hat{f}(\xi) = 0 for |\xi| > B, then f can be perfectly reconstructed from its uniform samples at rate $2B. Specifically, f(t) = \sum_{n=-\infty}^{\infty} f\left(\frac{n}{2B}\right) \operatorname{sinc}\left(2B\left(t - \frac{n}{2B}\right)\right), where \operatorname{sinc}(x) = \sin(\pi x)/(\pi x). This result, foundational to , ensures no information loss when sampling above the of $2B samples per unit time. Extensions to multidimensional settings address sampling on for bandlimited in higher , crucial for and . For a bandlimited to a compact set in \mathbb{R}^d, sampling on a \Lambda allows reconstruction if the satisfies certain conditions derived from the volume of the fundamental domain. Irregular sampling, governed by Beurling , requires the lower Beurling D^-(\Lambda) > 2B (in one dimension) for stable reconstruction, where D^-(\Lambda) = \liminf_{r \to \infty} \inf_{x \in \mathbb{R}} \#(\Lambda \cap B(x,r)) / (2r) measures the minimal local . These criteria generalize the uniform case and apply to non-uniform point sets in harmonic analysis. A key example illustrating bandlimited functions is the Paley-Wiener theorem, which characterizes them as s of exponential type. If f is bandlimited to [-B, B] and square-integrable on \mathbb{R}, then f extends to an on \mathbb{C} satisfying |f(z)| \leq C e^{2\pi B | \operatorname{Im} z |} for some constant C, and conversely, such functions restricted to \mathbb{R} are bandlimited. This property underpins sampling stability and links to in harmonic theory. Uncertainty principles in harmonic analysis quantify the inherent trade-off between time and frequency localization, asserting that a function and its Fourier transform cannot both be sharply concentrated. The Heisenberg uncertainty principle, in its variance formulation, states that for f \in L^2(\mathbb{R}) with \int |f(t)|^2 dt = 1 and finite variances \sigma_t^2 = \int (t - \mu_t)^2 |f(t)|^2 dt and \sigma_\xi^2 = \int (\xi - \mu_\xi)^2 |\hat{f}(\xi)|^2 d\xi, the inequality \sigma_t \sigma_\xi \geq 1/(4\pi) holds, with equality achieved for Gaussian functions f(t) = (2\pi \sigma_t^2)^{-1/4} e^{-(t-\mu_t)^2/(4\sigma_t^2)} e^{2\pi i \mu_\xi t}. This principle, analogous to its quantum mechanical counterpart, limits simultaneous resolution in time-frequency representations. In discrete settings, the Donoho-Stark uncertainty principle provides a support-based bound for vectors in \mathbb{C}^N. For a nonzero vector f \in \mathbb{C}^N with discrete Fourier transform \hat{f}, the product of the sizes of their supports satisfies |\operatorname{supp} f| \cdot |\operatorname{supp} \hat{f}| \geq N, where supports are the smallest subspaces containing f and \hat{f}. This discrete analog reveals phenomena absent in the continuous case, such as the possibility of perfect time-frequency localization for sparse signals, and has applications in and signal recovery. Abstract formulations extend these principles to locally compact abelian groups via , employing structures on G \times \hat{G}. For instance, on \mathbb{R}^d, the form \omega((x,\xi),(x',\xi')) = \xi \cdot x' - \xi' \cdot x induces a notion of compatible supports, leading to inequalities like |\operatorname{supp} f| \cdot |\operatorname{supp} \hat{f}| \geq 1 in measure-theoretic terms. More generally, covariance operators on L^2(G) yield variance-based bounds, with equality for Gaussian-like characters, unifying classical results under group-theoretic harmonic analysis.

References

  1. [1]
    [PDF] harmonic analysis - UCLA Mathematics
    Harmonic analysis1 focuses in particular on the quantitative properties of such functions, and how these quantitative properties change when apply various ( ...
  2. [2]
    [PDF] Lecture 1: Introduction 1 Harmonic Analysis 2 Applications
    Jan 23, 2008 · • Differential Equations: Harmonic analysis began as a method to assist in solving diffusion equations.
  3. [3]
    [PDF] Introduction to Harmonic Analysis - Christopher Heil
    Nov 12, 2010 · theory, and functional analysis. 1.1 Definition and Basic Properties. 1.1.1 The Fourier Transform on L1(R). We define the Fourier transform on ...
  4. [4]
    [PDF] Harmonic analysis - Purdue Math
    Jan 26, 2021 · In this course we study boundary value problems for PDE. The main method is called Harmonic analysis a. k. a. Fourier Analysis. The main.
  5. [5]
    [PDF] Harmonic Analysis and Group Representations
    Harmonic analysis can be interpreted broadly as a general principle that re- lates geometric objects and spectral objects. The two kinds of objects are.
  6. [6]
    Harmonic Analysis -- from Wolfram MathWorld
    Harmonic analysis is the mathematical study of overlapping waves, where integer multiples of a frequency are called harmonics. It is used in signal processing, ...
  7. [7]
    [PDF] Harmonic Analysis in Mathematics
    Harmonic analysis in mathematics links geometric objects with spectral objects, and is a general principle in mathematics.
  8. [8]
    Brief Description of Harmonic Analysis and Representation Theory
    Harmonic analysis, historically using Fourier series, now uses representations of groups, especially non-abelian ones, and is related to quantum mechanics.
  9. [9]
    The Earliest Astronomers: A Brief Overview of Babylonian Astronomy
    Sep 18, 2023 · Babylonian astronomers utilized a sexagesimal system in their calculations and measurements of celestial objects. Sexagesimal numbers are ...
  10. [10]
    Pythagoras - Biography
    ### Summary of Pythagoras' Contributions to Harmonics and Music Theory
  11. [11]
    Mathematical tables in Ptolemy's Almagest - ScienceDirect
    Feb 26, 2014 · As he states in Harmonics III.3, Ptolemy regarded astronomy and harmonics as the highest branches of mathematics, for which geometry and ...
  12. [12]
    Wave equation: d'Alembert's formula - The University of Queensland
    In 1747, Jean le Rond d'Alembert (1717-1783) published a paper on vibrating strings that included his famous solution to the one-dimensional wave equation.
  13. [13]
    Daniel Bernoulli - Biography
    ### Summary of Daniel Bernoulli's Series Solutions for Vibrating Strings
  14. [14]
    Analysis - Trigonometry, Series, Solutions | Britannica
    In 1748, in response to d'Alembert's work, the Swiss mathematician Leonhard Euler wrote a paper, Sur la vibration des cordes (“On the Vibrations of Strings”).
  15. [15]
    Joseph Fourier (1768 - 1830) - Biography - MacTutor
    His life was not without problems however since his theory of heat still provoked controversy. Biot claimed priority over Fourier, a claim which Fourier had ...
  16. [16]
    Henri Lebesgue - Biography
    ### Summary of Lebesgue's Integration Enabling Rigorous Convergence of Fourier Series in the 1900s
  17. [17]
  18. [18]
    Lev Pontryagin - Biography
    ### Summary of Pontryagin's Work on Abstract Harmonic Analysis and Pontryagin Duality (1930s-1940s)
  19. [19]
    [PDF] CHAPTER 4 FOURIER SERIES AND INTEGRALS
    The Gibbs phenomenon is the overshoot that moves closer and closer to the jumps. ... That jump has produced the slow 4/πn decay of the coefficients, with Gibbs ...
  20. [20]
    [PDF] Fourier Series - UMD MATH
    where fk ∈ C are the so-called Fourier coefficients and describe our signal in the “frequency domain”. Example: For the signal f(x) = sin(2πx) we have f(x) = 1.
  21. [21]
    [PDF] 3 Trigonometric Fourier Series - UNCW
    Definition of an orthogonal set of func- tions and orthonormal functions ... We have seen the Gibbs phenomenon when there is a jump discontinu- ity in ...
  22. [22]
    [PDF] MATH 461: Fourier Series and Boundary Value Problems - Chapter III
    Convergence of Fourier Series. Gibbs Phenomenon (cont.) Now, sin 2Nx = 0 if 2Nx = ±π,±2π,...,±2Nπ. The maximum overshoot occurs at x = π. 2N and its value is.
  23. [23]
    [PDF] Fourier Transforms - Mathematical Foundations of Data Sciences
    We recall that for f ∈ L1(R), its Fourier transform is defined as ∀ ω ∈ R,ˆf(ω) def. = ∫R f(x)e−ixωdx. and this is extended to L2(R) by density.Missing: seminal | Show results with:seminal
  24. [24]
    [PDF] EE 261 - The Fourier Transform and its Applications
    1 Fourier Series. 1. 1.1 Introduction and Choices to Make . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. 1.2 Periodic Phenomena .<|control11|><|separator|>
  25. [25]
    [PDF] Chapter 4: Frequency Domain and Fourier Transforms
    The properties tell us how the Fourier transform changes when the original signal undergoes certain changes. They are generally derived by going back to the ...
  26. [26]
    [PDF] Fourier inversion
    Thus we have the Fourier inversion formula at 0, J−1Jϕ(0) = ϕ(0) for Schwartz functions ϕ. As explained above, general Fourier inversion for Schwartz functions ...
  27. [27]
    [PDF] Fourier Transform and Schwartz Functions - UW Math Department
    Fourier inversion formula for f ∈ S(Rn) f(x) = 1. (2π)n. Z eix·ξ f(ξ)dξ. Proof: key fact in proof is an explicit calculation. Z eix·ξe. − |ξ|2. 2 dξ = −n. 2 (2π).Missing: source | Show results with:source
  28. [28]
    [PDF] Plancherel and spectral decomposition/synthesis
    The Plancherel theorem. |Ff| = |f|. (L2(R) norm, for f ∈ S (R)) extends the Fourier transform to L2(R) by continuity, despite the general divergence of the ...
  29. [29]
    [PDF] n-dimensional Fourier Transform
    The geometry of Rn is governed by the dot product, and using it will greatly help our understanding as well as streamline our notation. Page 2. 336. Chapter 8 n ...
  30. [30]
    Fourier Transform--Gaussian -- from Wolfram MathWorld
    The Fourier transform of a Gaussian function f(x)=e^(-ax^2) is given by F_x[e^(-ax^2)](k) = int_(-infty)^inftye^(-ax^2)e^(-2piikx)dx (1) ...
  31. [31]
    Fourier Transform--Delta Function -- from Wolfram MathWorld
    The Fourier transform of the delta function is given by F_x[delta(x-x_0)](k) = int_(-infty)^inftydelta(x-x_0)e^(-2piikx)dx (1) = e^(-2piikx_0). (2)
  32. [32]
    Abstract Harmonic Analysis: Volume I - SpringerLink
    Edwin Hewitt. Department of Mathematics, University of Oregon, Eugene, USA. Kenneth A. Ross. Accessibility Information. PDF accessibility summary. This PDF is ...
  33. [33]
    Pontryagin Duality and the Structure of Locally Compact Abelian ...
    1 - Introduction to topological groups · 2 - Subgroups and quotient groups of Rn · 3 - Uniform spaces and dual groups · 4 - Introduction to the Pontryagin-van ...
  34. [34]
    [PDF] Partial Differential Equations - UC Berkeley math
    Strongly related are Fourier series methods, which represent solutions of certain PDEs on bounded domains as infinite sums entailing sines and cosines. Another ...
  35. [35]
    Green's Function for the Helmholtz Equation
    The Green's function satisfies the homogeneous Helmholtz equation (HHE). Furthermore, clearly the Poisson equation is the $k \to 0$ limit of the Helmholtz ...
  36. [36]
    Fourier transform and Sobolev spaces
    The Fourier transform is a powerful operator; Sobolev spaces control smoothness of functions, and Sobolev norms help prove solutions for partial differential ...Missing: seminal | Show results with:seminal
  37. [37]
    [PDF] A Course in Harmonic Analysis - University of Oregon
    These notes were written to accompany the courses Math 6461 and Math. 6462 (Harmonic Analysis I and II) at Missouri University of Science & Tech- nology.
  38. [38]
    schur orthogonality relations and invariant sesquilinear forms
    Aug 29, 2001 · Abstract. Important connections between the representation theory of a compact group G and L2(G) are summarized by the Schur orthogonality.<|separator|>
  39. [39]
    [PDF] Harmonic Analysis and Group Representations
    Harmonic analysis can be interpreted broadly as a general principle that re- lates geometric objects and spectral objects. The two kinds of objects are.
  40. [40]
    [PDF] A Selective History of the Stone-von Neumann Theorem - UMD MATH
    The Stone-von Neumann Theorem answers these questions af- firmatively in the case of a single free particle, in the sense that all irreducible representations ...
  41. [41]
    [PDF] The Weyl Quantization: a brief introduction
    Nov 4, 2021 · Definition 4 (Fourier transform). Let f ∈ L2(Rn), the Fourier transform of f is. F(f)(ξ) = #Rn e−2πiξxf(x)dx, and the Fourier inverse transform ...
  42. [42]
    [PDF] SU(2) Representations and Their Applications
    In quantum mechanics the states are said to have “quantized” angular momentum, taking on the value l. Such quantum mechanical states are sometimes thought ...
  43. [43]
    on the role of the heisenberg group in harmonic analysis
    Thus simply the observation that the reduced Heisenberg group exists virtually suffices to establish the basic Plancherel and Fourier Inversion formulas.
  44. [44]
    [PDF] A Theory for Multiresolution Signal Decomposition: The Wavelet ...
    We have seen that the wavelet representation is com- plete. We now show that the original discrete signal can also be reconstructed with a pyramid transform.
  45. [45]
    Decomposition of Hardy Functions into Square Integrable Wavelets ...
    Decomposition of Hardy Functions into Square Integrable Wavelets of Constant Shape ... 1 | 1 Oct 1984. View Options. View options. PDF. View PDF. Figures. Tables ...
  46. [46]
    [PDF] Zur Theorie der orthogonalen Funktionensysteme.
    Die Untersuchungen des Kapitels I legen die Frage nahe: Gibt es überhaupt ein orthogonales Funktionensystem, das so be schaffen ist, daß jede st e t ig e F ...
  47. [47]
    [PDF] Orthonormal bases of compactly supported wavelets
    We construct orthonormal bases of compactly supported wavelets, with arbitrarily high regular-.
  48. [48]
    Theory of communication. Part 1: The analysis of information
    01 November 1946. Theory of communication. Part 1: The analysis of information. Author: D. GaborAuthors Info & Affiliations. Publication: Journal of the ...
  49. [49]
    [PDF] Unit 31: Parseval's theorem
    The following theorem is called the Parseval's identity. It is the Pythagoras theorem for Fourier series. Theorem: ||f||2 = a2. 0 +.
  50. [50]
    245C, Notes 2: The Fourier transform | What's new - Terry Tao
    Apr 6, 2009 · There is an analogue of the Plancherel theorem in this setting, closely related to the Peter-Weyl theorem in representation theory. We will not ...
  51. [51]
    [PDF] harmonic analysis on lca groups - UChicago Math
    Aug 28, 2017 · Harmonic analysis attempts to generalize many of these constructions to locally compact abelian groups.
  52. [52]
    [PDF] Lectures on Representation Theory - Princeton Math
    We start with basic definitions and examples, followed by the character theory, the Peter–Weyl theorem for finite groups. Then we introduce the Haar ...
  53. [53]
    [PDF] Sampling in Complex and Harmonic Analysis
    Abstract. This is a survey article on uniqueness, sampling and interpolation problems in complex analysis. Most of these problems are motivated by appli-.
  54. [54]
    [PDF] The uncertainty principle: A mathematical survey
    Uncertainty principles for time-frequency operators. Continuous and. Discrete Fourier Transforms, Extension Problems, and Wiener-Hopf Equations (I. Gohberg ...
  55. [55]
    Uncertainty Principles and Signal Recovery
    The classical uncertainty principle asserts that both a function and its Fourier transform cannot be largely concentrated on intervals of small measure. Donoho ...
  56. [56]
    Symplectic Methods in Harmonic Analysis and in Mathematical ...
    The aim of this book is to give a rigorous and complete treatment of various topics from harmonic analysis with a strong emphasis on symplectic invariance ...