Fact-checked by Grok 2 weeks ago

Eigenfunction

In mathematics, particularly within the field of , an eigenfunction of a linear operator \hat{A} acting on a is defined as a non-zero \psi that satisfies \hat{A} \psi = \lambda \psi, where \lambda is a scalar value known as the corresponding eigenvalue. This concept generalizes the notion of eigenvectors from finite-dimensional vector spaces to infinite-dimensional spaces of functions, allowing the operator to scale the eigenfunction by the eigenvalue without altering its form. The term "eigenfunction" derives from the word Eigenfunktion, meaning "" or "proper function." Eigenfunctions play a central role in solving linear partial differential equations (PDEs), where they provide the fundamental modes or basis functions for expanding general solutions via series representations, such as in Sturm-Liouville theory. In this context, the eigenvalues often correspond to discrete spectral values that satisfy specific boundary conditions, enabling the decomposition of complex problems into simpler components. For instance, in boundary value problems, non-trivial solutions exist only for certain eigenvalues, and the associated eigenfunctions form an under appropriate inner products. Beyond , eigenfunctions are indispensable in physics, especially , where they describe the stationary states of physical systems under self-adjoint operators like the , with eigenvalues representing observable quantities such as energy levels. In , sets of eigenfunctions associated with operators can be simultaneously diagonalized, facilitating the of multivariable systems. Applications extend to and data , including and the study of vibrations in continuous media, where eigenfunctions enable efficient computational methods for wave propagation and .

Definition and Fundamentals

Formal Definition

In , an eigenfunction of a linear operator \hat{T} defined on a is a non-zero function \psi in the domain of \hat{T} that satisfies the equation \hat{T} \psi = \lambda \psi, where \lambda is a scalar value known as the associated eigenvalue. This definition applies within the framework of Hilbert spaces, such as the Lebesgue space L^2 of square-integrable functions over a , where the linearity of \hat{T} ensures that it maps the space to itself while preserving addition and of functions. The operator \hat{T} is commonly notated to encompass forms like or operators, though the abstract linear structure remains central to the concept. The terminology "eigenfunction" was coined by in 1904 during his development of the theory of linear integral equations. This formulation generalizes the finite-dimensional notion of eigenvectors for matrices to infinite-dimensional settings.

Basic Properties

Eigenfunctions of a linear T on a satisfy several fundamental algebraic properties derived directly from the defining equation T\psi = \lambda \psi, where \psi is the eigenfunction and \lambda is the corresponding eigenvalue. These properties highlight the structure of eigenspaces and the behavior under linear operations. One key property is scaling invariance: if T\psi = \lambda \psi, then for any scalar c \neq 0, T(c\psi) = c T\psi = c\lambda \psi = \lambda (c\psi), meaning c\psi is also an eigenfunction with the same eigenvalue \lambda. This follows from the linearity of the operator T. Regarding linear combinations, suppose \psi_1 and \psi_2 are eigenfunctions with eigenvalues \lambda_1 and \lambda_2, respectively. Then, the combination a\psi_1 + b\psi_2 (for scalars a, b) is an eigenfunction only if \lambda_1 = \lambda_2 = \lambda; in that case, T(a\psi_1 + b\psi_2) = \lambda (a\psi_1 + b\psi_2). If \lambda_1 \neq \lambda_2, the combination is not an eigenfunction for either eigenvalue. Moreover, eigenfunctions corresponding to distinct eigenvalues are linearly independent. In the special case of a zero eigenvalue, where T\psi = 0, the eigenfunction \psi belongs to the (or null space) of T, denoted N(T) = \{\phi \mid T\phi = 0\}. This identifies the eigenspace for \lambda = 0 precisely as the of the . Degeneracy arises when the eigenspace for a given eigenvalue \lambda has greater than one, allowing multiple linearly independent eigenfunctions to share the same \lambda. The of this eigenspace, known as the geometric multiplicity of \lambda, quantifies the of degeneracy. For compact operators, eigenspaces corresponding to nonzero eigenvalues are finite-dimensional.

Connection to Linear Algebra

Analogy with Eigenvectors

In linear algebra, an eigenvector of a square matrix A is a non-zero v satisfying A v = \lambda v, where \lambda is the corresponding eigenvalue. This concept generalizes to eigenfunctions in , where functions serve as elements of infinite-dimensional spaces, analogous to vectors in finite dimensions, and linear operators act like infinite-dimensional matrices. A key illustration of this transition arises in discretizing differential ; for instance, the eigenvalue problem for the second operator \frac{d^2 u}{dx^2} = \lambda u on [0,1] with conditions u(0) = u(1) = 0 can be approximated using finite differences on a with spacing h = 1/(m+1), yielding a A whose eigenvalues \mu_p \approx - (p \pi)^2 for small p approximate the continuous eigenvalues \lambda_p = - (p \pi)^2. Unlike finite-dimensional cases, where spectra are always , infinite-dimensional settings permit continuous spectra, as seen in the of certain operators on Hilbert spaces where eigenvalues may form a . For example, the derivative operator on suitable function spaces exemplifies this generalization without eigenvalues.

Extension to Linear Operators

The concept of eigenfunctions extends from finite-dimensional linear algebra to infinite-dimensional settings, particularly on Hilbert spaces like L^2 spaces of functions, where linear operators replace matrices. These operators include differential, integral, and multiplication types, each acting on infinite-dimensional domains and requiring careful definition of the operator's domain to ensure well-posedness. Eigenfunctions f satisfy Tf = \lambda f for eigenvalue \lambda, but unlike matrix eigenvectors, they must lie within the operator's domain, often imposing boundary conditions on functions to close the operator graph. Differential operators, such as the differentiation \frac{d}{dx} on intervals like [0, \pi], demand domains consisting of sufficiently functions satisfying conditions, such as Dirichlet conditions f(0) = f(\pi) = 0, to make the densely defined and closable. For example, the -\frac{d^2}{dx^2} on this yields eigenfunctions that solve the , ensuring the solutions are in L^2 and respect the boundaries. Integral operators, like Fredholm operators defined by (Kf)(x) = \int_a^b k(x,y) f(y) \, dy with continuous k, operate on L^2[a,b] and are typically compact, leading to discrete eigenvalues with corresponding eigenfunctions forming a basis when the is . Multiplication operators, given by (Tf)(x) = a(x) f(x) where a \in L^\infty, act on L^2 s; their is the full since they are bounded. If a(x) is constant, say a(x) = c, then any non-zero f \in L^2 serves as an eigenfunction with eigenvalue c, as Tf = c f. However, if a(x) varies, proper eigenfunctions in L^2 generally do not exist, because the level sets \{x : a(x) = \lambda\} have measure zero for almost all \lambda, though delta-like distributions can be viewed as generalized eigenfunctions in rigged Hilbert s. The of these operators varies with : compact operators, such as Fredholm operators with square-integrable kernels, possess spectra consisting of eigenvalues (possibly accumulating at zero) with finite multiplicity, excluding zero which may be in the continuous . Non-compact operators, like unbounded operators on infinite domains or operators with non-constant a, often exhibit continuous spectra, reflecting the infinite-dimensional without point accumulations. Domain restrictions via boundary conditions are essential for eigenfunctions to belong to the , and extensions may be required for symmetric operators to achieve well-posed eigenvalue problems.

Advanced Properties

Self-Adjoint Operators

In the context of , a T on an V is defined as a linear operator satisfying \langle T f, g \rangle = \langle f, T g \rangle for all f, g \in V, where \langle \cdot, \cdot \rangle denotes the inner product. This property, also known as being Hermitian in the complex case, ensures symmetry with respect to the inner product and is fundamental to . A key result for self-adjoint operators is that all eigenvalues are real numbers. Furthermore, eigenfunctions corresponding to distinct eigenvalues are orthogonal with respect to the inner product. These properties arise directly from the self-adjoint condition: for an eigenvector f with eigenvalue \lambda, the relation \langle T f, f \rangle = \langle f, T f \rangle implies \lambda = \overline{\lambda}, confirming reality; orthogonality follows from (\lambda_1 - \lambda_2) \langle f_1, f_2 \rangle = 0 for distinct \lambda_1, \lambda_2. The provides a deeper : for a bounded on a separable , the spectrum consists of real numbers, and the operator can be represented via a spectral measure. In the case of compact operators, the guarantees an of eigenfunctions spanning the space (except possibly the ), with eigenvalues forming a converging to zero. This allows the operator to be expressed as T v = \sum_n \lambda_n \langle v, e_n \rangle e_n, where \{e_n\} is the eigenbasis. A concrete example is the Hermitian differential operator -\frac{d^2}{dx^2} on the interval [0, L] with Dirichlet boundary conditions f(0) = f(L) = 0. The eigenvalues are \lambda_n = \left( \frac{n \pi}{L} \right)^2 for n = 1, 2, 3, \dots, with corresponding eigenfunctions \sin\left( \frac{n \pi x}{L} \right). In quantum mechanics, self-adjoint operators represent physical observables, ensuring real eigenvalues correspond to measurable outcomes.

Orthogonality and Bases

A fundamental property of eigenfunctions associated with operators is their . For distinct eigenvalues λ_m ≠ λ_n, the corresponding eigenfunctions ψ_m and ψ_n satisfy the inner product relation ⟨ψ_m, ψ_n⟩ = 0, where the inner product is defined on the underlying . These eigenfunctions can be normalized such that ⟨ψ_n, ψ_n⟩ = 1 for each n, forming an orthonormal set. In specific classes of self-adjoint operators, such as those arising from regular Sturm-Liouville problems, the set of eigenfunctions {ψ_n} is complete, meaning it spans the entire (typically L²[a, b] with a suitable ). This completeness allows any function f in the space to be expanded as a series f = ∑ c_n ψ_n, where the coefficients are given by c_n = ⟨f, ψ_n⟩. Such expansions are central to solving boundary value problems and provide a basis analogous to . The completeness property leads to , which preserves the norm in the expansion: ∥f∥² = ∑ |c_n|². This identity quantifies the energy or L²-norm distribution across the eigenfunction components, ensuring the series converges in the L² sense. However, not all operators yield complete discrete bases; those with continuous spectrum require generalized expansions involving integrals over the measure. For instance, the serves as such an expansion for the differentiation operator on L²(ℝ), where the "eigenfunctions" are plane waves forming a continuous in the sense.

Applications in Physics

Classical Wave Problems

In classical wave problems, eigenfunctions arise prominently in the solution of the one-dimensional describing the transverse vibrations of a taut fixed at both ends. The governing is \frac{\partial^2 u}{\partial t^2} = c^2 \frac{\partial^2 u}{\partial x^2}, where u(x, t) represents the at x and time t, and c = \sqrt{T_0 / \rho_0} is the wave speed determined by the 's T_0 and linear \rho_0. Applying , assume u(x, t) = X(x) T(t), which yields the spatial eigenvalue problem X'' + \lambda X = 0 subject to boundary conditions X(0) = X(L) = 0 for a of length L. The eigenvalues are \lambda_n = (n \pi / L)^2 for n = 1, 2, \dots, with corresponding eigenfunctions X_n(x) = \sin(n \pi x / L). These sine functions form an for expanding initial conditions, allowing the general solution to be expressed as a superposition u(x, t) = \sum_{n=1}^\infty \left[ A_n \cos(\omega_n t) + B_n \sin(\omega_n t) \right] \sin(n \pi x / L), where \omega_n = c \sqrt{\lambda_n} = n \pi c / L is the of the nth . Each term represents a , manifesting as a that oscillates independently at harmonic frequencies, with the (n=1) having the lowest frequency. This approach extends to higher dimensions, such as the vibrations of a circular membrane (drumhead), governed by the two-dimensional wave equation \frac{\partial^2 u}{\partial t^2} = c^2 \nabla^2 u in polar coordinates, where \nabla^2 is the Laplacian. Separation of variables u(r, \theta, t) = \phi(r, \theta) T(t) leads to the Helmholtz equation \nabla^2 \phi + \lambda \phi = 0 with boundary condition \phi(a, \theta) = 0 on the radius a. Further separation \phi(r, \theta) = R(r) \Theta(\theta) separates the angular part into \Theta'' + \mu \Theta = 0, yielding eigenvalues \mu_n = n^2 and solutions \Theta_n(\theta) = \cos(n \theta) or \sin(n \theta) for n = 0, 1, 2, \dots. The radial equation becomes r^2 R'' + r R' + (\lambda r^2 - n^2) R = 0, a Bessel equation whose solutions are the Bessel functions of the first kind J_n(\sqrt{\lambda} r). Imposing the boundary condition gives eigenvalues \lambda_{n,m} = (z_{n,m} / a)^2, where z_{n,m} is the mth positive zero of J_n(z) for m = 1, 2, \dots. The eigenfunctions are thus \phi_{n,m}(r, \theta) = J_n(z_{n,m} r / a) \cos(n \theta) or J_n(z_{n,m} r / a) \sin(n \theta), enabling the expansion of the membrane's displacement as a sum over these modes, each evolving harmonically in time. For example, the lowest mode (n=0, m=1) uses J_0, corresponding to a radially symmetric vibration. In three dimensions, such as inside a spherical cavity of radius a, the wave equation \frac{\partial^2 u}{\partial t^2} = c^2 \nabla^2 u separates in spherical coordinates, leading to eigenfunctions composed of Y_l^m(\theta, \phi) for the angular part (with eigenvalues l(l+1)) and radial spherical Bessel functions j_l(kr) satisfying the boundary condition at r = a. The eigenvalues are determined by the zeros of j_l(kr) at r = a, yielding discrete frequencies for modes that describe resonant vibrations within the sphere.

Quantum Mechanics

In quantum mechanics, eigenfunctions of the operator \hat{H} describe stationary states, which are solutions to the time-independent \hat{H} \psi = E \psi, where \psi is the eigenfunction and E is the corresponding eigenvalue. For a single particle in one dimension, the takes the form \hat{H} = -\frac{\hbar^2}{2m} \frac{d^2}{dx^2} + V(x), with V(x) as the . These eigenfunctions form a complete basis for expanding the wave function of any state in the system, and the self-adjoint nature of the guarantees that the eigenvalues are real. A foundational example is the particle in an infinite square well potential, often called the particle in a box, where V(x) = 0 for $0 < x < L and infinite elsewhere. The eigenfunctions are \psi_n(x) = \sqrt{\frac{2}{L}} \sin\left(\frac{n \pi x}{L}\right) for n = 1, 2, 3, \dots, with corresponding energies E_n = \frac{n^2 \pi^2 \hbar^2}{2 m L^2}. This model illustrates energy quantization, where only discrete values of E_n are allowed, reflecting the confinement of the particle and leading to standing-wave-like probability distributions that vanish at the boundaries. Another key example is the , modeling systems like molecular vibrations, with potential V(x) = \frac{1}{2} m \omega^2 x^2. The eigenfunctions are \psi_n(\xi) = N_n H_n(\xi) e^{-\xi^2 / 2}, where \xi = \sqrt{m \omega / \hbar} \, x, H_n are , and N_n is a normalization constant; the energies are E_n = \hbar \omega \left(n + \frac{1}{2}\right) for n = 0, 1, 2, \dots, exhibiting equidistant spacing above the \frac{1}{2} \hbar \omega. This structure arises from the quadratic potential and underpins phenomena such as equally spaced vibrational spectra in . Physically, the modulus squared |\psi|^2 of an eigenfunction provides the probability density for locating the particle at a given position, as established by the . Stationary states evolve only by a e^{-i E t / \hbar}, maintaining constant probability densities over time, whereas general time-dependent wave functions are linear superpositions of these eigenfunctions, allowing for dynamic interference and evolution according to the full time-dependent .

Applications in Engineering and Analysis

Signal Processing

In signal processing, eigenfunctions play a central role in the decomposition and analysis of signals, particularly through transforms that exploit the spectral properties of linear operators. A foundational example is Fourier analysis, where sine and cosine functions serve as eigenfunctions of the second-derivative operator \frac{d^2}{dx^2} on periodic domains. For a periodic function on [0, 2\pi], the eigenfunctions are \sin(nx) and \cos(nx) for integers n, with corresponding eigenvalues -n^2, allowing any square-integrable periodic signal to be decomposed into a sum of these frequency components via the Fourier series. This decomposition enables efficient frequency-domain representation and filtering of signals, such as isolating specific harmonic content in audio or vibration data. For linear time-invariant (LTI) systems, complex exponentials e^{i \omega t} act as eigenfunctions of the convolution operator, which characterizes the system's response. When such an exponential input is applied, the output is the same exponential scaled by the system's H(i\omega), the eigenvalue, simplifying the analysis of system behavior across frequencies. This property underpins the frequency-domain approach in , where arbitrary signals are expressed as superpositions of these exponentials, and the system's output is obtained by multiplying each component by H(i\omega). In , a branch of dealing with continuous curves or functions, () extends to the Karhunen-Loève (KL) expansion, which uses eigenfunctions of the to decompose random signals. For a X(t) with function K(s,t) = \mathbb{E}[(X(s) - \mu(s))(X(t) - \mu(t))], the KL theorem provides an orthogonal expansion X(t) = \mu(t) + \sum_k \sqrt{\lambda_k} \xi_k \phi_k(t), where \{\phi_k\} are the eigenfunctions and \{\lambda_k\} the eigenvalues of the defined by K, solving \int K(s,t) \phi_k(s) \, ds = \lambda_k \phi_k(t). This expansion, analogous to for vector data, captures the principal modes of variation in functional signals like or spectra, prioritizing those with largest variances for . A key application of these eigenfunction decompositions is , where retaining only the dominant eigenmodes—those with the largest eigenvalues—filters out low-variance noise components while preserving signal structure. In multichannel , the KL transform applied to the of noisy observations identifies signal-dominated modes for reconstruction, achieving significant improvements, as demonstrated in seismic data suppression of random noise in empirical studies. Similarly, in processing, this approach yields enhanced speech signals by suppressing uncorrelated noise across channels.

Separation of Variables

The method of is a fundamental technique for solving linear partial differential equations (PDEs) by assuming a product of the form u(\mathbf{x}, t) = X(\mathbf{x}) T(t), where \mathbf{x} represents spatial variables and t is time, leading to an (ODE) in space that reduces to an eigenvalue problem L X = \lambda X for a spatial L. Substituting this into the PDE separates the variables, yielding T'/T = \lambda / \kappa (for diffusion-like equations) and the spatial eigenvalue problem, whose eigenfunctions form the basis for expanding the general as a superposition \sum c_n X_n(\mathbf{x}) e^{-\lambda_n t / \kappa}. This approach exploits the of the PDE and the self-adjointness of L to ensure orthogonal eigenfunctions, enabling efficient computation of coefficients via inner products. A canonical example is the one-dimensional \partial u / \partial t = \kappa \partial^2 u / \partial x^2 on [0, L] with Dirichlet boundary conditions u(0, t) = u(L, t) = 0. Assuming u(x, t) = X(x) T(t) yields the spatial Sturm-Liouville problem X'' + \lambda X = 0 with X(0) = X(L) = 0, having eigenvalues \lambda_n = (n \pi / L)^2 and eigenfunctions X_n(x) = \sin(n \pi x / L) for n = 1, 2, \dots. The time component satisfies T' + \kappa \lambda_n T = 0, giving T_n(t) = e^{-\kappa (n \pi / L)^2 t}, so the general solution is u(x, t) = \sum_{n=1}^\infty b_n \sin(n \pi x / L) e^{-\kappa (n \pi / L)^2 t}, where coefficients b_n are determined from initial conditions via the sine series. For the Helmholtz equation \nabla^2 u + k^2 u = 0 in polar coordinates (r, \theta) within a disk of radius a, separation of variables assumes u(r, \theta) = R(r) \Theta(\theta), leading to the angular eigenvalue problem \Theta'' + m^2 \Theta = 0 with periodic boundary conditions, yielding eigenvalues m^2 (non-negative integers m) and eigenfunctions \Theta_m(\theta) = \cos(m \theta) or \sin(m \theta). The radial part then satisfies the Bessel equation r^2 R'' + r R' + (k^2 r^2 - m^2) R = 0, with solutions involving Bessel functions J_m(k r) and Y_m(k r); for boundedness at the origin, Y_m is discarded. For boundary value problems like Dirichlet conditions on the disk, the eigenvalues k_{mn} are roots of J_m(k a) = 0. The full eigenfunctions are products J_m(k_{mn} r) \{ \cos(m \theta), \sin(m \theta) \}, forming a basis for expanding solutions inside the domain. Eigenfunction expansions from converge to the solution in the L^2 sense due to the completeness of the eigenfunctions for operators on compact domains, meaning the series \sum c_n X_n approximates any L^2 function with error tending to zero in the L^2 norm. holds under additional assumptions on the solution and boundary data, such as for analytic initial conditions in the , ensuring throughout the domain. This completeness theorem underpins the method's reliability for boundary value problems.

References

  1. [1]
    [PDF] 2. Introduction to Quantum Mechanics - MIT OpenCourseWare
    P: Eigenvalues and eigenfunctions of an operator are defined as the solutions of the eigenvalue problem: A[un(rx)] = anun(rx) where n = 1, 2,... indexes the ...
  2. [2]
    [PDF] QM notes
    An operator operates on a function. Eigenvalues and eigenfunctions satisfy A!j = aj!j. Linear operators have properties like O[f(x) + g(x)] = Of(x) + Og(x).
  3. [3]
    Eigenvalues and Eigenfunctions - HyperPhysics
    "Eigenfunction" is from "Eigenfunktion" meaning "proper or characteristic function". Index. Schrodinger equation concepts.Missing: definition | Show results with:definition
  4. [4]
    Differential Equations - Eigenvalues and Eigenfunctions
    Nov 16, 2022 · An eigenvalue (λ) for a BVP is a value that gives non-trivial solutions, and the corresponding non-trivial solutions are called eigenfunctions.
  5. [5]
    [PDF] Eigenfunction Expansions
    Eigenfunction expansions use eigenfunctions to represent arbitrary functions as infinite series, where constants characterize the function.
  6. [6]
    D.18 Eigenfunctions of commuting operators
    Commuting operators A and B share a common set of eigenfunctions if each has a complete set of eigenfunctions. This extends to more operators.
  7. [7]
    [PDF] MAT 280: Laplacian Eigenfunctions: Theory, Applications, and ...
    Mar 29, 2007 · Also, the eigenfunctions can be used for spectral analysis of data defined (or living) on Ω. Furthermore, Laplacian eigenfunctions allow us to.
  8. [8]
    Eigenfunction -- from Wolfram MathWorld
    If L^~ is a linear operator on a function space, then f is an eigenfunction for L^~ and lambda is the associated eigenvalue whenever L^~f=lambdaf.Missing: definition | Show results with:definition
  9. [9]
    [PDF] Eigenvalues and Eigenfunctions of the Laplacian
    In this article we will consider the eigenvalue problem for the Laplace operator acting on the L2 space of functions on a bounded domain in Rn. We prove that ...
  10. [10]
    Earliest Known Uses of Some of the Words of Mathematics (E)
    The eigen terms are associated with David Hilbert (1862-1943), though he may have been following such constructions as Eigentöne in acoustics (cf. H. L. F. ...
  11. [11]
    Math Origins: Eigenvectors and Eigenvalues
    The eigen- prefix appears in both papers. In the former, Hilbert coined the word Eigenzeit (in English, "eigentime") for eigenvalues of a particular matrix, ...
  12. [12]
    None
    Below is a merged summary of the eigenfunctions for linear operators based on the provided segments from Kreyszig's *Introductory Functional Analysis with Applications*. To retain all information in a dense and organized manner, I will use a table in CSV format to summarize the key properties across the segments, followed by a narrative summary that consolidates additional details and references. This approach ensures all details (e.g., page numbers, theorems, and URLs) are preserved and easily accessible.
  13. [13]
    [PDF] FUNCTIONAL ANALYSIS | Second Edition Walter Rudin
    Functional analysis/Walter Rudin.-2nd ed. p. em. -(international series in pure and applied mathematics). Includes bibliographical references (p. ) ...Missing: online | Show results with:online
  14. [14]
    [PDF] Finite Difference Methods for Ordinary and Partial Differential ...
    Jun 1, 2007 · The discrete approximation to this operator given by the matrix A has only m eigenvalues instead of an infinite number, and the ...
  15. [15]
    [PDF] Chapter 9: The Spectrum of Bounded Linear Operators
    The action of the operator on the subspace of eigen- vectors with a given eigenvalue is just multiplication by the eigenvalue. As we will see, the spectral ...
  16. [16]
    [PDF] 9.2. Eigenvalues and eigenfunctions of differential operators
    We will focus on the case P(z)=z², i-e. P(D)=D². 2. Goal: find the eigenvalues & eigenfunctions of the operator ...
  17. [17]
    [PDF] III. Linear Operators on Hilbert Spaces
    A is called a multiplication operator. Definition. An operator on a vector ... All eigenvalues of a self-adjoint operator on a Hilbert space are real.
  18. [18]
    [PDF] Continuous operators on Hilbert spaces
    Apr 28, 2014 · [1] A linear operator T : X → Y on. Hilbert spaces is compact when it maps the unit ball in X to a pre-compact set in Y . Equivalently, T is.
  19. [19]
    [PDF] 5: Inner Products, Adjoints, Spectral Theorems, Self-Adjoint Operators
    Self-Adjoint Operators. Definition 7.1 (Self-Adjoint Operator). Let F denote R or C. Let V be a finite- dimensional inner product space over F. Let T : V ...
  20. [20]
    [PDF] C Self-adjoint operators and complete orthonormal bases
    The main fact we will present here is that self-adjoint operators typically have orthogonal eigenvectors, and under some conditions, these eigenvectors can form ...
  21. [21]
    [PDF] Spectral theory in Hilbert spaces (ETH Zürich, FS 09) E. Kowalski
    We can now prove Theorem 3.1 for self-adjoint operators. Theorem 3.17 (Spectral theorem for self-adjoint operators). Let H be a separable. Hilbert space and ...
  22. [22]
    [PDF] Chapter 6 Sturm-Liouville Eigenvalue Problems - UNCW
    In physics many problems arise in the form of boundary value prob- lems involving second order ordinary differential equations. For example,.
  23. [23]
    Self-adjointness in Quantum Mechanics: a pedagogical path - arXiv
    Dec 28, 2020 · Observables in quantum mechanics are represented by self-adjoint operators on Hilbert space. Such ubiquitous, well-known, and very foundational ...
  24. [24]
    1.18 Linear Second Order Differential Operators and Eigenfunction ...
    Mar 27, 2024 · Note that eigenfunctions for distinct (necessarily real) eigenvalues of a self-adjoint operator are mutually orthogonal. If an eigenvalue is ...
  25. [25]
    [PDF] Proof of the spectral theorem
    Nov 5, 2013 · The theorem says first of all that a selfadjoint operator is diagonalizable, and that all the eigenvalues are real. The orthogonality of the ...
  26. [26]
    [PDF] Math 531 - Partial Differential Equations - Vibrating String
    As before, we apply our separation of variables technique: u(x, t) = φ(x)h(t), so φ00h = c2φh00 or h00 c2h = φ00 φ = −λ.Missing: classical | Show results with:classical
  27. [27]
    [PDF] MATH 461: Fourier Series and Boundary Value Problems - Chapter VII
    We will now look at two examples and see how to apply separation of variables in these different cases: vibrations of an arbitrarily shaped membrane, i.e., a 2D ...
  28. [28]
    None
    ### Summary of Vibrations of a Circular Membrane
  29. [29]
    Strings, drums, and the Laplacian
    Schrödinger equation: i. ∂U(t, x). ∂t. = −ΔU(t, x), where i2 = −1. In quantum mechanics, the solution U(t, x) of this equation is called the wave function.Missing: 2D | Show results with:2D<|control11|><|separator|>
  30. [30]
    [PDF] 3. Quantisation as an eigenvalue problem; by E. Schrödinger
    To the reduction of the quantisation of degenerate systems might correspond the arbitrary distribution of the energy over the eigenfunctions belonging to one ...
  31. [31]
    [PDF] Zur Quantenmechanik der Sto&#x00DF;vorg&#x00E4;nge - psiquadrat
    (Eingegangen am 25. Juni 1926.) Durch eine Untersuchung der S~oflvorg~nge wird die Auffassung entwickelt, daft die Quantenmechanik in der S chrSdingerschen Form ...
  32. [32]
    [PDF] Low-Pass Filters, Fourier Series and Partial Differential Equations
    Since the elements of the Fourier basis are also eigenfunctions of the second-derivative operator, as one can easily see by simply calculating the derivatives,.
  33. [33]
    [PDF] Fourier Analysis
    The Fourier series is used to figure out which sine and cosine waves should be summed, at what amplitude, to create a periodic waveform of interest. This is ...
  34. [34]
    [PDF] SIGNALS, SYSTEMS, and INFERENCE — Class Notes for 6.011
    2.2.2 Eigenfunction and Transform Representation of LTI Systems. Exponentials are eigenfunctions of LTI mappings, i.e., when the input is an expo nential for ...
  35. [35]
    [PDF] Lecture Notes EE301 Signals and Systems I
    Dec 2, 2016 · For LTI systems, complex exponential signals are eigenfunctions : For LTI systems, representation of signals as a linear combination of complex ...<|separator|>
  36. [36]
    [PDF] The Karhunen-Loève Theorem - CORE
    The Karhunen-Loève Expansion is to be seen as the stochastic parallel of one of the most famous analytical methods ever: the Fourier. Expansion. In the Fourier ...
  37. [37]
    [PDF] Review of Functional Data Analysis - arXiv
    Jul 18, 2015 · This paper provides an overview of FDA, starting with simple statistical notions such as mean and covariance functions, then covering some core ...
  38. [38]
    [PDF] SIGNAL-TO-NOISE RATIO ENHANCEMENT IN MULTICHANNEL ...
    The Karhunen-Loeve transform is useful for data compression for the transmission and storage of stacked seismic data. The second problem is the suppression of ...
  39. [39]
    Multichannel Noise Reduction in the Karhunen-Loève Expansion ...
    Typically the noise reduction problem is approached by passing the noisy microphone signals through a linear filter in order to obtain a cleaner version of the ...Missing: eigenmodes | Show results with:eigenmodes
  40. [40]
    [PDF] Notes on Separation of Variables
    Sep 25, 2012 · 1 Overview. Separation of variables is a technique to reduce the dimensionality of PDEs by writing their solutions.
  41. [41]
    Differential Equations - Separation of Variables
    Nov 16, 2022 · The method of separation of variables relies upon the assumption that a function of the form, u(x,t)=φ(x)G(t) (1) will be a solution to a linear homogeneous ...
  42. [42]
    [PDF] 2 Heat Equation
    The general solution of the ODE is given by X(x) = C + Dx. The boundary condition X(−l) = X(l) =⇒ D = 0. X0(−l) = X0(l) is automatically satisfied if D = 0. ...
  43. [43]
    [PDF] The 1-D Heat Equation
    Sep 8, 2006 · This is the Fourier Sine Series of f (x). To solve for the Bn's, we use the orthogonality property for the eigenfunctions sin (nπx),. 1. 0.<|separator|>
  44. [44]
    [PDF] MAT 280: Laplacian Eigenfunctions: Theory, Applications, and ...
    Recall in polar coordinates that (x, y) = (r cos θ, r sinθ) for 0 ≤ r < a ... This is Bessel's differential equation of order n. For more about Bessel ...
  45. [45]
    [PDF] The Laplace operator in polar coordinates
    This function is known as the Bessel function of order n, usually written Jn(s). Eigenvalues and Eigenfunctions of the Laplacian. Following the reasoning above ...Missing: angular | Show results with:angular
  46. [46]
    [PDF] SPECTRUM OF BOUNDED DOMAINS 1. Eigenfunction expansions ...
    ∆u + λu = 0 in Da, u = 0 on ∂Da. The Laplacian in polar coordinates (r, θ) is: ∆u = urr + 1 r ur + 1 r2 uθθ. Since we know the eigenvalues of the operator L[u] ...
  47. [47]
    [PDF] math 124b: introduction to pdes and fourier series
    Sep 5, 2016 · Convergence and completeness. Definition 1.4. We say that an infinite series Pfn(x) converges to f(x) pointwise in (a, b) if it.
  48. [48]
    [PDF] Applications of Partial Differential Equations To Problems in Geometry
    Step 5 L2 Completeness of the Eigenfunctions. Let PN be the (self-adjoint) ... Step 6 L2 Uniform Convergence of the Eigenfunction Expansion. Page 44. 3.3 ...