Fact-checked by Grok 2 weeks ago

Operator theory

Operator theory is a branch of that studies linear s on normed spaces, particularly bounded and unbounded s acting between Banach or Hilbert spaces. It focuses on their properties, such as , spectra, s, and functional calculi, forming a foundational framework for understanding infinite-dimensional linear mappings. Central concepts in operator theory include the of an , defined as the set of complex numbers \lambda for which T - \lambda I is not invertible, which determines the operator's behavior and eigenvalues. The T^* satisfies \langle T\xi, \eta \rangle = \langle \xi, T^*\eta \rangle and preserves the norm, enabling the study of self-adjoint, (T^*T = TT^*), unitary, and positive s. , a core component, extends finite-dimensional to infinite dimensions via tools like the Riesz functional calculus for s and the Hille-Yoshida for semigroups generated by unbounded s. The field originated in the late 19th century with work on integral equations by and Ivar Fredholm, evolving through Stefan Banach's 1932 textbook on linear operators in Banach spaces and John von Neumann's 1930s extensions of the to Hilbert spaces. Operator theory underpins applications in , where operators represent observables; partial differential equations, modeling evolution via operator semigroups; and , signal processing, and electromagnetics for solving boundary value problems and analyzing systems.

Fundamentals of Operators

Definitions and Classifications

In operator theory, a linear operator is defined as a linear map T: D(T) \to Y between normed vector spaces X and Y, where D(T) \subseteq X is a linear subspace called the domain of T, and the map satisfies T(\alpha x + \beta z) = \alpha T x + \beta T z for all scalars \alpha, \beta and x, z \in D(T). The primary settings for such operators are Banach spaces, which are complete normed vector spaces, and Hilbert spaces, which are Banach spaces equipped with an inner product inducing the norm. In these spaces, operators often act on function spaces like L^2 spaces, where linearity preserves addition and scalar multiplication of functions. Linear operators are classified based on and other structural properties. A is continuous and satisfies \|T\| = \sup_{\|x\| \leq 1} \|T x\| < \infty, where \|T\| is the operator norm measuring the maximum amplification of the norm. Unbounded operators, in contrast, lack a finite operator norm and are defined only on a dense subspace of the domain to ensure well-posedness. Further classifications include finite-rank operators, which have finite-dimensional range and are thus compact; compact operators, which map bounded sets to precompact sets (relatively compact closures); and trace-class operators, a subclass of compact operators where the sum of singular values is finite. A representative example is the integral operator (T f)(x) = \int_a^b k(x,y) f(y) \, dy on L^2[a,b], which is compact if the kernel k \in L^2([a,b] \times [a,b]) and trace-class under stronger conditions like k being Hilbert-Schmidt (i.e., \int \int |k(x,y)|^2 \, dx \, dy < \infty). Basic properties of linear operators include the range R(T) = \{ T x : x \in D(T) \}, a subspace of Y, and the kernel \ker(T) = \{ x \in D(T) : T x = 0 \}, a subspace of D(T) whose dimension relates to the operator's nullity. For operators on Hilbert spaces H, the adjoint T^* is formally defined by the relation \langle T x, y \rangle = \langle x, T^* y \rangle for all x \in D(T) and y \in D(T^*), where D(T^*) is the set of y \in H for which the functional x \mapsto \langle T x, y \rangle is bounded, ensuring T^* is densely defined.

Bounded Operators

Bounded linear operators between normed linear spaces play a central role in functional analysis due to their uniform continuity and preservation of bounded sets in a controlled manner. Specifically, a linear operator T: X \to Y is bounded if there exists a constant M \geq 0 such that \|Tx\| \leq M \|x\| for all x \in X, with the operator norm defined as \|T\| = \inf \{ M \geq 0 : \|Tx\| \leq M \|x\| \ \forall x \in X \}. This norm satisfies \|Tx\| \leq \|T\| \|x\| for all x \in X, ensuring that bounded operators map bounded sets to bounded sets with the image diameter at most \|T\| times the original. The space of bounded linear operators \mathcal{B}(X, Y) forms a Banach space when Y is complete, with the operator norm topology. Composition of bounded operators preserves boundedness: if S: Y \to Z and T: X \to Y are bounded, then ST: X \to Z is bounded with \|ST\| \leq \|S\| \|T\|. Equality holds if, for example, S or T attains its norm on the range of the other, such as when one is an . Bounded operators are precisely the continuous linear operators, as the inequality implies uniform continuity. A key example of bounded operators arises in L^p spaces, where the multiplication operator M_f: L^p(\mu) \to L^p(\mu) defined by (M_f g)(x) = f(x) g(x) is bounded if and only if f is essentially bounded, with \|M_f\| = \|f\|_\infty = \esssup |f|. This operator preserves the L^p norm up to the essential supremum of |f|, illustrating how bounded multipliers control function growth. Another class involves approximations to differential operators; for instance, operators with compact resolvent, such as the Dirichlet Laplacian on a bounded domain, admit bounded inverses (resolvents) that approximate the operator on finite-dimensional subspaces, facilitating numerical analysis. Several fundamental theorems characterize bounded operators on Banach spaces. The uniform boundedness principle, or Banach-Steinhaus theorem, states that a family \{T_\alpha: X \to Y\} of bounded operators is uniformly bounded (i.e., \sup_\alpha \|T_\alpha\| < \infty) if it is pointwise bounded (i.e., \sup_\alpha \|T_\alpha x\| < \infty for each x \in X). This prevents pathological pointwise behavior in infinite dimensions. The open mapping theorem asserts that a surjective bounded linear operator T: X \to Y between Banach spaces is open, meaning it maps open sets to open sets; equivalently, there exists c > 0 such that B_Y(0, c) \subseteq T(B_X(0, 1)), where B denotes the open unit ball. The closed graph theorem provides a criterion for boundedness: a linear operator T: X \to Y between Banach spaces, defined on all of X, is bounded if and only if its graph \Gamma(T) = \{ (x, Tx) \in X \times Y : x \in X \} is closed in the product topology. This equivalence simplifies verification of continuity by checking sequential closure rather than direct norm estimates. Adjoint operators of bounded operators, defined via duality, inherit boundedness with \|T^*\| = \|T\|.

Unbounded Operators

In operator theory, unbounded operators arise naturally when extending the framework of bounded operators to more general settings, such as differential operators in partial differential equations and . An T on a H is a defined on a dense D(T) \subset H such that there exists no constant M < \infty satisfying \|Tx\| \leq M \|x\| for all x \in D(T). The domain D(T) is chosen maximally as the set of all x \in H for which Tx can be defined in a way that ensures the operator is (densely defined and) continuous with respect to the norm topology on D(T). This contrasts with bounded operators, which are uniformly continuous and extend uniquely to the entire space. A key tool for analyzing unbounded operators is the graph of T, defined as the set G(T) = \{ (x, Tx) \mid x \in D(T) \} \subset H \oplus H, where H \oplus H carries the product Hilbert space structure. The operator T is closed if G(T) is a closed subspace of H \oplus H; equivalently, if x_n \in D(T) with x_n \to x and Tx_n \to y in H, then x \in D(T) and Tx = y. Closedness is characterized using the graph norm on D(T), given by \|x\|_G = \sqrt{\|x\|^2 + \|Tx\|^2}, which makes D(T) a pre-Hilbert space; T is closed if and only if D(T) is complete with respect to this norm. Many unbounded operators encountered in applications are not initially closed but admit extensions to closed operators. An operator T is closable if the closure \overline{G(T)} of its graph is itself the graph of a linear operator \overline{T}, called the closure of T, which is the minimal closed extension. For symmetric operators—those satisfying \langle Tx, y \rangle = \langle x, Ty \rangle for all x, y \in D(T)—special self-adjoint extensions exist under certain conditions. The Friedrichs extension applies to positive symmetric operators (where \langle Tx, x \rangle \geq 0 for x \in D(T)) and is obtained by completing the quadratic form domain and extending via the associated sesquilinear form, yielding a self-adjoint operator with the smallest form domain among all positive self-adjoint extensions. Dually, the Krein extension (also known as the Krein-von Neumann extension) is the maximal positive self-adjoint extension, characterized by having the largest possible domain among positive self-adjoint extensions. To determine the existence of self-adjoint extensions for symmetric operators, one computes the deficiency indices d_\pm(T) = \dim \ker(T^* \mp i I), where T^* is the adjoint operator. Self-adjoint extensions exist if and only if d_+(T) = d_-(T); in this case, the extensions are parameterized by unitary maps from \ker(T^* - i I) to \ker(T^* + i I). The Friedrichs and Krein extensions correspond to specific choices in this parameterization for positive operators, with equal deficiency indices ensuring their availability. Prominent examples illustrate the role of unbounded operators. The momentum operator on L^2(\mathbb{R}) is defined as P = -i \frac{d}{dx} with initial domain D(P) = C_c^\infty(\mathbb{R}), the space of smooth functions with compact support; this minimal operator is closable, symmetric, and essentially self-adjoint (deficiency indices (0,0)), admitting a unique self-adjoint extension to the maximal domain where the distributional derivative lies in L^2(\mathbb{R}). Similarly, the Laplacian operator \Delta on L^2(\Omega) for a domain \Omega \subset \mathbb{R}^n is unbounded, typically defined on Sobolev spaces such as D(\Delta) = H^2(\Omega) \cap H_0^1(\Omega) for the Dirichlet boundary conditions, where it is self-adjoint and positive; different boundary conditions yield distinct self-adjoint extensions via the deficiency index method. The modern theory of unbounded operators was developed in the 1930s by and , primarily to provide a rigorous mathematical foundation for quantum observables like position and momentum, which are inherently unbounded. 's introduction of the graph method in 1932 enabled the systematic study of domains and extensions, resolving foundational issues in quantum mechanics.

Spectral Theory for Single Operators

Spectrum and Resolvent Set

In operator theory, for a bounded linear operator T on a complex Banach space X, the resolvent set \rho(T) consists of all complex numbers \lambda \in \mathbb{C} such that \lambda I - T: X \to X is bijective and its inverse (\lambda I - T)^{-1} is a bounded operator on X. The resolvent operator is defined as R(\lambda, T) = (\lambda I - T)^{-1} for \lambda \in \rho(T), and it satisfies the resolvent identity R(\lambda, T) - R(\mu, T) = (\mu - \lambda) R(\lambda, T) R(\mu, T) for distinct \lambda, \mu \in \rho(T). This operator R(\cdot, T) is holomorphic (analytic) on \rho(T), as it can be expressed locally via the Neumann series expansion R(\lambda, T) = \sum_{n=0}^\infty \lambda^{-(n+1)} T^n for |\lambda| > \|T\|, and by the resolvent identity elsewhere in connected components. The \sigma(T) is the complement \mathbb{C} \setminus \rho(T), which is a nonempty compact subset of \mathbb{C} for bounded T on infinite-dimensional X. It partitions into three disjoint sets: the point spectrum \sigma_p(T) = \{\lambda \in \mathbb{C} : \lambda I - T \text{ is not injective}\}, consisting of eigenvalues; the continuous spectrum \sigma_c(T) = \{\lambda \in \mathbb{C} : \lambda I - T \text{ is injective, has dense range, but is not surjective}\}; and the spectrum \sigma_r(T) = \{\lambda \in \mathbb{C} : \lambda I - T \text{ is injective but has non-dense range}\}. For operators, the spectrum is empty, ensuring the spectrum lies on the approximate point spectrum. The spectral radius r(T) = \sup\{|\lambda| : \lambda \in \sigma(T)\} equals \lim_{n \to \infty} \|T^n\|^{1/n}, known as Gelfand's formula. To derive this, note first that the operator norms satisfy \|T^n\| \leq \|T\|^n, so \limsup_{n \to \infty} \|T^n\|^{1/n} \leq \|T\| by submultiplicativity \|T^{m+n}\| \leq \|T^m\| \|T^n\|. For the reverse, if |\lambda| > r(T), then \lambda \in \rho(T) and \|R(\lambda, T)\| \leq 1/(|\lambda| - r(T)) by the Neumann series convergence radius. The lower bound \liminf_{n \to \infty} \|T^n\|^{1/n} \geq r(T) follows by contradiction: if \liminf \|T^n\|^{1/n} = l < r(T), then for some \lambda with l < |\lambda| < r(T), the Neumann series for R(\lambda, T) would converge, implying \lambda \in \rho(T), which contradicts \sigma(T) being compact and containing points up to radius r(T). Thus, equality holds. A classic example is the unilateral (right) shift operator S on \ell^2(\mathbb{N}), defined by S(e_k) = e_{k+1} for the standard basis \{e_k\}, where \|S\| = 1. Its spectrum is the closed unit disk \sigma(S) = \{\lambda \in \mathbb{C} : |\lambda| \leq 1\}, with empty point spectrum, residual spectrum the open unit disk \{\lambda : |\lambda| < 1\}, and continuous spectrum the unit circle \{\lambda : |\lambda| = 1\}. For compact operators on infinite-dimensional Hilbert spaces, the spectrum consists of 0 and at most countably many nonzero eigenvalues of finite multiplicity, with 0 as the only possible accumulation point.

Normal Operators

A bounded linear operator T on a complex Hilbert space \mathcal{H} is defined to be normal if it commutes with its adjoint operator T^*, that is, if T T^* = T^* T. This defining property ensures that normal operators preserve the geometry of the space in a balanced way relative to their adjoints. An immediate consequence is that \| T x \| = \| T^* x \| for every x \in \mathcal{H}, which follows from the equality \langle T x, T x \rangle = \langle x, T^* T x \rangle = \langle x, T T^* x \rangle = \langle T^* x, T^* x \rangle. The Fuglede-Putnam theorem provides a key commutativity result for normal operators: if N is normal and S is bounded with S N = N S, then S N^* = N^* S. This theorem implies that if two normal operators commute, they share common eigenspaces and can be simultaneously diagonalized with respect to an orthonormal basis of \mathcal{H}. Additionally, the spectrum of a normal operator exhibits unitary invariance: for any unitary operator U on \mathcal{H}, \sigma(U T U^*) = \sigma(T). The spectrum itself, computed via the resolvent set, often displays circular symmetry for normal operators. Prominent examples of normal operators include unitary operators, for which U U^* = I = U^* U, and orthogonal projections P, satisfying P = P^* and thus P P^* = P^2 = P = P^* P. Diagonal operators in an orthonormal basis, represented by diagonal matrices with complex entries, are also normal, as their adjoints are the entrywise conjugates, which commute with the original. A fundamental norm estimate for polynomials applied to normal operators is \| p(T) \| = \sup_{\lambda \in \sigma(T)} |p(\lambda)|, where p is a polynomial; this reflects the operator's spectral behavior directly. The development of the spectral theorem for normal operators generalized David Hilbert's foundational 1906 work on self-adjoint operators arising from integral equations. This extension broadened the applicability of spectral decomposition to operators with complex spectra while maintaining key analytic properties.

Self-Adjoint Operators

In a complex Hilbert space \mathcal{H}, a linear operator T with dense domain \mathcal{D}(T) \subset \mathcal{H} is called symmetric if \langle Tx, y \rangle = \langle x, Ty \rangle for all x, y \in \mathcal{D}(T), where \langle \cdot, \cdot \rangle denotes the inner product. The adjoint T^* is defined on the domain \mathcal{D}(T^*) = \{ y \in \mathcal{H} \mid \exists z \in \mathcal{H} \text{ s.t. } \langle Tx, y \rangle = \langle x, z \rangle \ \forall x \in \mathcal{D}(T) \} by T^* y = z. A densely defined operator T is self-adjoint if T = T^*, meaning \mathcal{D}(T) = \mathcal{D}(T^*) and Tx = T^* x for all x \in \mathcal{D}(T). For bounded self-adjoint operators, the domain is the entire space \mathcal{H}, whereas unbounded cases require careful specification of \mathcal{D}(T) to ensure the adjoint is well-defined and the equality holds. Self-adjoint operators form a subclass of normal operators, as they commute with their adjoints. A fundamental property of self-adjoint operators is that their spectrum lies on the real line: \sigma(T) \subset \mathbb{R}. This ensures that eigenvalues, when they exist, are real, which is crucial for physical interpretations. A self-adjoint operator T is positive if \sigma(T) \subset [0, \infty) and \langle Tx, x \rangle \geq 0 for all x \in \mathcal{D}(T); more generally, lower semi-bounded operators (with spectrum bounded below) admit self-adjoint extensions. The Cayley transform provides a bijection between self-adjoint operators and certain unitary operators. For a self-adjoint T, the operator U = (T - iI)(T + iI)^{-1} is unitary on \mathcal{H}, with $1 \notin \sigma(U), since i \notin \sigma(T) and the ranges of T \pm iI are dense (and equal to \mathcal{H} for closed T). Conversely, given a unitary U with $1 \notin \sigma(U), the inverse transform T = i(I + U)(I - U)^{-1} yields a self-adjoint operator. This transform, introduced by von Neumann, facilitates the study of self-adjointness by reducing it to unitarity. In quantum mechanics, self-adjoint operators model physical observables, ensuring real-valued measurement outcomes. The position operator \hat{Q} on L^2(\mathbb{R}) is defined by (\hat{Q} \psi)(x) = x \psi(x) with domain \{ \psi \in L^2(\mathbb{R}) \mid x \psi \in L^2(\mathbb{R}) \}, which is self-adjoint. The momentum operator \hat{P} = -i \frac{d}{dx} acts on the Sobolev space H^1(\mathbb{R}) and is also self-adjoint. Hamiltonians, such as the Schrödinger operator H = -\frac{d^2}{dx^2} + V(x) for suitable potentials V, are self-adjoint on appropriate domains, representing total energy observables. For bounded self-adjoint operators, spectral projections arise from the resolution of the identity \{E(\Delta)\}_{\Delta \subset \mathbb{R}}, a family of orthogonal projections indexed by Borel sets \Delta, satisfying E(\emptyset) = 0, E(\mathbb{R}) = I, and T = \int_{\sigma(T)} \lambda \, dE(\lambda) in the strong sense. These projections E(\Delta) are orthogonal, with ranges corresponding to generalized eigenspaces, and for singletons \{\lambda\}, \operatorname{ran} E(\{\lambda\}) = \ker(T - \lambda I). The Hellinger-Toeplitz theorem asserts that a symmetric operator defined on the entire Hilbert space \mathcal{H} is bounded (and hence self-adjoint, by closure properties). More generally, any densely defined symmetric operator is closable, meaning its closure \overline{T} (the operator with graph \overline{G(T)}) is also symmetric, and T^{**} = \overline{T}. Unbounded self-adjoint operators often arise as extensions of symmetric operators. Von Neumann's extension theorem characterizes self-adjoint extensions via deficiency subspaces: for a closed symmetric T, define N_\pm = \ker(T^* \mp iI); then T is self-adjoint if and only if \dim N_+ = \dim N_- = 0, and in general, self-adjoint extensions exist if \dim N_+ = \dim N_- < \infty, parameterized by unitary maps from N_+ to N_-. These extensions preserve the physics in quantum models, such as boundary conditions for differential operators.

Functional Calculus and Decompositions

Spectral Theorem

The spectral theorem is a cornerstone of operator theory, providing a decomposition of normal and self-adjoint operators on Hilbert spaces into integrals with respect to spectral measures, analogous to diagonalization of matrices. This theorem enables the construction of functional calculi and reveals deep connections to measure theory and representation theory. It applies primarily to normal operators, which commute with their adjoints, and self-adjoint operators, a key subclass ensuring real spectra. For bounded normal operators on a complex separable Hilbert space H, the spectral theorem asserts that there exists a probability measure space (\Omega, \mu), a unitary operator U: H \to L^2(\Omega, \mu), and a bounded measurable function \phi: \Omega \to \mathbb{C} such that T is unitarily equivalent to the multiplication operator M_\phi given by (M_\phi f)(\omega) = \phi(\omega) f(\omega) for f \in L^2(\Omega, \mu), where the essential range of \phi coincides with the spectrum \sigma(T). Equivalently, T = \int_{\sigma(T)} \lambda \, dP(\lambda), where P is the unique projection-valued measure on the Borel subsets of \sigma(T) such that P(\mathbb{C}) = I and \|P(E)\| \leq 1 for Borel sets E. For unbounded self-adjoint operators T on H with dense domain D(T), the spectral theorem provides a spectral family \{E(\lambda)\}_{\lambda \in \mathbb{R}} of projections such that T = \int_{-\infty}^\infty \lambda \, dE(\lambda), where the integral is understood in the strong sense on D(T) = \{\xi \in H : \int_{-\infty}^\infty \lambda^2 \, d\|E(\lambda)\xi\|^2 < \infty\}, and \sigma(T) = \{\lambda \in \mathbb{R} : E has a discontinuity at \lambda \} \cup closure of points of increase of E. This representation holds unitarily equivalent to multiplication by \lambda on L^2(\mathbb{R}, \mu) for some measure \mu. Proofs of the spectral theorem can be sketched via the holomorphic functional calculus or the Gelfand-Naimark theorem. In the holomorphic approach, for a bounded self-adjoint A, define f(A) = \frac{1}{2\pi i} \oint_{\partial \Omega} f(\zeta) (\zeta - A)^{-1} d\zeta for holomorphic f on a neighborhood \Omega of \sigma(A); positivity preservation and extension to continuous functions on \sigma(A) yield the projection-valued measure via Riesz representation for the algebra generated by A. For general normals, polar decomposition reduces to the self-adjoint case. Alternatively, the Gelfand-Naimark theorem embeds the commutative C*-algebra generated by a normal T into C(\sigma(T)), whose irreducible representations yield the multiplication operator form via universal properties of C*-algebras. A primary consequence is the Borel functional calculus: for Borel measurable f: \sigma(T) \to \mathbb{C}, define f(T) = \int_{\sigma(T)} f(\lambda) \, dP(\lambda) (or strong integral for unbounded T), satisfying \|f(T)\| \leq \|f\|_\infty for bounded f and commuting with T. This extends polynomials in T and enables solving equations like e^{itT}. Stone's theorem follows: every strongly continuous one-parameter unitary group \{U(t)\}_{t \in \mathbb{R}} on H is of the form U(t) = e^{itA} for a unique self-adjoint A, generated via the spectral theorem applied to the infinitesimal generator. A canonical example is the Laplacian -\frac{d^2}{dx^2} on L^2(\mathbb{R}) with domain H^2(\mathbb{R}), an unbounded self-adjoint operator. The Fourier transform \mathcal{F}: L^2(\mathbb{R}) \to L^2(\mathbb{R}), defined by (\mathcal{F} f)(\xi) = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^\infty f(x) e^{-i x \xi} dx, is unitary and conjugates -\frac{d^2}{dx^2} to multiplication by \xi^2, so \sigma\left(-\frac{d^2}{dx^2}\right) = [0, \infty) with continuous spectrum, illustrating wave propagation via plane waves. The spectral theorem's development began with Frigyes Riesz in the 1910s–1920s, who established it for compact self-adjoint operators via integral representations and projection decompositions. The general form for normal and unbounded self-adjoint operators was independently achieved by Marshall Stone in his 1932 monograph and John von Neumann in papers from 1930–1932, motivated by quantum mechanics and extending Hilbert's earlier work on integral operators.

Polar Decomposition

In operator theory, the polar decomposition theorem provides a canonical factorization of bounded linear operators on a Hilbert space into a product of a partial isometry and a positive self-adjoint operator. For a bounded linear operator T on a complex Hilbert space \mathcal{H}, there exists a unique partial isometry U and a unique positive self-adjoint operator |T| such that T = U |T|, where |T| = \sqrt{T^* T} is defined via the spectral theorem applied to the positive self-adjoint operator T^* T. The equation T^* T = |T|^2 holds, and the initial space of U (the subspace where U acts as an isometry) is the closure of the range of |T|, while the final space is the closure of the range of T. This decomposition generalizes the polar form of complex numbers, where a nonzero complex number z factors as z = r e^{i\theta} with r = |z| \geq 0 and |e^{i\theta}| = 1. Uniqueness follows from the fact that |T| is the unique positive square root of T^* T, and U is uniquely determined on the range of |T| by U x = T |T|^{-1} x for x in that range, with U = 0 on the orthogonal complement. For unbounded operators, a similar polar decomposition exists under appropriate conditions. Specifically, if T is a closed densely defined linear operator from a Hilbert space \mathcal{H} to another Hilbert space \mathcal{K}, then T admits a polar decomposition T = U |T|, where |T| = \sqrt{T^* T} is defined on the domain of T^* T using the spectral theorem for unbounded self-adjoint operators, and U is a densely defined partial isometry with initial space the closure of the range of |T|. The uniqueness properties mirror those for the bounded case, with U unique on the range of |T|. The eigenvalues of |T|, known as the singular values of T, play a central role in this decomposition, as they measure the "size" of T in a basis-independent way and coincide with the square roots of the eigenvalues of T^* T. For compact operators, the polar decomposition connects directly to the singular value decomposition (SVD), where if T is compact, then |T| has a pure point spectrum of singular values, and U extends to a full SVD by choosing orthonormal bases aligned with the singular vectors. This provides a concrete example, as the SVD explicitly realizes the polar form T = U \Sigma V^* for matrices, with \Sigma = |T| diagonal and positive.

Singular Value Decomposition

The singular value decomposition (SVD) provides a canonical factorization for compact operators on Hilbert spaces, generalizing the finite-dimensional SVD to infinite dimensions and facilitating analysis in approximation theory and numerical methods. For a compact operator T: H \to K between separable Hilbert spaces H and K, the SVD expresses T as an infinite sum of rank-one operators weighted by singular values, which decay to zero. This decomposition is fundamental for understanding the structure of compact operators and their role in regularizing ill-posed problems. The concept of singular values for integral operators, precursors to the modern SVD, was introduced by Erhard Schmidt in 1907 as eigenvalues of associated symmetric kernels in the context of solving integral equations. The approximation-theoretic significance, particularly the optimality of low-rank truncations, was established by Carl Eckart and Gilbert Young in 1936 for matrices, with extensions to compact operators following naturally from spectral theory. Let T: H \to K be a compact operator between separable Hilbert spaces. There exist orthonormal sequences \{u_n\}_{n=1}^\infty \subset K and \{v_n\}_{n=1}^\infty \subset H, and a non-increasing sequence of non-negative real numbers \sigma_n \to 0 (the singular values) such that Tx = \sum_{n=1}^\infty \sigma_n \langle x, v_n \rangle_H u_n, \quad x \in H. The series converges in the norm topology of K, and the partial sums define finite-rank approximations to T. The operator norm satisfies \|T\| = \sigma_1, the largest singular value. The singular values \sigma_n are precisely the eigenvalues of the positive square root |T| = (T^* T)^{1/2}, counted with multiplicity and ordered decreasingly; this connects the SVD to the polar decomposition, where T = U |T| with U a partial isometry. An operator T belongs to the trace class (the ideal \mathcal{I}_1 of nuclear operators) if and only if \sum_{n=1}^\infty \sigma_n < \infty, in which case the trace norm is \|T\|_1 = \sum_{n=1}^\infty \sigma_n. The finite-rank truncations T_N x = \sum_{n=1}^N \sigma_n \langle x, v_n \rangle_H u_n converge to T in the operator norm, i.e., \|T - T_N\| \to 0 as N \to \infty, since compact operators are precisely the norm limits of finite-rank operators; moreover, the Eckart-Young theorem implies that T_N is the optimal rank-N approximation to T in the operator norm. A prominent example arises with compact integral operators on L^2([a,b]) defined by continuous kernels K(s,t), where Mercer's theorem for symmetric positive kernels yields an eigenexpansion, but the general SVD applies to non-symmetric cases via the sequences \{u_n\} and \{v_n\}. The SVD also enables construction of the Moore-Penrose pseudoinverse T^\dagger = \sum_{\sigma_n > 0} \sigma_n^{-1} \langle \cdot, v_n \rangle_H u_n, which minimizes the least-squares error for solving Tx = y in a stable manner.

Connections to Complex Analysis

Holomorphic Functional Calculus

The holomorphic functional calculus provides a method to define functions of operators in a way that extends the classical notion from scalar functions to bounded linear operators on Banach spaces, allowing the construction of f(T) for a holomorphic function f defined on a neighborhood of the spectrum \sigma(T). This calculus is fundamental in operator theory as it enables the application of complex analysis tools to operators, facilitating the study of their spectral properties and behaviors under analytic mappings. It was developed as part of the integral calculus framework in the mid-20th century, building on earlier work in spectral theory. For a bounded linear operator T on a complex Banach space, the holomorphic functional calculus is constructed using Cauchy's integral formula adapted to the resolvent. Specifically, if f is holomorphic on an open set containing \sigma(T) and \Gamma is a positively oriented contour enclosing \sigma(T) within its region of holomorphy, then f(T) = \frac{1}{2\pi i} \int_\Gamma f(\zeta) R(\zeta, T) \, d\zeta, where R(\zeta, T) = (\zeta I - T)^{-1} is the resolvent operator. This definition ensures that f(T) is a bounded linear operator, and it agrees with the polynomial calculus when f is a polynomial. The construction relies on the analyticity of the resolvent outside \sigma(T) and the holomorphy of f, guaranteeing the integral's convergence and well-definedness independent of the choice of contour \Gamma. Key properties of this calculus include linearity in f, meaning (\alpha f + \beta g)(T) = \alpha f(T) + \beta g(T) for scalars \alpha, \beta and suitable holomorphic f, g; the adjoint relation f(T)^* = \overline{f}(T^*), where \overline{f}(z) = \overline{f(\overline{z})} is the function; and commutativity with the functional calculus, so f(T) commutes with any polynomial in T. For normal operators T on Hilbert spaces, the holomorphic functional calculus coincides with the spectral integral representation from the , given by f(T) = \int_{\sigma(T)} f(\lambda) \, dE(\lambda), where E is the spectral measure of T. This equivalence highlights the calculus's consistency with more specialized decompositions. Practical examples illustrate the utility of the . For instance, the e^T is defined via the with f(z) = e^z, which generates contraction semigroups for certain dissipative operators T. Another special case is the resolvent itself, recovered as f(T) with f(\lambda) = 1/(\lambda - \zeta) for \zeta \notin \sigma(T), underscoring the calculus's foundational role in . For unbounded operators, the Riesz-Dunford calculus extends this framework to sectorial operators—those whose spectrum lies in a sector of the with the resolvent bounded appropriately outside. In this setting, the is defined over contours in the , yielding bounded operators f(A) for holomorphic f in suitable sectors, with analogous properties holding. The integral form of the was pioneered by Nelson Dunford and Jacob Schwartz in their comprehensive treatment during the 1950s, as detailed in their multi-volume work Linear Operators. Their development integrated with operator theory, providing a robust tool for handling non-normal operators beyond approximations.

Runge Approximation

The Runge approximation provides a foundational result in concerning the uniform of holomorphic functions by s on compact sets. Specifically, let \Omega \subset \mathbb{C} be an open set, K \subset \Omega a compact , and f: \Omega \to \mathbb{C} holomorphic. Then, for any \epsilon > 0, there exists a r(z) = P(z)/Q(z), where P and Q are s with no common zeros and all poles of r lie in \mathbb{C} \setminus \Omega, such that \sup_{z \in K} |f(z) - r(z)| < \epsilon. This approximation is uniform on K. In the context of operator theory, the theorem underpins the density properties of the holomorphic functional calculus for a bounded linear operator T on a Banach space, where the spectrum \sigma(T) is contained in a compact set K \subset \Omega with f holomorphic on \Omega. The functional calculus defines f(T) via the Dunford-Cauchy integral f(T) = \frac{1}{2\pi i} \int_\Gamma f(\zeta) (\zeta - T)^{-1} d\zeta, where \Gamma is a contour enclosing K. By Runge's theorem, f(T) can be approximated in the operator norm by r(T), where r is a rational function with poles outside \Omega, and thus avoiding \sigma(T). For operators with compact resolvent, such as those on Hilbert spaces with discrete spectrum accumulating only at zero, the spectrum is compact, allowing polynomials in T to densely approximate the image of the calculus when \mathbb{C} \setminus \sigma(T) is connected. This density extends to Laurent polynomials for spectra contained in annuli around the origin, enabling approximations like f(T) \approx \sum_{k=-n}^m a_k T^k uniformly on the spectrum. Applications of the theorem in operator theory include the approximation of resolvents and functions in spectral decompositions. For instance, the resolvent operator R(\lambda, T) = (\lambda - T)^{-1} for \lambda \notin \sigma(T) is itself rational in T, and Runge's theorem facilitates its polynomial approximation on compact subsets of the resolvent set, which is crucial for numerical methods in solving operator equations and perturbation analyses. Another example arises in approximating exponentials e^{tT} for t > 0, where the f(z) = e^{tz} on a neighborhood of \sigma(T) can be uniformly approximated by rationals on compact sets, yielding stable numerical schemes for generation. The theorem also connects to the numerical range W(T) = \{\langle Tx, x \rangle : \|x\| = 1\}, as quadratic forms can be approximated via the calculus, providing bounds on norms through inclusion. A sketch of the proof relies on : represent f(z) = \frac{1}{2\pi i} \int_{\partial U} \frac{f(\zeta)}{\zeta - z} d\zeta for z \in K and a suitable \partial U in \Omega enclosing K. To approximate, select points a_j in each bounded component of \mathbb{C} \setminus \overline{\Omega} and deform the to pass near these points, generating simple poles at the a_j. Merging multiple poles via partial fractions yields a rational approximant with controlled error, uniform on K by the . The theorem was originally proved by in 1885 for simply connected domains, emphasizing the role of the complement's connectivity in pole placement. It was generalized by Sergei Mergelyan in 1951, who showed that if \mathbb{C} \setminus K is connected (i.e., K has no holes), then holomorphic functions on a neighborhood of K can be uniformly approximated by polynomials alone, without needing rational functions with poles. This extension is particularly impactful in operator theory for operators with connected complements to their spectra, such as operators on Hilbert spaces.

Operator Algebras

C*-Algebras

A C*-algebra is defined as a complex Banach algebra equipped with an involution operation * that satisfies the C*-identity: for every element a in the algebra, \|a^* a\| = \|a\|^2. This condition ensures that the norm is compatible with the involution, distinguishing C*-algebras from more general Banach -algebras. The involution extends the adjoint operation on bounded operators, and the C-identity captures the self-adjointness essential for spectral properties. In the commutative case, every unital commutative is isometrically *-isomorphic to the C(X) of continuous complex-valued functions on a compact X, via the Gelfand transform. The Gelfand transform maps each element a to its Gelfand , the set of complex homomorphisms from the to \mathbb{C}, topologized as the , yielding the . This representation theorem, developed by in the 1940s, highlights the geometric interpretation of commutative as function . For non-commutative , the bounded operators B(\mathcal{H}) on a \mathcal{H} form a universal example, serving as the enveloping algebra for any C*-algebra via the Gelfand-Naimark theorem. This theorem states that every C*-algebra admits a faithful *-representation as a norm-closed -subalgebra of B(\mathcal{H}) for some \mathcal{H}, established by Gelfand and Naimark in 1943 and extended by in 1947 to the general non-commutative setting. Ideals in C-algebras are closed under the involution and form a rich structure; quotients by closed two-sided -ideals inherit the C-algebra properties, enabling the study of extensions and short exact sequences. The Gelfand-Naimark-Segal (GNS) provides a fundamental tool for representations, associating to each (a positive linear functional of norm 1) on a a cyclic *-representation on a Hilbert space. Given a \phi, the builds a pre-Hilbert space from the left ideal of elements with \phi(a^* a) < \infty, completes it to a Hilbert space, and defines operators \pi_\phi(a) \xi = a \xi for \xi in the space, yielding an irreducible representation when \phi is pure. This method, originating from Segal's 1947 work, ensures every has faithful representations from its states. Through representations, the spectral theorem for normal elements in C*-algebras manifests as multiplication operators on an L^2-space over the spectrum. A normal element a (satisfying a^* a = a a^*) in a represented C*-algebra acts as multiplication by a continuous function on its spectrum, generalizing the finite-dimensional diagonalization and relying on the faithful embedding into B(\mathcal{H}). This connects back to normal operators, which are faithfully represented in such algebras. A prominent example is the Toeplitz algebra, the norm closure in B(\ell^2(\mathbb{N})) of the -algebra generated by the unilateral shift operator S e_n = e_{n+1}, where \{e_n\} is the standard basis. It contains the compact operators as an ideal and serves as a model for extensions in index theory. Basic K-theory for C-algebras defines K_0(A) as equivalence classes of projections in matrix algebras over A, and K_1(A) from unitaries in the unitization, providing invariants for classification; this framework, influenced by the , was systematized by Brown, Douglas, and Fillmore in the 1970s for extensions.

Von Neumann Algebras

A von Neumann algebra is defined as a unital -subalgebra M of the bounded linear operators B(\mathcal{H}) on a Hilbert space \mathcal{H} that is closed in the weak operator topology. Equivalently, it is a self-adjoint subalgebra containing the identity operator and closed under weak limits. This closure distinguishes von Neumann algebras from C-algebras, which are defined via norm closure, though von Neumann algebras are the weak closures of C*-algebras in B(\mathcal{H}). The double commutant theorem, a foundational result, characterizes von Neumann algebras via commutants: for any *-subalgebra A \subseteq B(\mathcal{H}), the double commutant A'' = \{ T \in B(\mathcal{H}) \mid T S = S T \ \forall S \in A' \}, where A' is the commutant of A, is itself a von Neumann algebra, and the von Neumann algebra generated by A is precisely A''. This theorem implies that any von Neumann algebra is the double commutant of the set of its elements, providing an algebraic characterization without explicit reference to topology. Projections in a von Neumann algebra M form a complete orthomodular lattice under the operations of least upper bound (supremum) and greatest lower bound (infimum), both existing within M. Two projections p, q \in M are equivalent, denoted p \sim q, if there exists a partial isometry v \in M with initial projection p and final projection q; the partial order on projections is then defined by p \preceq q if p \sim r \leq q for some projection r \in M. This structure allows for a dimension theory analogous to that in type I algebras, enabling comparisons and decompositions. Von Neumann algebras are classified into types based on the equivalence classes of their projections: type I (admitting minimal nonzero projections, finite or infinite dimensional), type II_1 (finite, with no minimal projections but admitting a normalized trace taking values in [0,1]), type II_\infty (infinite, with a semifinite trace taking values in [0,\infty]), and type III (lacking nonzero finite projections, with traces only 0 or \infty). Every von Neumann algebra decomposes uniquely as a direct integral of factors of these types, where a factor is a von Neumann algebra with trivial center Z(M) = \mathbb{C} I. This classification, due to , relies on the lattice of projections and traces to distinguish the types. Tomita-Takesaki theory associates to a von Neumann algebra M and a faithful normal state \phi (or more generally, a cyclic and separating vector \Omega \in \mathcal{H}) a modular operator \Delta > 0 and a modular conjugation J, generating a one-parameter group of automorphisms \sigma_t^\phi (the modular automorphism group) satisfying \sigma_t^\phi(a) = \Delta^{it} a \Delta^{-it} for a \in M. This theory, developed by Tomita and Takesaki in the late and early , reveals the structure of factors through modular theory, with the sub-classification of type III factors determined by the Connes spectrum of the modular automorphism group. It provides a duality between states and automorphisms, essential for understanding non-tracial weights and states in . Examples of von Neumann algebras include the abelian algebra L^\infty(X, \mu) of essentially bounded measurable functions acting by multiplication on L^2(X, \mu), which is an abelian type I von Neumann algebra, a only if X is a point; otherwise, it is diffuse type I. The full algebra B(\mathcal{H}) is the canonical type I_\infty when \dim \mathcal{H} = \infty. Type II_1 factors arise in group-measure space constructions, such as the crossed product L^\infty(X, \mu) \rtimes \Gamma for a free ergodic probability-preserving action of a \Gamma on (X, \mu), admitting a finite . Historically, algebras emerged in the 1930s through John 's work on rings of operators in , with the double commutant theorem appearing in his 1936 paper. The type classification was developed by F. J. Murray and in a series of papers from 1936 to 1943, motivated by the need to generalize decompositions to infinite dimensions. Later advancements, including Tomita-Takesaki theory around 1970, extended the framework to handle type III factors and modular structures.

References

  1. [1]
    [PDF] 7. Operator Theory on Hilbert spaces - KSU Math
    These are often called bounded operators, and the branch of Functional Analysis that studies these objects is called “Operator Theory.” The standard notations ...
  2. [2]
    Introduction to Operator Theory - Georgia Tech Math
    This course covers linear operators on Hilbert space, spectral theory of bounded and unbounded operators, Hilbert spaces, projections, and idempotents.Missing: overview | Show results with:overview
  3. [3]
    [PDF] A Short History of Operator Theory - NYU Stern
    The most thorough history of operator theory of which I am aware is Jean Dieudonné's History of Functional Analysis, on which I draw in this account, along with ...
  4. [4]
    math7334 - Michael Loss
    ... operator theory. The prime example are linear evolution equations. They can be viewed as initial value problems involving linear operators on a Banach space ...
  5. [5]
    [PDF] NARRATIVE STATEMENT ON RESEARCH - University of Toledo
    Aug 17, 2023 · Applications of Operator Theory have also appeared in image processing, signal processing and control theory.
  6. [6]
    [PDF] Operator Theory For Electromagnetics An Introduction
    At its core, operator theory deals with operators, which are mathematical objects that transform one function into another. In electromagnetics, these functions ...
  7. [7]
    [PDF] Lecture Notes on Operator Theory - Fakultät für Mathematik
    Chapter 1 introduces basic definitions of Banach and Hilbert spaces and linear operators on them, following mostly Teschl's book. (§0.2–0.6, §1.1–1.3, and ...
  8. [8]
    [PDF] Operator theory on Hilbert spaces
    Its content is quite standard and this theory can be seen as a special instance of bounded linear operators on more general Banach spaces. 1.1 Hilbert space.
  9. [9]
    [PDF] banach and hilbert space review - Christopher Heil
    We will focus on operators on Hilbert spaces, although compact operators on Banach spaces are equally important. Definition 6.1 (Compact Operator). Let H, K be ...
  10. [10]
    [PDF] FREDHOLM, HILBERT, SCHMIDT Three Fundamental Papers on ...
    Dec 15, 2011 · The papers by Fredholm, Hilbert, and Schmidt advanced integral equations, focusing on Fredholm equations of the second kind, and moving from ...
  11. [11]
    [PDF] Functionalanalysis 1
    Nov 27, 2018 · ... composition ST = S ◦ T : X → Z and observe that. Proposition 2.4. The composition ST of two bounded linear operators S ∈ L(Y,Z) and T ∈. L ...
  12. [12]
    [PDF] Bounded linear operators - Rice University
    If b is essentially bounded, then. Mb determines a bounded linear operator on Lp for each p, with operator norm equal to the L∞ norm kbk∞ of b. Suppose ...
  13. [13]
    [PDF] Spectral theory of differential operators
    ). The operator T2R has compact resolvent (as Ω2R is bounded), so specess T2R = ∅ and. (Remark 5.12) inf specess(T2R ⊕ TC. R) ≡ inf( specess T2R. | {z }.
  14. [14]
    [PDF] Brief summary of functional analysis APPM 5440 Fall 2014 Applied ...
    Oct 27, 2017 · An open mapping is one that maps open sets to open sets. Theorem 4 (Open Mapping Theorem). Let X and Y be Banach spaces. Then any bounded linear ...
  15. [15]
    [1601.02600] A simple proof of the closed graph theorem - arXiv
    Jan 9, 2016 · The proof can be easily extended to Banach spaces. Subjects: Functional Analysis (math.FA). MSC classes: 47A05. Cite as: arXiv:1601.02600 ...
  16. [16]
    [PDF] UNBOUNDED OPERATORS ON HILBERT SPACES Let X and Y be ...
    Therefore if one wants to study unbounded linear operators on a Hilbert space. H that have an adjoint, such operators can not be defined on all of H. Definition ...
  17. [17]
    [PDF] Notes on unbounded operators - UNM Math
    Introduction and examples. Definition 1.1. A linear operator on X is a linear mapping A : D(A) → X defined on some subspace D(A) ⊂ X. A is densely defined ...
  18. [18]
    [PDF] Unbounded operators on Hilbert spaces
    The graph of T is the closure of the graph of T in H1 × H2. Hence T is a closed operator. Lemma 1.3. Let T1 : dom(T1) −→ H2 be a densely defined operator and.
  19. [19]
    [PDF] 11. Unbounded operators and relations - OU Math
    Unbounded operators and relations. Many important operators on Hilbert spaces are not bounded. For example, differential operators on L2(Rn) are never bounded.
  20. [20]
    [PDF] Unbounded Operators on Hilbert Spaces - Amy DeCelles
    Aug 6, 2016 · An unbounded operator T is closable if there is an operator, the closure T of T, whose graph is the closure of the graph of. T. For everywhere- ...<|control11|><|separator|>
  21. [21]
    Symmetric pairs and self-adjoint extensions of operators, with ... - arXiv
    Dec 10, 2015 · We provide a streamlined construction of the Friedrichs extension of a densely-defined self-adjoint and semibounded operator A on a Hilbert ...
  22. [22]
    [PDF] On the Stone-von Neumann Uniqueness Theorem and Its Rami ...
    Abstract. A brief history of the Stone-von Neumann uniqueness theorem and its rami cations is provided. The in uence of this theorem on the development of ...
  23. [23]
    [PDF] Self-adjoint extensions of positive symmetric operators: the Krein ...
    Mar 16, 2018 · The Friedrichs extension has an acknowledged importance in quantum mechanics while Grubb has shown that the Krein-von Neumann extension has a ...
  24. [24]
    [PDF] Operator Algebras and Unbounded Self-Adjoint Operators
    Jun 16, 2015 · The theory of unbounded operators arosed between the 1920s and 1930s. In particular, von Neumann and Marshall H. Stone (1903 − 1989) developed ...
  25. [25]
    [PDF] arXiv:2107.09638v2 [math.FA] 23 Jul 2021
    Jul 23, 2021 · ... operators on. X) and its complement σ(A) := C\ρ(A) are called the operator's resolvent set and spectrum, respectively. The spectrum is ...
  26. [26]
    [PDF] Chapter 9: The Spectrum of Bounded Linear Operators
    In this section, we analyze the spectrum of a compact, self-adjoint operator. The spectrum consists entirely of eigenvalues, with the possible exception of ...
  27. [27]
    [PDF] Rickart's proof of σ(a) 6= ∅ and of the Beurling-Gelfand formula
    Oct 19, 2020 · The spectral radius of a is r(a) = sup{|λ| | λ ∈ σ(a)}. 1.3 Theorem Let A be a unital normed algebra and a ∈ A. Then σ(a) 6= ∅, and r(a) ...<|separator|>
  28. [28]
    [PDF] an elementary proof of the spectral radius formula for matrices
    The spectral radius formula holds for any matrix and any norm: ‖An‖1/n −→ ρ(A). Proof. Theorem 2.1 on the equivalence of norms yields the inequality. C.
  29. [29]
    [PDF] The spectra of the unilateral shift and its adjoint - Jordan Bell
    Apr 3, 2014 · Here we review general statements about the spectrum of a bounded linear operator. If H is a Hilbert space and T ∈ B(H), the spectrum σ(T) of T ...
  30. [30]
    [PDF] Compact Operators on Hilbert Space
    Feb 18, 2012 · • The only possible accumulation point of the set of eigenvalues is 0, and if X is infinite-dimensional it is an accumulation point. • The ...
  31. [31]
    A Commutativity Theorem for Normal Operators - PNAS
    A Commutativity Theorem for Normal Operators. Bent FugledeAuthors Info & Affiliations. January 15, 1950. 36 (1) 35-40. https://doi.org/10.1073/pnas.36.1.35.
  32. [32]
  33. [33]
  34. [34]
  35. [35]
    [PDF] The Cayley Transform and Self–adjoint Extensions - UBC Math
    Nov 27, 2018 · Throughout these notes H is a Hilbert space. We shall use D(B) and R(B) to denote the domain and range, respectively, of the linear operator B.
  36. [36]
  37. [37]
    [PDF] Lecture Notes on the Spectral Theorem - Dartmouth Mathematics
    Aug 10, 2018 · The object of these lectures was to give a reasonably self-contained proof of the Spectral Theorem for bounded normal operators on an infinite ...
  38. [38]
    [PDF] The Spectral Theorem for Unbounded Operators.
    Nov 29, 2001 · We then apply the spectral theorem for bounded normal operators to derive the spectral theorem for unbounded self-adjoint operators. This is ...
  39. [39]
    [PDF] The Spectral Theorem for Self-Adjoint and Unitary Operators
    The Spectral Theorem states that for self-adjoint A and unitary U, there exists a measure space, unitary map Φ, and a, u, such that ΦAΦ-1f(x) = a(x)f(x) and ΦU ...
  40. [40]
    [PDF] Gelfand representation
    May 8, 2013 · Together with the spectral radius formula, this shows that В is a subset of the unit ball of A* and as such can be given the relative weak-* ...
  41. [41]
    [PDF] Stone's Theorem. - Harvard Mathematics Department
    Dec 6, 2001 · Stone proved his theorem to meet the needs of quantum mechanics, where a unitary one parameter group corresponds, via Wigner's theorem to a one ...
  42. [42]
    [PDF] Introduction to the spectral theory
    Example 5.19. The usual Fourier transform is a classical example of a spectral representation. For example, Take H = L2(R) and T = −id/dx with the natural.
  43. [43]
    [PDF] The singular value decomposition of compact operators on Hilbert ...
    Apr 3, 2014 · The purpose of these notes is to present material about compact operators on. Hilbert spaces that is special to Hilbert spaces, ...
  44. [44]
    [PDF] Early History of the Singular Value Decomposition - UC Davis Math
    Jan 17, 2002 · In his treatment of integral equations with unsymmetric kernels, Erhard Schmidt (of Gram-Schmidt fame and a student of Hilbert) introduced the ...
  45. [45]
    [PDF] The Approximation of One Matrix by Another of Lower Rank
    The mathematical problem of approximating one matrix by an- other of lower rank is closely related to the fundamental postulate of factor-theory.
  46. [46]
    [PDF] Essays in analysis Compact operators
    Apr 16, 2020 · If T is a compact operator from a Hilbert space to itself, then its spectrum σ(T) is a discrete set having no limit points except possibly 0.
  47. [47]
    Singular Values of Compact Operators - SpringerLink
    The singular values of a compact operator A are by definition the eigenvalues of (A*A) 1/2. These numbers are important characteristics for compact operators.
  48. [48]
    [PDF] Trace Class Operators - Paul Skoufranis
    Now that we have developed the singular values of a compact operator, we can finally define what it means for a compact operator to be a trace class operator.
  49. [49]
    [PDF] The SVD and applications
    By Mercer's theorem, the corresponding integral operator on. TK has an orthonormal basis of eigenvectors, {ei(t)}. Define: Zi = ∫ 1. 0. Xtei(t)dt. Then Zi are ...
  50. [50]
    [PDF] Lectures on Functional Calculus
    Mar 19, 2018 · Proof of the Spectral Theorem. In the remainder of this chapter we shall present a proof of the spectral theorem. This will happen in four ...
  51. [51]
    [PDF] The Functional Calculus for Sectorial Operators and ... - Uni Ulm
    ... calculus we arrive at is called the Dunford-Riesz calculus, see [Con90,. Chapter VII,3 4]. This can be generalized to situations when is no longer bounded or ...
  52. [52]
    [PDF] Operator Theory - Spectra and Functional Calculi
    Aug 25, 2009 · Useful books include: Tosio Kato, Perturbation theory for linear operators [Kat76]; Paul Halmos,. Introduction to Hilbert space [Hal98]; Edgar ...
  53. [53]
    [PDF] Lecture Notes on C -algebras - Self-serve web hosting
    Jan 3, 2019 · We begin with the definition of a C∗-algebra. Definition 1.1.1. A C∗-algebra A is a (non-empty) set with the following algebraic operations:.
  54. [54]
    [PDF] NOTES ON C∗-ALGEBRAS Contents 1. A first look at C
    Definition 1.2. A C∗-algebra is a Banach ∗-algebra whose norm satisfies the C∗-identity. Remark. Calling these C∗-algebras is already highly suggestive.
  55. [55]
    [PDF] A (Very) Short Course on C -Algebras - Dartmouth Mathematics
    Feb 14, 2024 · ABSTRACT: These are a minimally polished set of lecture notes for a course on C∗-algebras which I've given several times at Dartmouth ...<|separator|>
  56. [56]
    Two-Sided Ideals in Operator Algebras - jstor
    the present paper we prove several results concerning the ideals of uniformly closed self-adjoint algebras of operators on Hilbert spaces,-which we call C*-.
  57. [57]
    [PDF] The Gelfand-Naimark-Segal construction
    Dec 22, 2017 · A representation of a C∗-algebra is a ∗-homomorphism π : A → B(H). Two representa- tions π1 : A → B(H1), π2 : A → B(H2), are (unitarily) ...
  58. [58]
    [PDF] GNS and all that: a rough guide to algebras and states - LSE
    Apr 2, 2021 · This theorem will say that the GNS representation of a C∗-algebra that is defined by a state is irreducible (so that by Schur's lemma: the ...
  59. [59]
    [PDF] C*-algebras and the Spectral Theorem - DiVA portal
    Feb 23, 2022 · The name “Gelfand spectrum” suggests that sp(A) is somehow related to the spectrum of an element, and this is indeed the case. Proposition 3.19.
  60. [60]
    [PDF] The Spectral Theorem - UC Berkeley Mathematics
    We will do this in the context of. C∗-algebra representations. For future reference we note the following proposition which follows from the general theory of ...
  61. [61]
    [PDF] K-theory for C*-Algebras
    Nov 20, 2018 · An introduction to K-theory for C∗- algebras. Vol. 49. London Mathematical Society Student Texts. Cambridge Univer- sity Press, Cambridge ...
  62. [62]
    [PDF] Notes on the type classification of von Neumann algebras - arXiv
    Abstract: These notes provide an explanation of the type classification of von Neumann algebras, which has made many appearances in recent work on ...
  63. [63]
    [PDF] Concise lectures on selected topics of von Neumann algebras - arXiv
    Apr 6, 2020 · The commutant theorem. (M1 ⊗ M2)′ = M′1 ⊗ M′2 holds for tensor products of von Neumann algebras. 1.3 States, weights, and traces. We write M ...