Fact-checked by Grok 2 weeks ago

Spectral theorem

The spectral theorem is a of that establishes the diagonalizability of operators on Hilbert spaces via a . In finite dimensions, it asserts that every real or complex admits an of eigenvectors and is thus orthogonally or unitarily diagonalizable, with all eigenvalues real. This result extends to infinite-dimensional Hilbert spaces, where a bounded T can be represented as T = \int_{\sigma(T)} \lambda \, dE(\lambda), with E a on the \sigma(T). The theorem's finite-dimensional version, often traced to early work on symmetric matrices, provides a complete that simplifies computations involving quadratic forms and matrices. In the unbounded case, it applies to densely defined operators, ensuring a similar form over the real line, which is crucial for handling differential operators. These decompositions guarantee that functions of the operator, such as exponentials used in time evolution, can be defined unambiguously via . Beyond , the spectral theorem underpins key applications in physics and ; in , operators model observables, with eigenvalues representing measurable values and the spectral measure encoding probabilities. It also facilitates the analysis of vibrations in materials and solutions to partial differential equations, where eigenfunctions form natural bases for expansions. The theorem's generalizations to normal operators and its role in unitary representations further extend its influence across operator algebras and .

Finite-dimensional case

Self-adjoint matrices

A matrix, also known as a , is a square matrix A that equals its own , denoted A^*, so A = A^*. This condition implies that the diagonal entries of A are real numbers, while the off-diagonal entries satisfy a_{ji} = \overline{a_{ij}}, where the bar denotes complex conjugation. The spectral theorem for self-adjoint matrices states that every self-adjoint matrix has real eigenvalues and is unitarily diagonalizable. Specifically, there exists a unitary matrix U (satisfying U^* U = I) and a real diagonal matrix D such that A = U D U^*. This decomposition reveals the matrix's eigenvalues on the diagonal of D and an orthonormal basis of eigenvectors as the columns of U. To establish this theorem, the proof proceeds in steps, often by on the matrix dimension. First, all eigenvalues are shown to be real: for any eigenvalue and corresponding eigenvector x with \|x\| = 1, the inner product satisfies \langle Ax, x \rangle = \lambda \langle x, x \rangle = \lambda, but self-adjointness also gives \langle Ax, x \rangle = \langle x, Ax \rangle = \overline{\lambda}, so \lambda = \overline{\lambda} and \lambda is real. Next, eigenvectors corresponding to distinct eigenvalues are orthogonal, as \langle Ax, y \rangle = \lambda \langle x, y \rangle and \langle x, Ay \rangle = \mu \langle x, y \rangle imply (\lambda - \mu) \langle x, y \rangle = 0. The induction step constructs an of eigenvectors by finding one eigenvalue in the current space, adjoining its eigenvector, and applying the theorem recursively to the . Consider the 2×2 self-adjoint matrix A = \begin{pmatrix} 3 & 2+i \\ 2-i & 1 \end{pmatrix}. The characteristic polynomial is \det(A - \lambda I) = (\lambda - 3)(\lambda - 1) - |2+i|^2 = \lambda^2 - 4\lambda - 2, with roots \lambda_1 = 2 + \sqrt{6} and \lambda_2 = 2 - \sqrt{6}, both real. The eigenvectors are found by solving (A - \lambda_i I) \mathbf{v}_i = 0 and normalized to form the columns of U, confirming A = U D U^* with D = \operatorname{diag}(\lambda_1, \lambda_2). The early development of this result traces to , who in proved that real symmetric matrices (a special case of matrices over the ) have real eigenvalues and are orthogonally diagonalizable.

Normal matrices

A matrix A \in \mathbb{C}^{n \times n} is called if it commutes with its , that is, A A^* = A^* A, where A^* denotes the () of A. This condition generalizes the case, where A = A^*, which forms a special subclass of matrices. The theorem for asserts that every is unitarily diagonalizable: there exists a U (satisfying U U^* = I) and a D such that A = U D U^*. The diagonal entries of D are the eigenvalues of A, which may be in general, unlike the real eigenvalues guaranteed for matrices. This diagonalization occurs over the numbers \mathbb{C}, providing an of eigenvectors for the matrix. To establish this result, Schur's triangulation theorem first guarantees that any complex matrix is unitarily similar to an . For a , the normality condition A A^* = A^* A implies that this triangular form must actually be diagonal, as the off-diagonal entries vanish; this follows from the fact that normal matrices preserve the norm of vectors (\|A v\| = \|A^* v\| for all v), which forces the superdiagonal elements to be zero in the Schur form. Thus, the columns of the unitary matrix from Schur's theorem form an of eigenvectors. A concrete example is the 2D by an angle \theta \neq 0, \pi, given by R = \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix}. This matrix is orthogonal, hence (R R^* = R^* R = I), but not unless \theta = 0 or \pi, since R^* = R^T equals R only in those cases. Its eigenvalues are the complex numbers e^{i\theta} and e^{-i\theta}, confirming over \mathbb{C} with an of eigenvectors.

Spectral decomposition

The spectral decomposition of a self-adjoint matrix A on a finite-dimensional expresses A as a of its eigenvalues multiplied by orthogonal s onto the corresponding eigenspaces. Specifically, if \lambda_1, \dots, \lambda_k are the distinct eigenvalues of A, and P_i is the orthogonal projection onto the eigenspace E_{\lambda_i}, then A = \sum_{i=1}^k \lambda_i P_i, where the P_i satisfy P_i P_j = \delta_{ij} P_i and \sum_{i=1}^k P_i = I. This decomposition arises from the existence of an of eigenvectors for operators, allowing A to be unitarily diagonalized. For normal matrices, which commute with their (A A^* = A^* A), the spectral decomposition takes a similar form, but the eigenvalues \lambda_i may be . In this case, there exists a U such that U^* A U = D, where D is diagonal with entries \lambda_i, leading to A = \sum_{i=1}^n \lambda_i |u_i\rangle \langle u_i|, with \{u_i\} forming an of eigenvectors and |u_i\rangle \langle u_i| the rank-one projections. This extends the self-adjoint case, as self-adjoint matrices are a special class of normal matrices with real eigenvalues. The spectral decomposition connects to the singular value decomposition (SVD) of arbitrary matrices through the self-adjoint case. For any m \times n matrix A, the SVD is A = U \Sigma V^*, where the singular values \sigma_i (non-negative diagonal entries of \Sigma) are the eigenvalues of the self-adjoint positive semidefinite matrix \sqrt{A^* A}. Thus, applying the spectral theorem to A^* A yields the singular values and right singular vectors V, while left singular vectors U follow from those of A A^*. The projections P_i in the decomposition are uniquely determined as the spectral projections E(\{\lambda_i\}), which are the unique orthogonal projections onto the eigenspaces satisfying the resolution of the identity for the discrete spectrum. This decomposition enables the for or matrices, where for a f, f(A) = \sum_i f(\lambda_i) P_i, facilitating computations of powers, exponentials, or other functions by reducing to the diagonal form.

Compact self-adjoint operators

Discrete spectrum and eigenvalues

A compact T on a H is a bounded linear that is self-adjoint (T = T^*) and can be approximated in the by a sequence of finite-rank operators. Such operators arise naturally in applications like integral equations and play a central role in due to their "finite-dimensional-like" behavior in infinite dimensions. The theorem for compact operators asserts that the \sigma(T) consists solely of real eigenvalues \{\lambda_n\}, which form a set accumulating only at zero (i.e., |\lambda_n| \to 0 as n \to \infty, assuming they are ordered by decreasing absolute value). Moreover, H admits an \{e_n\} consisting of eigenvectors of T, so T e_n = \lambda_n e_n for each n, and the closed of these eigenvectors is all of H (with the corresponding to the eigenvalue 0 if it occurs). This result extends the finite-dimensional analog, where every is diagonalizable by an . In this basis, the operator admits the spectral decomposition T = \sum_{n=1}^\infty \lambda_n \langle \cdot, e_n \rangle e_n, where the series converges in the operator norm; if T has finite rank, the sum is finite. Proofs of the theorem typically rely on finite-dimensional approximations: since T is the norm limit of finite-rank self-adjoint operators T_k, each of which is diagonalizable, perturbation theory for ensures that the eigenvalues of T are limits of those of the T_k, yielding the discrete spectrum and orthonormal eigenbasis. A example is the on L^2[a,b] defined by (Tf)(x) = \int_a^b K(x,y) f(y) \, dy, where K is a continuous symmetric ; such operators are compact and , with eigenvalues accumulating at zero corresponding to the of K.

Eigenfunction expansion

For a compact T on a separable H, the spectral theorem guarantees the existence of a countable \{e_n\}_{n=1}^\infty consisting of eigenvectors with corresponding real eigenvalues \{\lambda_n\}_{n=1}^\infty satisfying \lambda_n \to 0 as n \to \infty. For any f \in H, the action of T admits the eigenfunction expansion Tf = \sum_{n=1}^\infty \lambda_n \langle f, e_n \rangle e_n, where the series converges in the norm topology of H. This representation follows from the completeness of the eigenbasis and the fact that T maps H into the closed span of the eigenvectors. If an eigenvalue \lambda_k has multiplicity greater than one, the sum incorporates an orthonormal basis for the associated eigenspace, ensuring the expansion remains valid across degenerate cases. The of the expansion relies on the properties of the discrete . The eigenvalues converge absolutely to zero due to the of T, which implies that the only of the is at zero. For the coefficients, ensures \sum_{n=1}^\infty |\langle f, e_n \rangle|^2 \leq \|f\|^2, bounding the series terms and guaranteeing norm of the partial sums to Tf. This framework allows for the approximation of Tf by finite-rank projections onto the eigenspaces, with the error controlled by the of the eigenvalue . A key application arises in the context of positive compact operators on L^2 spaces. states that if K(x,y) is a continuous symmetric on a compact domain, defining the operator (Tf)(x) = \int K(x,y) f(y) \, dy, then K(x,y) = \sum_{n=1}^\infty \lambda_n \phi_n(x) \overline{\phi_n(y)}, where \{\phi_n\} are the orthonormal eigenfunctions and \lambda_n > 0 with \lambda_n \to 0, and the series converges absolutely and uniformly. This expansion facilitates the solution of equations Tf = g by projecting g onto the eigenbasis, yielding f = \sum (\langle g, \phi_n \rangle / \lambda_n) \phi_n for \lambda_n \neq 0, with convergence in L^2. Such decompositions are foundational for numerical methods and approximation theory in solving Fredholm equations of the second kind.

Bounded self-adjoint operators

General statement and continuous spectrum

The spectral theorem for bounded operators provides a canonical decomposition that generalizes the finite-dimensional of matrices to infinite-dimensional s. Specifically, for a bounded A acting on a separable H, the \sigma(A) is a closed subset of the real numbers \mathbb{R}, and there exists a unique resolution of the identity, or \{E(\Delta)\}_{\Delta \subset \mathbb{R}}, such that A = \int_{\sigma(A)} \lambda \, dE(\lambda), where the integral is understood in the strong operator . This representation allows A to be expressed as a "continuous sum" of scalar multiples of orthogonal projections, capturing the operator's action across its entire . The theorem was originally formulated by in his foundational work on . The \sigma(A) decomposes into disjoint parts: the point spectrum (eigenvalues), the continuous , and the singular continuous (though the latter is often empty or absent in simple cases). The continuous \sigma_c(A) consists of those \lambda \in \sigma(A) for which the spectral E(\{\lambda\}) = 0, implying that \lambda is not an eigenvalue and there are no corresponding eigenvectors in H. In this regime, \lambda belongs to the because A - \lambda I fails to be invertible, typically as its range is dense but not closed, leading to approximate eigenvectors that become arbitrarily good in the limit but never exact within the space. A canonical example of an operator with purely continuous spectrum is the operator M_x on L^2[0,1], defined by (M_x f)(t) = t f(t) for f \in L^2[0,1]. Here, \sigma(M_x) = [0,1], which is entirely continuous, as there are no eigenvalues: for any \lambda \in [0,1], the equation M_x f = \lambda f forces f(t) = 0 almost everywhere except possibly at t = \lambda, yielding only the zero function in L^2[0,1]. The spectral projections correspond to by characteristic functions of intervals in [0,1], illustrating how the operator's action is distributed continuously without discrete jumps. When the spectrum includes a continuous part, the spectral theorem implies that no complete of eigenvectors exists in H, unlike the compact case where the spectrum is purely point-like and admits such a basis. This follows from the structure of L^2 spaces, where the Riesz-Fischer theorem identifies completeness but precludes a discrete for operators with continuous multiplicity, necessitating representations for expansions. Thus, vectors in H are decomposed via the continuous measure dE(\lambda), reflecting the operator's "smeared" spectral support.

Projection-valued measures

In the spectral theorem for a bounded A acting on a H, the (also known as a spectral measure or resolution of the ) is a E: \mathcal{B}(\mathbb{R}) \to \mathcal{P}(H), where \mathcal{B}(\mathbb{R}) denotes the \sigma-algebra on \mathbb{R} and \mathcal{P}(H) is the set of orthogonal projections on H. This map assigns to each \Delta \subseteq \mathbb{R} an orthogonal projection E(\Delta) such that E(\mathbb{R}) = I_H, with I_H the on H, and satisfies A = \int_{\mathbb{R}} \lambda \, dE(\lambda), where the is taken in the strong topology. The projections E(\Delta) are and idempotent by definition, ensuring \|E(\Delta)\| \leq 1 for all \Delta. The projection-valued measure E possesses several key properties that underpin its role in . It is countably additive: for a countable collection of pairwise disjoint Borel sets \{\Delta_n\}_{n=1}^\infty, E\left(\bigcup_{n=1}^\infty \Delta_n\right) = \sum_{n=1}^\infty E(\Delta_n) in the strong topology, with E(\emptyset) = 0. For any vector f \in [H](/page/H+), the scalar function \Delta \mapsto \langle f, E(\Delta) f \rangle defines a positive finite measure \mu_f on \mathcal{B}(\mathbb{R}), capturing the "spectral distribution" of f with respect to A. These properties ensure that E behaves analogously to a classical measure but with values in the of projections, facilitating the of A into its spectral components. The construction of the projection-valued measure E for a bounded self-adjoint A relies on foundational results in operator theory. One standard approach uses Stone's theorem on one-parameter unitary groups: the family \{e^{itA}\}_{t \in \mathbb{R}} forms a strongly continuous unitary group on H, and Stone's theorem guarantees the existence of a unique E on \mathbb{R} satisfying e^{itA} = \int_{\mathbb{R}} e^{it\lambda} \, dE(\lambda) for all t \in \mathbb{R}. Differentiating formally with respect to t at t=0 (justified by strong continuity) yields iA = \int_{\mathbb{R}} i\lambda \, dE(\lambda), confirming A = \int_{\mathbb{R}} \lambda \, dE(\lambda). An alternative construction employs the U = (A - iI_H)(A + iI_H)^{-1}, which is a on H; the spectral measure for U on the unit circle can then be mapped back to a measure on \mathbb{R} via the inverse transform, yielding E. For each \Delta \subseteq \mathbb{R}, the H_\Delta = E(\Delta) H (the range of E(\Delta)) is a closed of H under A, known as a reducing subspace: A H_\Delta \subseteq H_\Delta and A (H_\Delta)^\perp \subseteq (H_\Delta)^\perp. These subspaces provide a direct decomposition of H according to the support of A, with the spectrum of A|_{H_\Delta} contained in the closure of \Delta. The projection-valued measure E is unique for the given operator A, as ensured by the uniqueness clause in Stone's theorem and the bijective correspondence it establishes between self-adjoint operators and such measures.

Multiplication operator representation

In the context of bounded s on a H, the spectral theorem provides a concrete realization through unitary equivalence to s on suitable L^2 spaces. This representation, often referred to as the multiplication operator form, expresses the operator in terms of by a real-valued , offering an explicit model for its action and spectral properties. Consider a bounded A on H. The theorem states that there exists a space (\Sigma, \mu), a real-valued bounded \phi: \Sigma \to \mathbb{R}, and a U: H \to L^2(\Sigma, \mu) such that U A U^{-1} = M_\phi, where M_\phi denotes the defined by (M_\phi g)(\sigma) = \phi(\sigma) g(\sigma) for g \in L^2(\Sigma, \mu). This equivalence implies that the action of A on s in H corresponds directly to in the transformed space, simplifying the analysis of eigenvalues and spectral projections. In the case where A admits a cyclic f \in H, the measure \mu can be taken as the scalar spectral measure \mu(B) = \langle E(B) f, f \rangle for Borel sets B \subseteq \mathbb{R}, where E is the associated with A, and \phi(\lambda) = \lambda. This isomorphism preserves the spectrum of the operator: the spectrum \sigma(A) coincides with the essential support of \phi with respect to \mu, defined as the smallest closed set outside which \phi is \mu-almost everywhere zero. Thus, the essential range of \phi captures the possible "eigenvalues" in the continuous case, reflecting the distribution of the spectrum across discrete and continuous components. A canonical example arises in quantum mechanics, where the position operator Q on the Hilbert space L^2(\mathbb{R}, dx) acts as multiplication by the coordinate function: (Q \psi)(x) = x \psi(x). Here, Q is already in its spectral multiplication form with \phi(x) = x and Lebesgue measure \mu = dx, and its spectrum is the entire real line \mathbb{R}, corresponding to all possible position measurements. This representation underscores the theorem's role in modeling observables with continuous spectra. The multiplication operator representation was originally developed by in his foundational 1932 paper, where he established the unitary equivalence for s using integral representations tied to their resolvents.

Direct integral decomposition

The direct integral decomposition provides the canonical representation of a bounded on a separable , accommodating arbitrary multiplicities across the . This formulation, introduced by , asserts that for every bounded A on a H, there exists a standard (\Sigma, \mu) with \Sigma \subseteq \mathbb{R}, a measurable family of Hilbert spaces \{\mathcal{H}_\lambda\}_{\lambda \in \Sigma}, and a U: H \to \int^\oplus_\Sigma \mathcal{H}_\lambda \, d\mu(\lambda) such that U A U^{-1} = \int^\oplus_\Sigma \lambda \, I_\lambda \, d\mu(\lambda), where I_\lambda denotes the identity operator on \mathcal{H}_\lambda and \Sigma = \sigma(A), the of A. The fibers \mathcal{H}_\lambda capture the multiplicity of the spectrum at each point \lambda, defined as \dim \mathcal{H}_\lambda, which may be finite, countably infinite, or uncountably infinite and can vary measurably with \lambda. This multiplicity function reflects the dimension of the eigenspaces for discrete points or the "degeneracy" in the continuous spectrum, allowing the decomposition to handle operators where the spectral behavior changes across different parts of the spectrum. The construction of this direct integral relies on the spectral projection-valued measure E associated with A, which generates a measurable Hilbert bundle over \sigma(A). Specifically, the fibers \mathcal{H}_\lambda are constructed as the ranges of the spectral projections E(\{\lambda\}) for discrete points or via local cyclic decompositions in the continuous case, ensuring the direct integral reproduces the original operator through integration against the identity functions on each fiber. This approach unifies the discrete and continuous cases under a single framework. As an illustrative example, consider a bounded with a simple eigenvalue at \lambda = 0 (multiplicity 1) and continuous spectrum on [1, 2] with multiplicity 2. The direct integral would decompose H into \mathbb{C} \oplus \int^\oplus_{[1,2]} \mathbb{C}^2 \, d\lambda, with the operator acting as multiplication by 0 on the first and by \lambda on the second, highlighting how varying multiplicities are encoded in the dimensions.

Cyclic vectors and multiplicity

In the context of the spectral theorem for bounded self-adjoint operators, a vector f \in H in a separable H is called a cyclic vector for the operator A if the closed of the set \{ p(A) f : p \text{ [polynomial](/page/Polynomial)} \}, denoted \vee \{ p(A) f : p \text{ [polynomial](/page/Polynomial)} \}, equals the entire space H. This condition implies that A has simple spectrum, meaning the spectral multiplicity is at most 1, as the operator can be unitarily equivalent to multiplication by the independent variable on L^2(\sigma(A), \mu) for some probability measure \mu supported on the \sigma(A). The multiplicity of the spectrum of A is determined through the cyclic decomposition of H. Specifically, H decomposes as an orthogonal direct sum H = \bigoplus_{n=1}^m H_n, where each H_n is a cyclic generated by a cyclic for A, and the multiplicity m (possibly infinite) is the minimal number of such copies required to span H; this multiplicity corresponds to the dimension of the in the direct integral over the . A fundamental states that every bounded admits such a into cyclic subspaces, with the multiplicity function constant on the support of the spectral measure if the spectrum is Lebesgue. The spectrum of A is simple if and only if there exists a cyclic vector for A, providing a practical criterion to verify spectral simplicity without computing the full decomposition.

Functional calculus

The Borel functional calculus for a bounded self-adjoint operator A on a Hilbert space H provides a way to define f(A) for any Borel measurable function f: \mathbb{R} \to \mathbb{C}, extending the notion of applying functions to eigenvalues in the finite-dimensional case. Given the spectral projection-valued measure E associated with A by the spectral theorem, such that A = \int_{\mathbb{R}} \lambda \, dE(\lambda), the operator f(A) is defined by the integral f(A) = \int_{\mathbb{R}} f(\lambda) \, dE(\lambda). If f is bounded, then f(A) is a bounded operator with operator norm satisfying \|f(A)\| \leq \|f\|_\infty = \sup_{\lambda \in \mathbb{R}} |f(\lambda)|. This construction is unique and depends continuously on f in the appropriate topology. The satisfies several key algebraic that mirror function on the . For Borel functions f and g, the and multiplication rules hold: f(A) g(A) = (f g)(A), and if h = f \circ g, then h(A) = f(g(A)). Additionally, f(A) commutes with A, since \lambda f(\lambda) = f(\lambda) \lambda with respect to the spectral measure. The gives \mathrm{id}(A) = A, and constant functions yield scalar multiples of the . These ensure that the map f \mapsto f(A) is a ^*-homomorphism from the of bounded Borel functions on (\mathbb{R}to theC^*-[algebra](/page/*-algebra) generated by A$. The Borel calculus is constructed by first establishing a continuous for bounded continuous functions on the \sigma(A), using the Stone-Weierstrass theorem to approximate such functions uniformly by polynomials in A. Polynomials in A are defined in the usual way via or repeated application, and the density of polynomials in the on C(\sigma(A)) allows extension to all continuous functions. This continuous calculus then extends to Borel functions via limits along simple functions or measurable approximations, preserving the integral representation. For operators with compact , the calculus aligns with the continuous case on C_0(\mathbb{R}) restricted to bounded functions vanishing at . Representative applications include defining functions like the for positive operators: if A \geq 0, then f(\lambda) = \sqrt{\lambda} yields \sqrt{A}, which is also positive with (\sqrt{A})^2 = A. Another example is the exponential f(\lambda) = e^{-t\lambda} for t > 0, giving e^{-tA}, which generates the analytic solving the \partial_t u = -A u. These constructions rely on the boundedness of f to ensure f(A) is well-defined and bounded.

Bounded normal operators

Statement for normal operators

A bounded linear operator N on a complex H is called if it commutes with its , that is, NN^* = N^*N. The spectral theorem for bounded operators asserts that for such an N, the \sigma(N) is a nonempty compact of the \mathbb{C}. There exists a unique E, also called a spectral measure, defined on the Borel \sigma-algebra of \mathbb{C}, such that for every Borel set \Delta \subset \mathbb{C}, E(\Delta) is an orthogonal projection on H satisfying E(\mathbb{C}) = I and E(\Delta_1) E(\Delta_2) = E(\Delta_1 \cap \Delta_2) for Borel sets \Delta_1, \Delta_2. Moreover, N admits the representation N = \int_{\sigma(N)} \lambda \, dE(\lambda), where the integral is understood in the strong operator topology, and the support of E is contained in \sigma(N). The projections E(\Delta) satisfy \|E(\Delta)\| \leq 1 and resolve the identity, meaning the family \{E(\Delta)\} is a resolution of the identity with values in the orthogonal projections on H. The spectrum \sigma(N) coincides with the smallest K \subset \mathbb{C} such that E(\mathbb{C} \setminus K) = 0. When N is (hence ), the spectrum \sigma(N) is real and contained in \mathbb{R}, reducing to the spectral theorem for operators. In infinite-dimensional Hilbert spaces, normal operators can exhibit complex spectra, including nonreal eigenvalues.

Unitary equivalence to multiplication operators

A central result in the spectral theorem for bounded normal operators on a separable \mathcal{H} is that every such N is unitarily equivalent to a M_\phi on an L^2 space over a . Specifically, there exists a standard space (X, \mathcal{B}, \mu), a bounded \phi: X \to \mathbb{C}, and a U: \mathcal{H} \to L^2(X, \mu) such that U N U^* f = \phi f for all f \in L^2(X, \mu), with the \sigma(N) coinciding with the essential range of \phi. This representation simplifies the study of N by reducing it to pointwise , preserving norms and spectral properties. The construction of this equivalence proceeds via the measure E associated with N, which is a on \mathbb{C} satisfying N = \int_{\mathbb{C}} \lambda \, dE(\lambda). For separable \mathcal{H}, one can select a cyclic vector (if it exists) to generate a scalar spectral measure, leading to an L^2 over the ; otherwise, the general case employs a direct decomposition over Borel subsets of \sigma(N). This approach ensures the unitary U intertwines the actions of N and M_\phi, with \mu derived from the of the projections in E. To account for possible infinite multiplicity, the full representation takes the form of a direct \mathcal{H} \cong \int^\oplus_{\sigma(N)} \mathbb{C}^{m(\lambda)} \, d\rho(\lambda), where \rho is a scalar measure on \sigma(N) and m: \sigma(N) \to \mathbb{N} \cup \{\infty\} is the multiplicity determining the dimension of each . In this , N acts as by the \lambda on each \mathbb{C}^{m(\lambda)}, capturing the operator's eigenspace dimensions or generalized eigenspaces across the continuous . An illustrative example is the Laurent operator on the H^2(\mathbb{T}) of the unit , defined as by a Laurent polynomial \phi(z) = \sum_{k=-n}^m a_k z^k. This operator is , and by the spectral theorem, it is unitarily equivalent to by \phi on L^2(\mathbb{T}, d\theta/2\pi), with given by the image of \phi on the unit . For the specific case of the bilateral shift ( by z), the is the unit , and the multiplicity is 1 everywhere. The Fuglede-Kadison provides a key analytic tool arising from this , particularly for invertible operators in finite von Neumann factors. Defined as \Delta(N) = \exp(\tau(\log |N|)), where \tau is the , it leverages the form to compute \Delta(N) = \exp\left( \int_X \log |\phi| \, d\mu \right), generalizing the classical and facilitating estimates in operator inequalities. This is multiplicative and unitarily invariant, reflecting the spectral structure.

Unbounded self-adjoint operators

Statement and domain considerations

In a complex H, an unbounded A is defined as a densely defined closed linear A: D(A) \to H, where the domain D(A) is a dense of H, the domain of the adjoint coincides with D(A^*) = D(A), and A satisfies \langle Af, g \rangle = \langle f, Ag \rangle for all f, g \in D(A). This dense domain condition ensures that A can be extended continuously in a weak sense, distinguishing unbounded operators from bounded ones, where the domain is the entire space H. The spectral theorem for such operators asserts the existence of a unique E, also called a , defined on the Borel \sigma-algebra of \mathbb{R}, such that the operator A is given by Af = \int_{\mathbb{R}} \lambda \, dE(\lambda) f for all f \in D(A), where the integral is understood in the strong sense. The domain D(A) consists precisely of those vectors f \in H for which the \int_{\mathbb{R}} |\lambda|^2 \, d\mu_f(\lambda) < \infty, with \mu_f(B) = \langle E(B) f, f \rangle denoting the scalar spectral measure induced by f. This formulation captures the operator's action through a "weighted" L^2 condition with respect to the spectral measure, ensuring the integral converges. The spectrum \sigma(A) of A is the closed subset of \mathbb{R} supporting the measure E (i.e., the smallest closed set S \subseteq \mathbb{R} such that E(\mathbb{R} \setminus S) = 0), and it may be unbounded, reflecting the operator's potentially infinite range. Considerations of the essential spectrum involve parts where the spectral projections have infinite-dimensional range or where the operator cannot be approximated by compact perturbations, but the theorem guarantees the spectrum lies on the real line due to self-adjointness. A canonical example is the momentum operator P = -i \frac{d}{dx} on the Hilbert space L^2(\mathbb{R}), defined on the dense Sobolev domain D(P) = H^1(\mathbb{R}) = \{ f \in L^2(\mathbb{R}) : f' \in L^2(\mathbb{R}) \}, which is self-adjoint and has spectrum \sigma(P) = \mathbb{R}. Here, the spectral measure E corresponds to multiplication by the identity function in the Fourier representation, with domain elements satisfying \int_{\mathbb{R}} |\xi|^2 |\hat{f}(\xi)|^2 \, d\xi < \infty, where \hat{f} is the of f. This theorem was originally established by Marshall Stone in 1932, building on his work on strongly continuous one-parameter unitary groups generated by self-adjoint operators.

Spectral form and resolvent

For an unbounded self-adjoint operator A on a separable Hilbert space \mathcal{H}, the spectral theorem guarantees the existence of a unique projection-valued measure E, supported on the real line \mathbb{R}, such that the domain D(A) is the set of vectors f \in \mathcal{H} satisfying \int_{\mathbb{R}} |\lambda|^2 \, d\|E(\lambda) f\|^2 < \infty, and the action of A on this domain is represented by the spectral integral Af = \int_{\mathbb{R}} \lambda \, dE(\lambda) f. This integral form diagonalizes A with respect to the spectral measure E, allowing the operator to be understood as multiplication by the identity function on a suitable L^2 space. The projection-valued measure E is countably additive and satisfies E(\mathbb{R}) = I, the identity operator on \mathcal{H}. The resolvent operator R(\zeta, A) = (A - \zeta I)^{-1}, defined for \zeta \in \mathbb{C} \setminus \sigma(A), admits an integral representation in terms of the spectral measure: R(\zeta, A) = \int_{\mathbb{R}} \frac{1}{\lambda - \zeta} \, dE(\lambda). This expression shows that R(\zeta, A) is a bounded operator analytic in the resolvent set \rho(A) = \mathbb{C} \setminus \sigma(A), with the spectrum \sigma(A) consisting precisely of those points where the resolvent fails to exist as a bounded operator; specifically, \sigma(A) is the closure of the support of E. The resolvent integral converges in the strong operator topology for \Im \zeta \neq 0 and provides a means to recover the spectral measure via limits, such as Stone's formula for the density of states. Stone's theorem establishes a deep connection between self-adjoint operators and one-parameter unitary groups, stating that A is self-adjoint if and only if there exists a strongly continuous group of unitaries \{ U(t) \}_{t \in \mathbb{R}} on \mathcal{H} generated by A (in the sense that U(t) = e^{itA} for bounded functions), satisfying U(t) = \int_{\mathbb{R}} e^{it\lambda} \, dE(\lambda) for all t \in \mathbb{R}. This representation links the time evolution in quantum mechanics to the spectral decomposition, with the generator A determining the frequencies in the integral. A concrete illustration arises in the quantum harmonic oscillator, where the self-adjoint H = -\frac{d^2}{dx^2} + x^2 on L^2(\mathbb{R}) has discrete spectrum \sigma(H) = \{2n + 1 \mid n = 0, 1, 2, \dots \} and orthonormal eigenbasis \{\psi_n\} consisting of . The spectral measure is E = \sum_{n=0}^\infty E(\{2n+1\}) with E(\{2n+1\}) = |\psi_n\rangle \langle \psi_n|, yielding the resolvent R(\zeta, H) = \sum_{n=0}^\infty \frac{1}{2n + 1 - \zeta} |\psi_n\rangle \langle \psi_n| for \zeta \notin \sigma(H), which explicitly resolves (H - \zeta I)^{-1} via the discrete spectral decomposition.

Functional calculus for unbounded operators

The Borel functional calculus provides a framework for defining functions of unbounded self-adjoint operators on a Hilbert space, extending the spectral theorem to construct new operators from Borel measurable functions. For an unbounded self-adjoint operator A with spectral resolution of the identity E, given a Borel measurable function f: \mathbb{R} \to \mathbb{C}, the operator f(A) is defined by the integral representation f(A) x = \int_{\mathbb{R}} f(\lambda) \, dE(\lambda) x for all x in the domain D(f(A)) = \left\{ x \in H \ \middle|\ \int_{\mathbb{R}} |f(\lambda)|^2 \, d \langle E(\lambda) x, x \rangle < \infty \right\}, where H is the Hilbert space and \langle \cdot, \cdot \rangle denotes the inner product. This domain is dense in H provided f is not zero almost everywhere with respect to the spectral measure induced by E. The properties of this calculus mirror those for bounded operators but account for potential unboundedness when f is unbounded. Specifically, if f is real-valued and Borel measurable, then f(A) is self-adjoint on D(f(A)); the calculus is multiplicative, meaning f(A) g(A) = (f g)(A) where defined; and it commutes with A under suitable conditions on f. Moreover, the resolvent operators fit into this framework via the Herglotz representation: for \operatorname{Im} z > 0, the resolvent R(z, A) = (A - z I)^{-1} satisfies R(z, A) = \int_{\mathbb{R}} \frac{1}{\lambda - z} \, dE(\lambda), which is an H^\infty-function in the upper half-plane with positive imaginary part, enabling and representations central to the . A key aspect of the spectral theorem's completeness is the ability to define inverses and pseudo-inverses through this , such as |A|^{-1} via f(\lambda) = 1/|\lambda| for \lambda \neq 0, with consisting of vectors where the corresponding is finite; this is bounded if 0 is not an eigenvalue of A. An illustrative example is the Gaussian e^{-t A^2} for t > 0, defined by f(\lambda) = e^{-t \lambda^2}, which generates solutions to parabolic partial differential equations like the u_t + A^2 u = 0 with initial data in suitable Sobolev spaces, where A might represent the Laplacian on a .

References

  1. [1]
    [PDF] What Does the Spectral Theorem Say? - NYU Stern
    A more precise statement of the result is that every Hermitian matrix is unitarily equivalent to a diagonal one. The spectral theorem is widely and correctly ...
  2. [2]
    [PDF] Unit 17: Spectral theorem
    17.5. This immediately implies the spectral theorem. Theorem: Every symmetric matrix A has an orthonormal eigenbasis. Proof. Wiggle A so that all eigenvalues of ...
  3. [3]
    The Spectral Theorem
    Let A be an n × n matrix of real numbers. Then A is orthogonally diagonalizable if and only if A is symmetric.Missing: mathematics | Show results with:mathematics
  4. [4]
    [PDF] Proof of the spectral theorem
    Nov 5, 2013 · If dimV = 0, then S = 0 and there are no eigenvalues; the theorem says that the zero vector space is an empty direct sum, which is true by ...
  5. [5]
    [PDF] Spectral theorems, SVD, and Quadratic forms
    Definition. The spectrum of a square matrix A is the set of its eigenvalues. Theorem (Borel functional calculus). Let A be an n×n ...Missing: mathematics | Show results with:mathematics
  6. [6]
    [PDF] Spectral Decomposition of Quantum-Mechanical Operators
    In this work, applications of the spectral theorem for self-adjoint operators in quantum mechanics are examined. While classical physics provides ...
  7. [7]
    [PDF] Applications of Spectral Theory in the Material Sciences Math 7390 ...
    It describes, for example, fundamental modes of vibration or fundamental shapes that are preserved through time in the dynamics of material structures.
  8. [8]
    Self-Adjoint Matrix -- from Wolfram MathWorld
    A matrix A for which A^(H)=A^(T)^_=A, where the conjugate transpose is denoted A^(H), A^(T) is the transpose, and z^_ is the complex conjugate.
  9. [9]
    [PDF] Diagonalization of a general 2 × 2 hermitian matrix
    Diagonalization of a general 2 × 2 hermitian matrix. Consider a general 2 × 2 hermitian matrix. A = a c c∗ b. , where a and b are real numbers and c is a ...
  10. [10]
    Matrix Analysis - Roger A. Horn, Charles R. Johnson - Google Books
    This second edition of this acclaimed text presents results of both classic and recent matrix analysis using canonical forms as a unifying theme and ...
  11. [11]
    [PDF] The Spectral Theorem for normal linear maps 1 Self-adjoint or ...
    Mar 14, 2007 · Every eigenvalue of a self-adjoint operator is real. Proof. Suppose λ ∈ C is an eigenvalue of T and 0 = v ∈ V the corresponding eigenvector such ...Missing: authoritative source<|separator|>
  12. [12]
    [PDF] The Spectral Theorem Let V be a finite-dimensional inner-product ...
    The Spectral Theorem states that a linear operator T on a finite-dimensional inner product space is orthogonally diagonalizable if and only if T is normal.
  13. [13]
    [PDF] Math 108b: Notes on the Spectral Theorem
    The Spectral Theorem states that a linear operator T is normal (complex) or self-adjoint (real) if and only if it has an orthonormal basis of eigenvectors.
  14. [14]
    [PDF] 10. The Spectral Theorem - OU Math
    The spectral theorem states that in finite-dimensional Hilbert spaces, a normal matrix can be diagonalized. In infinite-dimensional spaces, a continuous analog ...
  15. [15]
    [PDF] Eigenvalues of Self Adjoint Operators: Spectral theorem ... - IIT Madras
    Mar 11, 2004 · Spectral theorem and Singular Value Decomposition ... Now we prove the spectral theorem for a self adjoint operator on a finite dimensional.
  16. [16]
    [PDF] Spectral theorem for unbounded self-adjoint operator
    Thus, the theorem says that every unbounded self-adjoint operator admits a spectral resolution. In fact, this spectral resolu- tion is unique (that is, the ...
  17. [17]
    [PDF] Lecture Notes on the Spectral Theorem - Dartmouth Mathematics
    Aug 10, 2018 · Here we want to look at a version of the spectral theorem for unbounded self- adjoint operators in a Hilbert space H. We are only going to ...
  18. [18]
    [PDF] methods of - modern mathematical physics - 1: functional analysis
    This book is the first of a multivolume series devoted to an exposition of func- tional analysis methods in modern mathematical physics.
  19. [19]
    [PDF] spectral theorem for compact, self-adjoint operators
    The following theorem is known as the spectral theorem for compact, self-adjoint operators. We will give two proofs which connect as much.
  20. [20]
    [PDF] 18.102 S2021 Lecture 22. The Spectral Theorem for a Compact Self ...
    May 11, 2021 · We discussed previously that for a self-adjoint operator, the spectrum is contained within a line segment on the real line, and in the finite- ...
  21. [21]
    A Course in Functional Analysis - SpringerLink
    John B. Conway. Pages 124-165. Linear Operators on a Banach Space. John B. Conway. Pages 166-186. Banach Algebras and Spectral Theory for Operators on a Banach ...
  22. [22]
    [PDF] Chapter 9: The Spectrum of Bounded Linear Operators
    The spectral theorem for compact, self-adjoint operators states that any x ~ can be expanded in the form ( .7), and that A can be expressed as a sum of.<|control11|><|separator|>
  23. [23]
    [PDF] The Spectral Theorem for Self-Adjoint and Unitary Operators
    If A ∈ L(H) is self-adjoint, there exists a measure space (X, F,µ), a ... Let U be a unitary operator on a Hilbert space H, and consider. (C.1). AN = 1.Missing: statement | Show results with:statement
  24. [24]
    [PDF] The Spectral Theorem - UC Berkeley Mathematics
    In (ii), Φ(f) will also be denoted by f(T). A projection-valued measure (or resolution of the identity) P on a measurable space (X, B) is a map. P : B→B(H) ...
  25. [25]
    [PDF] chapter ix projection-valued measures definition
    DEFINITION Let S be a set and let B be a σ-algebra of subsets of S. We refer to the elements of B as Borel subsets of S and we call the pair.Missing: spectral construction
  26. [26]
    [PDF] Spectral Measures and the Spectral Theorem - UChicago Math
    We develop properties about Hilbert spaces and spectral measures in order to give a generalization of the spectral theorem to infinite dimensions.
  27. [27]
    [PDF] 13. Stone's theorem and quantum dynamics - OU Math
    Stone's Theorem asserts that, conversely, every strongly continuous unitary group is of this type. Recall from (D) that a unitary group is a map U : R → B(H) ...
  28. [28]
    [PDF] Spectral Theorem for Bounded Self-Adjoint Operators.
    We state the spectral theorem for bounded self-adjoint operators in the cyclic case. We also compute the spectrum and the spectral measure in two concrete ...
  29. [29]
    [PDF] Spectral theory in Hilbert spaces (ETH Zürich, FS 09) E. Kowalski
    done in the proof of the spectral theorem for unbounded self-adjoint operators, where it will be shown that the spectrum (unsurprisingly) is real, so any ...
  30. [30]
    [PDF] BOUNDED OPERATORS - Math@LSU
    Definition A family of projections obeying (a)–(c) is called a (bounded) projection-valued measure (p.v.m.). We remark that (d) follows from (a) and (c) by ...
  31. [31]
    [PDF] Second lecture: The spectral theorem
    Sep 17, 2018 · In this lecture, we will state the spectral theorem for bounded self-adjoint operators and sketch a proof. We closely follow Chapter VII of M.
  32. [32]
    [PDF] Spectral Theory - Michael Taylor
    Appendix A contains a proof of the spectral theorem for a compact, self- adjoint operator A on a Hilbert space H. In that case, H has an orthonor- mal basis {uj} ...
  33. [33]
    [PDF] Lectures on Functional Calculus
    Mar 19, 2018 · for bounded self-adjoint operators. Indeed, for a bounded self-adjoint operator. A on a Hilbert space H we first find the continuous ...
  34. [34]
    [PDF] 13 Spectral theory
    Theorem 13.21. A bounded self-adjoint operatorT has a unique continuous functional calculus. Proof: SinceT is self-adjoint the spectrum σ(T) is contained ...
  35. [35]
  36. [36]
    [PDF] some notes on the spectral theorem
    The point of normal operators is that self-adjoint, anti-self-adjoint and orthogonal, resp. unitary, operators are all examples of normal operators. So the ...
  37. [37]
    SPECTRAL MULTIPLICITY FOR DIRECT INTEGRALS OF NORMAL ...
    The theory of spectral multiplicity solves the unitary equivalence problem for normal operators. Briefly recalling the theory, one associates with each ...
  38. [38]
    [PDF] 1. The spectral theorem for unitary operators
    Any Laurent polynomial p(z) can be written z−N q(z) where N is an integer and q(z) is a polynomial. Hence p(V ) = V −N q(V ) . We know from the lecture that q(σ ...
  39. [39]
    Determinant Theory in Finite Factors - jstor
    FUGLEDE and R. KADISON, On determinants and a property of the trace infinite factors,. Proc. Nat. Acad. Sci., U. S. A., vol. 37 (1951), pp. 425-431. 5. E ...
  40. [40]
    Fuglede–Kadison determinant: theme and variations - PNAS
    We review the definition of determinants for finite von Neumann algebras, due to Fuglede and Kadison [Fuglede B, Kadison R (1952) Ann Math 55:520–530],
  41. [41]
    [PDF] The Spectral Theorem for Unbounded Operators.
    Nov 29, 2001 · For example the operator D = zi dd}on L2(R). This operator is not defined on all of L2, and where it is defined it is not bounded as an operator ...Missing: differential | Show results with:differential
  42. [42]
    The spectral theorem and its converses for unbounded symmetric ...
    Dec 20, 2011 · The spectral theorem for finite-dimensional self-adjoint operators (ie Hermitian matrices, when viewed in coordinates), which provides a sequence.
  43. [43]
    The spectral theorem (Chapter 2) - Spectral Theory and Differential ...
    Introduction. The general spectral theorem for self-adjoint operators was proved independently by Stone and von Neumann during the period 1929–1932.
  44. [44]
    [PDF] Chapter 4 Spectral theory for self-adjoint operators
    In this chapter we develop the spectral theory for self-adjoint operators. As already seen in Lemma 2.2.6, these operators have real spectrum, however much ...
  45. [45]
    [PDF] Spectral theory and applications. An elementary introductory course ...
    Jan 7, 2023 · We say in this case that T0 is not essentially selfadjoint. Application 2: The harmonic oscillator. We can start from. H0 = −∆ + |x|2 + 1 ,.
  46. [46]
    [PDF] Schrödinger Operators and their Spectra - GWDG
    This lecture begins with a brief overview about the spectral theorem and its conse- quences for the spectrum of self-adjoint operators in Hilbert spaces.
  47. [47]