Fact-checked by Grok 2 weeks ago

Normal operator

In functional analysis, a normal operator on a complex Hilbert space is a bounded linear operator T that commutes with its adjoint T^*, satisfying T T^* = T^* T. This property ensures that normal operators preserve the norm of vectors under application, meaning \| T x \| = \| T^* x \| for all x in the space. Examples of normal operators include self-adjoint operators (where T = T^*) and unitary operators (where T^* = T^{-1}), both of which play central roles in quantum mechanics and representation theory. The significance of normal operators lies in the , which states that any normal operator on a finite-dimensional admits an of eigenvectors, allowing it to be diagonalized in an appropriate basis. In infinite-dimensional settings, the theorem extends to a involving projections onto eigenspaces or more general spectral measures, facilitating the analysis of even when they are not discrete. Additionally, the eigenspaces of a normal operator for T coincide with those for T^*, and the kernels of T and T^* are equal. These features make normal operators a cornerstone for studying operator algebras and applications in physics, such as observables in formulations of .

Definition and Fundamentals

Definition

A Hilbert space H is a complete inner product space over the complex numbers, equipped with an inner product \langle \cdot, \cdot \rangle that induces a norm \|x\| = \sqrt{\langle x, x \rangle} and allows for the definition of orthogonality and convergence. In this setting, a bounded linear operator N: H \to H is a linear map satisfying \|N x\| \leq M \|x\| for some constant M \geq 0 and all x \in H, with the operator norm given by \|N\| = \sup_{\|x\|=1} \|N x\|. The operator N^* of a bounded linear N on H is uniquely defined by the \langle N x, y \rangle = \langle x, N^* y \rangle for all x, y \in H, ensuring that N^* is also bounded with \|N^*\| = \|N\|. A bounded linear N on a complex H is called if it commutes with its , that is, N N^* = N^* N. operators, which satisfy N = N^*, form a special case of operators, as they trivially commute with themselves. The concept of normal operators was introduced by in the early 1930s as part of his foundational work on the mathematical structure of , particularly in developing the for operators on Hilbert spaces. A key consequence is the , which allows normal operators to be diagonalized in a suitable sense, facilitating their analysis in physical and mathematical applications (detailed in later sections).

Basic Examples

A fundamental example of a normal operator arises in the context of operators on the L^2(\mathbb{R}). For a bounded complex-valued f, the operator M_f is defined by (M_f g)(x) = f(x) g(x) for all g \in L^2(\mathbb{R}). The of M_f is M_{\bar{f}}, the by the \bar{f}. These operators commute because M_f M_{\bar{f}} = M_{|f|^2} = M_{\bar{f}} M_f, where |f|^2 is real-valued and thus commutes with , making M_f . Orthogonal projections provide another straightforward class of normal operators. Let H be a and P the orthogonal onto a closed K \subseteq H. Such a satisfies P^2 = P and is , meaning P = P^*. operators are normal because P P^* = P^2 = P = P^* P, illustrating the commutativity condition directly. In sequence spaces, diagonal operators on \ell^2(\mathbb{N}) serve as basic normal examples. Consider the operator D defined on the standard \{e_n\} by D e_n = \lambda_n e_n, where \{\lambda_n\} is a bounded complex sequence; in matrix terms, D is diagonal with entries \lambda_n. The D^* has diagonal entries \bar{\lambda_n}, and both being diagonal ensures they commute: D D^* e_n = |\lambda_n|^2 e_n = D^* D e_n, confirming . Unitary operators form a significant subclass of normal operators. A bounded linear operator U on a H is unitary if U^* U = U U^* = I, the identity . This relation immediately implies U U^* = U^* U, satisfying the condition, as the vanishes by definition. To contrast, consider the unilateral on \ell^2(\mathbb{N}), which fails to be normal. Defined by S e_n = e_{n+1} for the \{e_n\}, its is the backward shift S^* e_1 = 0 and S^* e_{n+1} = e_n for n \geq 1. Computing on the basis vector e_1, S S^* e_1 = S(0) = 0, but S^* S e_1 = S^* e_2 = e_1 \neq 0, so S S^* \neq S^* S, violating the commutativity required for .

Core Properties

Commutativity and Adjoint Relations

A fundamental consequence of the normality condition N^* N = N N^* for a bounded normal operator N on a is the equality of norms \|N\| = \|N^*\|. To see this, consider an arbitrary vector x with \|x\| = 1. Then, \|N^* x\|^2 = \langle N^* x, N^* x \rangle = \langle N N^* x, x \rangle = \langle N^* N x, x \rangle = \langle N x, N x \rangle = \|N x\|^2, which holds because N commutes with its . Taking the supremum over all such unit vectors yields \|N\| = \|N^*\|. Another key relation arising from normality is given by Fuglede's commutativity theorem: if N is a normal operator and A is a bounded operator such that A N = N A, then A N^* = N^* A. This result extends the commuting property to the adjoint and plays a role in broader operator theory. A generalization is Putnam's theorem, which states that if M and N are normal operators and T is bounded with T M = N T, then T^* M = N T^*. The set of normal operators is closed under polynomial functions. That is, if N is normal and p(z) is any polynomial with complex coefficients, then p(N) is also normal. The proof relies on the fact that polynomials in N and N^* can be formed such that p(N)^* = p(N^*), and since N and N^* commute, so do p(N) and p(N)^*. For instance, N - \lambda I is normal for any scalar \lambda. Normal operators also exhibit special behavior with respect to isometries. An isometry V satisfies V^* V = I, and if V is additionally normal, then V V^* = I, making V unitary. This follows directly from the normality condition V V^* = V^* V = I, ensuring V is both isometric and co-isometric. Unitary operators, being normal isometries, exemplify this relation.

Norm and Spectrum Basics

The of a bounded normal operator N on a is defined as the set \sigma(N) = \{\lambda \in \mathbb{C} : N - \lambda I \text{ is not invertible}\}. For any bounded linear operator on a complex , the is a non-empty compact of \mathbb{C}. A key property of normal operators is that their spectrum coincides with the approximate point spectrum \sigma_{ap}(N), defined as the set of \lambda \in \mathbb{C} for which there exists a sequence of unit vectors \{x_n\} such that \|(N - \lambda I)x_n\| \to 0 as n \to \infty. This equivalence holds because, for \lambda \in \sigma(N), if \lambda is not an eigenvalue, the normality of N - \lambda I ensures the range is dense while the kernel is trivial, placing \lambda in the approximate point spectrum; eigenvalues are already in \sigma_{ap}(N). The C*-algebra generated by a normal operator N and the identity I, denoted C^*(N, I), is commutative. This follows from the fact that N commutes with its adjoint N^*, so the polynomials in N and N^* commute, and the closure in the operator norm preserves commutativity. In a unital commutative C*-algebra, the spectral radius r(a) = \sup\{|\lambda| : \lambda \in \sigma(a)\} equals the norm \|a\| for any element a. For a normal operator N, Gelfand's formula gives r(N) = \lim_{n \to \infty} \|N^n\|^{1/n}. Since N^n is normal for each n and C^*(N, I) is commutative, \|N^n\| = r(N^n) = [r(N)]^n, so \lim_{n \to \infty} \|N^n\|^{1/n} = r(N), yielding r(N) = \|N\|. Thus, \sup\{|\lambda| : \lambda \in \sigma(N)\} = \|N\|.

Finite-Dimensional Operators

Diagonalization in Finite Dimensions

In finite-dimensional complex Hilbert spaces, a cornerstone result establishes that normal operators are unitarily diagonalizable. Specifically, every normal operator T on \mathbb{C}^n, satisfying T^* T = T T^*, is unitarily equivalent to a . That is, there exists a U such that U^* T U = D, where D is diagonal with entries consisting of the eigenvalues of T. This diagonalizability can be proved by induction on the dimension n of the space. For the base case n = 1, T is simply multiplication by a scalar, which is diagonal. Assume the result holds for dimensions less than n. For dimension n, let \lambda be an eigenvalue of T with corresponding eigenspace E_\lambda, and let P be the orthogonal projection onto E_\lambda with Q = I - P projecting onto the orthogonal complement. Decompose T = P T P + Q T Q, noting that Q T Q is normal on the subspace of dimension less than n, so by the induction hypothesis, it is unitarily diagonalizable on that subspace with an orthonormal basis of eigenvectors. Extending this basis by an orthonormal basis of eigenvectors for P T P (which acts as scalar multiplication by \lambda), yields an orthonormal basis of eigenvectors for T on the full space, preserving unitarity. The diagonal entries of D are precisely the eigenvalues of T, and for normal operators, the algebraic multiplicity of each eigenvalue equals its geometric multiplicity, ensuring no deficiency in the eigenspace dimensions. This equivalence implies that the eigenvectors form an , as unitary transformations preserve inner products. In contrast, non-normal operators on \mathbb{C}^n are not necessarily diagonalizable and may require the Jordan canonical form, which introduces off-diagonal 1's in blocks corresponding to eigenvalues with deficient eigenspaces, reflecting generalized eigenvectors rather than a pure diagonal .

Unitary Equivalence to Diagonal Form

A fundamental extension of the unitary diagonalization of individual normal operators concerns families of such operators that commute with one another. Specifically, let N_1, \dots, N_k be normal operators on a finite-dimensional complex \mathcal{H}, satisfying N_i N_j = N_j N_i for all i, j = 1, \dots, k. Then there exists a U: \mathcal{H} \to \mathcal{H} such that U^* N_i U is diagonal for each i = 1, \dots, k. This simultaneous unitary diagonalization implies the existence of an of \mathcal{H} consisting of common eigenvectors for all N_i, where the diagonal entries correspond to the joint eigenvalues. The proof relies on the preservation of eigenspaces under commuting operators and proceeds by induction on k, the number of operators. For k=1, the result follows from the unitary diagonalizability of a single normal operator, as established in the spectral theorem for finite dimensions. Assume the statement holds for k-1 operators. Consider an eigenvalue \lambda of N_k with corresponding eigenspace E_\lambda. Since each N_i (for i < k) commutes with N_k, it maps E_\lambda to itself and restricts to a normal operator on this invariant subspace. Moreover, these restrictions commute among themselves and are diagonalizable. By the induction hypothesis, there is a unitary operator on E_\lambda that simultaneously diagonalizes the restrictions N_1|_{E_\lambda}, \dots, N_{k-1}|_{E_\lambda}. Extending this across all eigenspaces of N_k yields a full orthonormal basis of common eigenvectors for the family. This construction ensures the overall transformation is unitary, preserving the inner product structure. This theorem finds a key application in , where operators (a subclass of normals) represent physical observables. observables are termed compatible, allowing simultaneous precise measurements; their common eigenbasis corresponds to states that are simultaneous eigenstates, with joint eigenvalues yielding possible measurement outcomes. For instance, in the quantum description of a particle in a central potential, the and operators commute, enabling simultaneous to reveal energy-angular momentum eigenstates. The result traces its origins to John von Neumann's foundational 1932 monograph on , where he proved the simultaneous for operators as part of developing the rigorous mathematical framework for the theory. This work generalized earlier finite-dimensional insights from matrix theory, such as those by Frobenius and Schur, to the operator setting essential for quantum applications.

Normal Elements in Operator Algebras

In C*-Algebras

In a A, an element a \in A is called normal if it commutes with its , that is, if a^* a = a a^*. This condition generalizes the notion of normal operators on Hilbert spaces to the abstract setting of , where the * preserves the algebraic structure and norm properties. A key property of normal elements is that the C*-subalgebra generated by a and the unit $1 (in the unital case) is commutative and isometrically *-isomorphic to C(\sigma(a)), the C*-algebra of continuous complex-valued functions on the compact spectrum \sigma(a) of a. This isomorphism arises from the continuous functional calculus for normal elements, which maps a continuous function f \in C(\sigma(a)) to an element \tilde{f}(a) \in A such that \tilde{id}(a) = a and the spectrum satisfies \sigma(\tilde{f}(a)) = f(\sigma(a)). In the non-unital case, the subalgebra generated by a and a^* is instead isomorphic to C_0(\sigma(a) \setminus \{0\}), the functions vanishing at 0. The Gelfand transform plays a central role in understanding normal elements within commutative C*-algebras. For a unital commutative A, the Gelfand transform \Gamma: A \to C(\Delta(A)) is an *- onto the continuous functions on the space \Delta(A), which coincides with the of A. Since the generated by a element is commutative, this transform restricts to yield the , associating normal elements with their spectral representations as multiplication operators by continuous functions. Examples of normal elements abound in concrete C*-algebras. In the C*-algebra B(\mathcal{H}) of bounded linear operators on a \mathcal{H}, the normal elements are precisely the usual normal operators, recovering the classical theory. Moreover, in any commutative , every element is normal, as all elements commute with one another and thus with their adjoints.

In Von Neumann Algebras

In von Neumann algebras, a normal element is defined analogously to the case as an operator a \in M satisfying a a^* = a^* a, where M is a , meaning M is the double commutant of some set of bounded operators on a and is closed in the weak topology (as opposed to the norm topology in C*-algebras). This weak closure ensures that the algebra generated by normal elements inherits additional structural properties tied to the topology. The generated by a element a \in M and the is commutative and, under suitable conditions such as when the generated is maximal abelian, forms a maximal abelian (masa) within M. This commutativity arises directly from the normality condition, allowing the to be represented as L^\infty(\mu) for some via the Gelfand-Naimark theorem adapted to the setting. Such play a key role in decomposing M into factors and analyzing its type classification. For a bounded normal element a \in M, the spectral theorem provides a projection-valued measure E on the Borel subsets of the spectrum \sigma(a) with values in the projections of M, such that a = \int_{\sigma(a)} \lambda \, dE(\lambda), where the integral converges in the weak operator topology, and all spectral projections E(B) for Borel sets B \subseteq \sigma(a) belong to M. This representation enables the , mapping bounded Borel functions f on \sigma(a) to elements f(a) \in M, preserving the . Unbounded normal operators affiliated to a von Neumann algebra M are closed, densely defined operators T on the underlying such that T commutes with every element of the commutant M' (i.e., for every unitary U \in M', U maps the domain of T to itself and U T U^* = T). Equivalently, the resolvent (T - zI)^{-1} \in M for all z \in \rho(T), the of T. Unlike self-adjoint affiliated operators, whose spectrum lies on the real line, affiliated operators can have complex spectra. The extends to provide a E with values in M such that T = \int \lambda \, dE(\lambda), where the integral is over the extended , converging appropriately, and the affiliation condition guarantees consistency with the weak closure of M. This framework is essential for studying spectral invariants like Brown measures in finite algebras.

Unbounded Normal Operators

Definition and Closure Properties

In functional analysis, an unbounded densely defined linear operator N on a Hilbert space H is termed normal if it is closable, its adjoint N^* is densely defined, and the equality N N^* = N^* N holds on the intersection of the domains D(N) \cap D(N^*). This definition extends the notion of normality from bounded operators, where commutativity with the adjoint is unrestricted, but accounts for the domain restrictions inherent to unbounded operators. The closability condition ensures that the graph of N has a closed extension that preserves the operator structure, while the dense domain of N^* guarantees the adjoint is well-defined and the commutativity can be meaningfully interpreted on the common domain. For closed normal operators, the domains of N and N^* coincide, i.e., D(N) = D(N^*). This equality follows from the commutativity condition and the closedness, implying that the graph norms induced by N and N^* are equivalent on the shared domain. Equivalently, closed normality can be characterized by D(N) = D(N^*) and \|N x\| = \|N^* x\| for all x \in D(N), which underscores the balanced behavior of the operator and its adjoint with respect to the norm. A key closure property is that if a normal operator N (in the densely defined sense) is closable, then its \overline{N} is also normal. The \overline{N} inherits the commutativity with its because the closure preserves the domain and the equality on it, ensuring \overline{N} satisfies the closed normal operator criteria. This stability under is essential for extending results from core domains to maximal ones in applications. A example of an unbounded operator is the P = -i \frac{d}{dx} on the L^2(\mathbb{R}), defined initially on the dense domain of smooth functions with compact support (or the ) and extended to its closure on the H^1(\mathbb{R}). Since P is , it commutes with its (P = P^*), making it , and its unboundedness arises from functions whose derivatives grow without bound in L^2 .

Spectral Theorem Extension

The spectral theorem for closed densely defined normal operators on a separable extends the bounded case by providing a multiplicative representation that accounts for the operator's possible unboundedness. Specifically, every such operator N is unitarily equivalent to by the \lambda on the space L^2(\sigma(N), \mu), where \sigma(N) \subseteq \mathbb{C} is the of N and \mu is a positive on \sigma(N). The domain of this operator consists of those f \in L^2(\sigma(N), \mu) for which \int_{\sigma(N)} |\lambda f(\lambda)|^2 \, d\mu(\lambda) < \infty. This representation ensures that N acts as Nf = \lambda f on its domain, preserving the structure while handling the lack of boundedness through the weighted integrability condition. The proof builds on the bounded spectral theorem by leveraging the resolvent family. For \zeta \notin \sigma(N), the resolvent R(\zeta) = (\zeta I - N)^{-1} is a bounded normal operator, to which the bounded applies directly, yielding a spectral measure E for N via limits of resolvent expressions like Stone's formula. To extend to the full representation, the continuous functional algebra generated by the resolvents is approximated using the Stone-Weierstrass theorem on the compactified spectrum, allowing uniform approximation of continuous functions and construction of the measure \mu supported on \sigma(N). The unitary equivalence then follows from the multiplication form for the bounded case, adjusted for the domain of N. This approach ensures the theorem holds for closed normal operators without requiring self-adjointness. Associated with this representation is a : for any Borel measurable function f: \sigma(N) \to \mathbb{C}, the operator f(N) is defined via the spectral integral f(N) = \int_{\sigma(N)} f(\lambda) \, dE(\lambda), where E is the unique spectral measure such that N = \int_{\sigma(N)} \lambda \, dE(\lambda) in the strong sense, and the domain of f(N) comprises vectors \xi \in H for which \int_{\sigma(N)} |f(\lambda)|^2 \, d\langle E(\lambda) \xi, \xi \rangle < \infty. This calculus preserves normality, with \sigma(f(N)) = f(\sigma(N)), and enables the definition of functions like exponentials or powers of N essential for evolution equations. In quantum mechanics, this theorem underpins the treatment of observables like the position operator Q and momentum operator P on L^2(\mathbb{R}), which are unbounded self-adjoint (hence normal) operators with continuous spectra \sigma(Q) = \sigma(P) = \mathbb{R}. The spectral representation allows expectation values \langle \psi | Q \psi \rangle = \int_{\mathbb{R}} x |\psi(x)|^2 \, dx (in the position representation) and facilitates uncertainty relations via the multiplication form.

Generalizations and Extensions

Hyponormal Operators

Hyponormal operators extend the class of by replacing the equality N^* N = N N^* with an inequality in the sense of positive semi-definiteness. For a bounded linear N on a , N is hyponormal if N^* N \geq N N^*. This condition is equivalent to \|[N, N^*] x\| \geq 0 for all x in the space, where [N, N^*] = N^* N - N N^* is the self-commutator. For unbounded densely defined operators, the notion is generalized to account for domains. An unbounded operator N is hyponormal if D(N) \subseteq D(N^*) and \|N^* x\| \leq \|N x\| for all x \in D(N). Under these conditions, the self-commutator extends to a positive operator on the appropriate domain. Normal operators satisfy the definition of hyponormal operators as the special case where equality holds in the inequality. Self-adjoint operators are hyponormal, since they are and thus satisfy the stricter equality. Subnormal operators, which are restrictions of operators to subspaces, are also hyponormal, as the inequality inherits from the normal extension. For certain classes of hyponormal operators, such as those with trace-class self-commutators, the lies in the closed right half-plane \{ z \in \mathbb{C} : \operatorname{Re} z \geq 0 \}. A representative example of a hyponormal that is not is the unilateral weighted shift on \ell^2(\mathbb{N}) with weights \{\alpha_n\}_{n=0}^\infty satisfying \alpha_n^2 \geq \alpha_{n-1}^2 for all n \geq 1, with the convention \alpha_{-1} = 0. For instance, the standard unilateral shift (with all \alpha_n = 1) is hyponormal but not , as its self-commutator is the rank-one onto the first basis vector. More generally, weighted shifts with strictly increasing weights, such as \alpha_n = \sqrt{n+1}, yield hyponormal operators whose self-commutators have infinite rank, distinguishing them further from s. Putnam's inequality provides a key norm estimate for functions of hyponormal operators. For a hyponormal operator N and an analytic function p on a neighborhood of the spectrum, \|p(N)\| \leq \sup \{ |p(z)| : \operatorname{Re} z \geq 0 \}. This bound reflects the containment of the numerical range in the right half-plane for relevant classes and controls the growth of functional calculus applications.

Quasi-Normal Operators

A bounded linear operator T on a complex Hilbert space H is said to be quasi-normal if it commutes with T^* T, that is,
T (T^* T) = (T^* T) T
or equivalently,
T T^* T = T^* T T.
This condition represents a weakening of the normality condition T T^* = T^* T, as the commutator [T, T^*] need not vanish but satisfies [T, T^*] T = 0. The concept was first introduced by Arlen Brown in 1953, who showed that every quasi-normal operator admits a canonical decomposition T = B \oplus C, where C is normal on the subspace \ker((T^* T)^\infty) (the normal kernel) and B is unitarily equivalent to an operator of the form V P_0 on a suitable subspace, with V an isometry and P_0 a positive operator.
Quasi-normal operators possess several properties analogous to those of normal operators but with notable distinctions. In particular, every quasi-normal operator is hyponormal, meaning T^* T - T T^* \geq 0 in the sense of positive semi-definiteness, since the decomposition into a normal part and a pure part (unitarily equivalent to a certain weighted shift) ensures the hyponormality condition holds. Moreover, quasi-normal operators are subnormal, admitting a normal extension on a larger . Regarding the spectrum, quasi-normal operators share with normal operators the property that \sigma(T) = \sigma(T^*), as follows from their subnormality and the fact that the spectrum coincides with that of the minimal normal extension; however, the point spectrum \sigma_p(T) may differ from \sigma_p(T^*), unlike in the normal case. For instance, the approximate point spectrum of a pure quasi-normal operator aligns with the boundary of its , but isolated eigenvalues in the point spectrum of T need not be eigenvalues of T^*. Examples of quasi-normal operators include all normal operators, as they trivially satisfy the commutation relation, and the unilateral S on the H^2(\mathbb{D}), which satisfies S S^* S = S^* S S = S since S^* S = I. More generally, analytic Toeplitz operators T_f with analytic symbols f (powers of the shift) are quasi-normal, as are certain compressions of normal operators to invariant subspaces where the commutation holds. However, not all subnormal operators are quasi-normal; for example, some weighted shifts that are subnormal fail the stronger commutation condition. The subnormality of quasi-normal operators implies that they can be "lifted" to normal operators via extensions: specifically, there exists a larger K \supseteq H and a normal operator N on K such that H is for N and N|_H = T. This extension is unique up to unitary equivalence in the minimal case, and the measure of N determines key of T, such as the multiplicity function. This lifting property distinguishes quasi-normal operators among hyponormals, providing a bridge to the full for normals while allowing for non-trivial defect structures in the pure part.

References

  1. [1]
    [PDF] Topics in Hilbert Spaces, Spectral Theory, and Harmonic Analysis
    Oct 1, 2019 · We say an operator is normal if it commutes with its adjoint, and we say an operator is self-adjoint if it is equal to its adjoint. Theorem ...
  2. [2]
    [PDF] some notes on the spectral theorem
    The linear transfor- mation is normal if T∗T equals TT∗, i.e., T commutes with its adjoint. ... For a normal operator T, the kernel of T∗ equals the kernel of T.
  3. [3]
  4. [4]
    [PDF] Bounded Linear Operators on a Hilbert Space - UC Davis Math
    An operator T : H→H is said to be normal if it commutes with its adjoint, meaning that TT∗ = T∗T. Both self-adjoint and unitary operators are normal. An.Missing: source | Show results with:source
  5. [5]
    [PDF] JOHN VON NEUMANN - National Academy of Sciences
    The quantum theory led von Neumann naturally to a re-examina- tion of the theory of operators in Hilbert space and, while still in. Berlin, he published in ...<|control11|><|separator|>
  6. [6]
    [PDF] Chapter 9: The Spectrum of Bounded Linear Operators
    For example, an operator may have a continuous spectrum in addition to, or instead of, a point spectrum of eigenvalues.
  7. [7]
    [PDF] Operators on Hilbert space
    The operator Mf is called the multiplication operator defined by the function f. ... Theorem 4.12 is easily generalisable to so-called normal operators, i.e. ...
  8. [8]
    [PDF] Self-Adjoint and Normal Operators, Part 3 - Linear Algebra Done Right
    Definition: normal. An operator on an inner product space is called normal if it commutes with its adjoint. In other words, T ∈ L(V) is normal if.
  9. [9]
    [PDF] Class notes, Functional Analysis 7212 - OSU Math
    Apr 1, 2019 · functional calculus with normal operators ... Under a unitary transformation using the spectral theorem, A becomes a multiplication operator.
  10. [10]
    [PDF] Normal operator
    Mar 12, 2013 · Eigenvectors of a normal operator corresponding to different eigenvalues are orthogonal, and it stabilizes orthogonal complements to its ...
  11. [11]
    [PDF] lecture 28: adjoints and normal operators
    If T is normal, then ker(T) = ker(Tr) for all r ≥ 1. Proof. Let S = T∗T. This is a self-adjoint linear operator. We can see that ker(S) = ker(Sr) for all ...<|control11|><|separator|>
  12. [12]
  13. [13]
    [PDF] Spectral Decomposition of Quantum-Mechanical Operators
    For a normal operator T, the spectrum of T is entirely approximate point spectrum. Proof. Let λ be in σ(T) for a normal operator T, and consider the ways in.
  14. [14]
    [PDF] An introduction to C*-algebras - Gábor Szabó
    Let x ∈ A be a normal element in a C∗-algebra. Let Ax = C∗(x,1) be the commutative C∗-subalgebra generated by x. Then. ˆ. Ax. ∼. = σ(x) by observing that ...
  15. [15]
    [PDF] chapter x the spectral theorem of gelfand definition
    An element x that commutes with its adjoint x∗ is called a normal element of A. ... sertion in the Spectral Theorem for a normal operator. (c) Apply parts ...
  16. [16]
    [PDF] The Spectral Theorem for normal linear maps 1 Self-adjoint or ...
    Mar 14, 2007 · Also, any self-adjoint operator T is normal. We now give a different characterization for normal operators in terms of norms.
  17. [17]
    Proof of the Spectral Decomposition Theorem in Finite Dimension ...
    PDF | Using second Principle of Mathematical Induction, Spectral Decomposition Theorem is proved. It is applicable for normal matrices, which are.
  18. [18]
    [PDF] Contents 4 Eigenvalues, Diagonalization, and the Jordan Canonical ...
    ◦ The Jordan canonical form therefore serves as an approximate diagonalization for non-diagonalizable ... normal operator, meaning that T∗T = TT∗.
  19. [19]
    [PDF] Two classical theorems on commuting matrices
    (2) An arbitrary set of commuting normal matrices may be simultaneously brought to diagonal form by a unitary similarity.
  20. [20]
    [PDF] Simultaneous commutativity of operators - Keith Conrad
    If each Ai is diagonalizable on V then they are simultaneously diagonal- izable. This choice of basis for Eλ is not made by Ar, but by the other operators ...
  21. [21]
    C*-Algebras and Operator Theory - ScienceDirect.com
    This book is a first or second-year graduate course in operator theory, covering C*-algebras and Hilbert space operators, and is important in math and physics.
  22. [22]
    [PDF] XI.3 Continuous functional calculus in C*-algebras
    Let A be a C∗-algebra with a unit e and let x ∈ A be a normal element. Let B be the closed subalgebra of A generated by the set {e, x, x∗}. Then: • B is ...
  23. [23]
    [PDF] Notes on von Neumann algebras
    Apr 5, 2013 · Let A be a C∗-algebra, and x ∈ A a normal element, then if f ∈ C ... formly by the span of its spectral projections it then follows that.
  24. [24]
    [PDF] Von Neumann Algebras. - Berkeley Math
    Oct 1, 2009 · An operator u ∈ B(H) is called a partial isometry if u∗u is a projection. The last three definitions extend to bounded linear operators between ...
  25. [25]
    [PDF] Operator Algebras and Unbounded Self-Adjoint Operators
    Jun 16, 2015 · At the end we will connect unbounded operators to von Neumann algebras by the method of affiliated von Neumann algebras. The. 3. Page 4. main ...
  26. [26]
    [PDF] Brown Measures of Unbounded Operators Affiliated with a Finite von ...
    In this paper we generalize Brown's spectral distribution measure to a large class of unbounded operators affiliated with a finite von Neumann algebra.
  27. [27]
    [PDF] Unbounded linear operators
    An operator N in a Hilbert space H is called normal if N is closed and densely defined, and NN∗ = N∗N. Show that N is normal if and only if N is a densely ...
  28. [28]
    [PDF] XIII.4 Normal unbounded operators
    (c) If S ⊃ T is normal, then S = T. Theorem 21 (spectral decomposition of an unbounded normal operator). If. T is a normal operator on H, then there exists ...
  29. [29]
    A note on the definition of an unbounded normal operator
    The inner product space (D(T), ·T ) is a Hilbert space if and only if T is closed. Proof. First, assume T is closed. Let (xn)n≥1 ⊆ D(T) be a Cauchy sequence ...
  30. [30]
    [PDF] 1 Bounded and unbounded operators - OSU Math
    As we have seen, the spectrum of an unbounded closed operator can be any closed set (that it is necessarily closed we will see shortly), including C and. ∅.
  31. [31]
    [PDF] Let's solve Schrödinger's equation for a particle on R with different ...
    Just like with bounded normal operators, unbounded normal operators can be diagonalized. ... It will turn out that the momentum operator is self-adjoint.
  32. [32]
    [PDF] The Spectral Theorem for Unbounded Operators.
    Nov 29, 2001 · Many important operators in Hilbert space that arise in physics and math- ematics are "unbounded". For example the operator D = zi dd}on L2(R).
  33. [33]
    The spectral theorem for unbounded normal operators - MSP
    The spectral theorem for unbounded normal operators states that T = \xE(dX), where E is the unique spectral measure. T is a normal, but not necessarily bounded ...
  34. [34]
    [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.
  35. [35]
    Hyponormal operators. - Project Euclid
    1962 Hyponormal operators. Joseph G. Stampfli · DOWNLOAD PDF + SAVE TO MY LIBRARY. Pacific J. Math. 12(4): 1453-1458 (1962).
  36. [36]
    Lectures on hyponormal operators, by M. Martin and M. Putinar.
    A hyponormal operator is a bounded operator T on a Hubert space %f such that T*T> TT*. This rather innocent definition was introduced by Paul Halmos [11] in ...
  37. [37]
    spectral pictures of hyponormal bilateral operator weighted shifts
    This note provides a complete description of the spectral picture of a hyponormal bilateral operator weighted shift of finite multiplicity. If the weights are ...
  38. [38]
    None
    ### Definition of Quasi-Normal Operators
  39. [39]
    The Local Spectra of Pure Quasinormal Operators - ScienceDirect
    The local spectra of pure quasinormal operators are computed and some results concerning the structure of the local spectra of operators satisfying ...
  40. [40]
    Subnormal and quasinormal Toeplitz operators with matrix-valued ...
    Apr 1, 2014 · A Toeplitz operator T φ (with symbol φ ∈ L ∞ ≡ L ∞ ( T ) ) is defined by the expression T φ f : = P ( φ f ) for each f ∈ H 2 , where P is the ...