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.[1][2] This property ensures that normal operators preserve the norm of vectors under application, meaning \| T x \| = \| T^* x \| for all x in the space.[1] 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.[2][1] The significance of normal operators lies in the spectral theorem, which states that any normal operator on a finite-dimensional Hilbert space admits an orthonormal basis of eigenvectors, allowing it to be diagonalized in an appropriate basis.[1][2] In infinite-dimensional settings, the theorem extends to a spectral decomposition involving projections onto eigenspaces or more general spectral measures, facilitating the analysis of eigenvalues and eigenvectors even when they are not discrete.[1] Additionally, the eigenspaces of a normal operator for T coincide with those for T^*, and the kernels of T and T^* are equal.[2] These features make normal operators a cornerstone for studying operator algebras and applications in physics, such as observables in Hilbert space formulations of quantum theory.[1]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.[3] 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\|.[4] The adjoint operator N^* of a bounded linear operator N on H is uniquely defined by the relation \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\|.[4] A bounded linear operator N on a complex Hilbert space H is called normal if it commutes with its adjoint, that is, N N^* = N^* N.[4] Self-adjoint operators, which satisfy N = N^*, form a special case of normal operators, as they trivially commute with themselves.[4] The concept of normal operators was introduced by John von Neumann in the early 1930s as part of his foundational work on the mathematical structure of quantum mechanics, particularly in developing the spectral theory for operators on Hilbert spaces.[5] A key consequence is the spectral theorem, which allows normal operators to be diagonalized in a suitable sense, facilitating their analysis in physical and mathematical applications (detailed in later sections).[6]Basic Examples
A fundamental example of a normal operator arises in the context of multiplication operators on the Hilbert space L^2(\mathbb{R}). For a bounded complex-valued measurable function f, the multiplication operator M_f is defined by (M_f g)(x) = f(x) g(x) for all g \in L^2(\mathbb{R}). The adjoint of M_f is M_{\bar{f}}, the multiplication by the complex conjugate \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 multiplication, making M_f normal.[7] Orthogonal projections provide another straightforward class of normal operators. Let H be a Hilbert space and P the orthogonal projection onto a closed subspace K \subseteq H. Such a projection satisfies P^2 = P and is self-adjoint, meaning P = P^*. Self-adjoint operators are normal because P P^* = P^2 = P = P^* P, illustrating the commutativity condition directly.[8] In sequence spaces, diagonal operators on \ell^2(\mathbb{N}) serve as basic normal examples. Consider the operator D defined on the standard orthonormal basis \{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 adjoint 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 normality.[9] Unitary operators form a significant subclass of normal operators. A bounded linear operator U on a Hilbert space H is unitary if U^* U = U U^* = I, the identity operator. This relation immediately implies U U^* = U^* U, satisfying the normality condition, as the commutator vanishes by definition.[8] To contrast, consider the unilateral shift operator on \ell^2(\mathbb{N}), which fails to be normal. Defined by S e_n = e_{n+1} for the standard basis \{e_n\}, its adjoint 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 normality.[9]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 Hilbert space 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 adjoint. Taking the supremum over all such unit vectors yields \|N\| = \|N^*\|.[10] 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.[11] 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^*.[12] 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.[13] 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.[14]Norm and Spectrum Basics
The spectrum of a bounded normal operator N on a Hilbert space is defined as the set \sigma(N) = \{\lambda \in \mathbb{C} : N - \lambda I \text{ is not invertible}\}.[6] For any bounded linear operator on a complex Banach space, the spectrum is a non-empty compact subset of \mathbb{C}.[6] 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.[15] 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).[15] The C*-algebra generated by a normal operator N and the identity I, denoted C^*(N, I), is commutative.[16] 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.[16] 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.[17] For a normal operator N, Gelfand's formula gives r(N) = \lim_{n \to \infty} \|N^n\|^{1/n}.[17] 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\|.[17] Thus, \sup\{|\lambda| : \lambda \in \sigma(N)\} = \|N\|.[17]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 diagonal matrix. That is, there exists a unitary matrix U such that U^* T U = D, where D is diagonal with entries consisting of the eigenvalues of T.[18][19] 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.[19] 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 orthonormal basis, as unitary transformations preserve inner products.[18][19] 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 structure.[20]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 Hilbert space \mathcal{H}, satisfying N_i N_j = N_j N_i for all i, j = 1, \dots, k. Then there exists a unitary operator U: \mathcal{H} \to \mathcal{H} such that U^* N_i U is diagonal for each i = 1, \dots, k.[21] This simultaneous unitary diagonalization implies the existence of an orthonormal basis of \mathcal{H} consisting of common eigenvectors for all N_i, where the diagonal entries correspond to the joint eigenvalues.[22] 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.[22][21] This theorem finds a key application in quantum mechanics, where self-adjoint operators (a subclass of normals) represent physical observables. Commuting 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 Hamiltonian and angular momentum operators commute, enabling simultaneous diagonalization to reveal energy-angular momentum eigenstates. The result traces its origins to John von Neumann's foundational 1932 monograph on quantum mechanics, where he proved the simultaneous spectral resolution for commuting self-adjoint operators as part of developing the rigorous mathematical framework for the theory.[23] 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.[21]Normal Elements in Operator Algebras
In C*-Algebras
In a C*-algebra A, an element a \in A is called normal if it commutes with its adjoint, that is, if a^* a = a a^*.[24] This condition generalizes the notion of normal operators on Hilbert spaces to the abstract setting of C*-algebras, where the involution * 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.[24] 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)).[25] 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.[25] The Gelfand transform plays a central role in understanding normal elements within commutative C*-algebras. For a unital commutative C*-algebra A, the Gelfand transform \Gamma: A \to C(\Delta(A)) is an isometric *-isomorphism onto the continuous functions on the maximal ideal space \Delta(A), which coincides with the spectrum of A. Since the subalgebra generated by a normal element is commutative, this transform restricts to yield the functional calculus isomorphism, 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 Hilbert space \mathcal{H}, the normal elements are precisely the usual normal operators, recovering the classical theory.[24] Moreover, in any commutative C*-algebra, every element is normal, as all elements commute with one another and thus with their adjoints.[24]In Von Neumann Algebras
In von Neumann algebras, a normal element is defined analogously to the C*-algebra case as an operator a \in M satisfying a a^* = a^* a, where M is a von Neumann algebra, meaning M is the double commutant of some set of bounded operators on a Hilbert space and is closed in the weak operator topology (as opposed to the norm topology in C*-algebras).[26] This weak closure ensures that the algebra generated by normal elements inherits additional structural properties tied to the topology.[27] The von Neumann subalgebra generated by a normal element a \in M and the identity is commutative and, under suitable conditions such as when the generated algebra is maximal abelian, forms a maximal abelian self-adjoint subalgebra (masa) within M.[26] This commutativity arises directly from the normality condition, allowing the subalgebra to be represented as L^\infty(\mu) for some measure space via the Gelfand-Naimark theorem adapted to the von Neumann setting.[27] Such subalgebras 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.[26] This representation enables the Borel functional calculus, mapping bounded Borel functions f on \sigma(a) to elements f(a) \in M, preserving the algebraic structure.[27] Unbounded normal operators affiliated to a von Neumann algebra M are closed, densely defined operators T on the underlying Hilbert space 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 resolvent set of T. Unlike self-adjoint affiliated operators, whose spectrum lies on the real line, affiliated normal operators can have complex spectra. The spectral theorem extends to provide a projection-valued measure E with values in M such that T = \int \lambda \, dE(\lambda), where the integral is over the extended complex plane, converging appropriately, and the affiliation condition guarantees consistency with the weak closure of M.[28] This framework is essential for studying spectral invariants like Brown measures in finite von Neumann algebras.[28]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^*).[29] 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.[30] 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.[31] For closed normal operators, the domains of N and N^* coincide, i.e., D(N) = D(N^*).[30] 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.[31] 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 Hilbert space norm.[30] A key closure property is that if a normal operator N (in the densely defined sense) is closable, then its closure \overline{N} is also normal.[29] The closure \overline{N} inherits the commutativity with its adjoint because the graph closure preserves the domain intersection and the equality on it, ensuring \overline{N} satisfies the closed normal operator criteria.[32] This stability under closure is essential for extending results from core domains to maximal ones in applications. A canonical example of an unbounded normal operator is the momentum operator P = -i \frac{d}{dx} on the Hilbert space L^2(\mathbb{R}), defined initially on the dense domain of smooth functions with compact support (or the Schwartz space) and extended to its self-adjoint closure on the Sobolev space H^1(\mathbb{R}).[33] Since P is self-adjoint, it commutes with its adjoint (P = P^*), making it normal, and its unboundedness arises from functions whose derivatives grow without bound in L^2 norm.[32]Spectral Theorem Extension
The spectral theorem for closed densely defined normal operators on a separable Hilbert space 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 multiplication by the identity function \lambda on the space L^2(\sigma(N), \mu), where \sigma(N) \subseteq \mathbb{C} is the spectrum of N and \mu is a positive Borel measure on \sigma(N). The domain of this multiplication operator consists of those f \in L^2(\sigma(N), \mu) for which \int_{\sigma(N)} |\lambda f(\lambda)|^2 \, d\mu(\lambda) < \infty.[34] This representation ensures that N acts as Nf = \lambda f on its domain, preserving the Hilbert space structure while handling the lack of boundedness through the weighted integrability condition.[35] 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 spectral theorem 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.[34] This approach ensures the theorem holds for closed normal operators without requiring self-adjointness.[36] Associated with this representation is a Borel functional calculus: 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.[36] 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 normal operators by replacing the equality N^* N = N N^* with an inequality in the sense of positive semi-definiteness. For a bounded linear operator N on a Hilbert space, N is hyponormal if N^* N \geq N N^*.[37] 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.[38] 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.[37] Self-adjoint operators are hyponormal, since they are normal and thus satisfy the stricter equality.[38] Subnormal operators, which are restrictions of normal operators to invariant subspaces, are also hyponormal, as the inequality inherits from the normal extension.[38] For certain classes of hyponormal operators, such as those with trace-class self-commutators, the spectrum lies in the closed right half-plane \{ z \in \mathbb{C} : \operatorname{Re} z \geq 0 \}.[38] A representative example of a hyponormal operator that is not normal 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 normal, as its self-commutator is the rank-one projection onto the first basis vector.[39] 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 normals.[39] 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 \}.[38] 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.[40] 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.[40] Moreover, quasi-normal operators are subnormal, admitting a normal extension on a larger Hilbert space. 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 spectrum, but isolated eigenvalues in the point spectrum of T need not be eigenvalues of T^*.[41] Examples of quasi-normal operators include all normal operators, as they trivially satisfy the commutation relation, and the unilateral shift operator S on the Hardy space 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.[42] The subnormality of quasi-normal operators implies that they can be "lifted" to normal operators via extensions: specifically, there exists a larger Hilbert space K \supseteq H and a normal operator N on K such that H is invariant for N and N|_H = T. This extension is unique up to unitary equivalence in the minimal case, and the spectral measure of N determines key invariants of T, such as the multiplicity function. This lifting property distinguishes quasi-normal operators among hyponormals, providing a bridge to the full spectral theorem for normals while allowing for non-trivial defect structures in the pure part.