Fact-checked by Grok 2 weeks ago

Normal matrix

In linear algebra, a normal matrix is a square matrix A over the numbers that commutes with its A^\dagger, satisfying the equation A A^\dagger = A^\dagger A. This defining property distinguishes normal matrices from more general square matrices and ensures they share key algebraic behaviors with simpler forms like diagonal matrices. Normal matrices encompass several important subclasses, including Hermitian matrices (where A = A^\dagger), unitary matrices (where A^\dagger A = I), real symmetric matrices (Hermitian with real entries), and skew-Hermitian matrices (where A^\dagger = -A). These subclasses arise naturally in applications such as , , and , where the normality condition preserves and stability under transformations. Beyond these, normal matrices can be non-Hermitian, as illustrated by the matrix \begin{pmatrix} 1 & -1 \\ 1 & 1 \end{pmatrix}, which satisfies the commutativity but belongs to none of the above categories. The most notable theorem associated with normal matrices is the , which states that a matrix A is normal if and only if it is unitarily diagonalizable: there exists a U such that U^\dagger A U is diagonal, with the diagonal entries being the eigenvalues of A. This diagonalization implies that eigenvectors corresponding to distinct eigenvalues are orthogonal, facilitating computations like eigenvalue decomposition in finite-dimensional Hilbert spaces. Consequently, normal matrices play a central role in the study of linear operators, extending to infinite-dimensional settings via normal operators in .

Definition and Fundamentals

Definition

In linear algebra, a normal matrix is defined as a that commutes with its . Specifically, for an n \times n A \in \mathbb{C}^{n \times n}, A is if A A^\dagger = A^\dagger A, where A^\dagger denotes the conjugate transpose (adjoint) of A, given by A^\dagger = \overline{A}^T. Simple examples of normal matrices include the I, for which I I^\dagger = I^\dagger I = I, and any D = \operatorname{diag}(d_1, \dots, d_n) with complex entries d_i \in \mathbb{C}, since D^\dagger = \operatorname{diag}(\overline{d_1}, \dots, \overline{d_n}) and both products D D^\dagger and D^\dagger D yield the diagonal matrix with entries |d_i|^2. Normal matrices generalize (Hermitian) operators while encompassing a broader class, including unitaries that preserve the Euclidean norm, making them essential in applications such as , where they represent operators that can be diagonalized in orthonormal bases to model observables and evolutions. A distinguishing feature is that every normal matrix is unitarily diagonalizable over the complex numbers. Although the definition is formulated for complex matrices to leverage the inner product structure of \mathbb{C}^n, an analogous notion for real matrices involves commuting with the transpose A^T instead, but such real normal matrices do not necessarily share all spectral properties with their complex counterparts.

Notation and Conventions

In discussions of normal matrices, the conjugate transpose (also known as the adjoint) of a matrix A is standardly denoted by A^\dagger, obtained by transposing A and taking the complex conjugate of each entry. The elementwise complex conjugate of A, without transposition, is denoted by A^*. The operator norm of a matrix A, which measures its maximum amplification of vectors, is denoted by \|A\|. A normal matrix must be square, as the definition involves commutativity between the matrix and its , a relation that presupposes equal dimensions for rows and columns. Such matrices typically have entries and act on finite-dimensional vector spaces, distinguishing them from rectangular matrices where the relation does not apply in the same way. Unless otherwise specified, this article assumes matrices over the field of complex numbers, reflecting the standard finite-dimensional setting for normal matrices. Extensions to other skew fields, such as the s, are noted where relevant, adapting the via the quaternion conjugate. The term "normal matrix" was coined by in 1932, originating in his foundational work on in Hilbert spaces.

Properties and Consequences

Basic Properties

A normal matrix A satisfies \|Ax\| = \|A^\dagger x\| for every x, where \|\cdot\| denotes the and A^\dagger is the conjugate transpose of A. This equality follows directly from the definition A A^\dagger = A^\dagger A, since \|Ax\|^2 = x^\dagger A^\dagger A x = x^\dagger A A^\dagger x = \|A^\dagger x\|^2. This property implies that A acts as an isometry between the range of A^\dagger and the range of A. The class of normal matrices is closed under unitary similarity: if A is normal and U is a , then B = U A U^\dagger is also normal. To see this, note that B^\dagger = U A^\dagger U^\dagger, so B B^\dagger = U A U^\dagger U A^\dagger U^\dagger = U A A^\dagger U^\dagger and B^\dagger B = U A^\dagger A U^\dagger, with the normality of A ensuring the equality. For a normal matrix A, the trace satisfies \operatorname{Tr}(A A^\dagger) = \operatorname{Tr}(A^\dagger A) = \sum_i |\lambda_i|^2, where \lambda_i are the eigenvalues of A. This follows from the cyclic property of the trace, which equates \operatorname{Tr}(A A^\dagger) and \operatorname{Tr}(A^\dagger A), and the unitary diagonalizability of normal matrices, which aligns the Frobenius norm with the sum of squared eigenvalue moduli. A concrete example of normal matrices that are not Hermitian is provided by scalar multiples of the . Consider \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}; then A = i \sigma_x = \begin{pmatrix} 0 & i \\ i & 0 \end{pmatrix}. The adjoint is A^\dagger = -i \sigma_x, and A A^\dagger = (i \sigma_x)(-i \sigma_x) = \sigma_x^2 = I, while A^\dagger A = (-i \sigma_x)(i \sigma_x) = \sigma_x^2 = I, confirming normality. Similar computations hold for i \sigma_y and i \sigma_z, yielding A A^\dagger = A^\dagger A = I in each case, with \operatorname{Tr}(A A^\dagger) = 2.

Spectral Properties and Diagonalization

A fundamental consequence of normality is the spectral theorem, which asserts that every normal matrix over the complex numbers is unitarily diagonalizable. Specifically, for a normal matrix A \in \mathbb{C}^{n \times n}, there exists a unitary matrix U and a D such that A = U D U^\dagger, where U^\dagger denotes the of U, and the diagonal entries of D are the eigenvalues of A. This decomposition implies that normal matrices possess an orthonormal basis of eigenvectors, facilitating analysis in the eigenbasis. The eigenvalues of a normal matrix are complex numbers, and a key property is that the spectral radius \rho(A), defined as the maximum modulus of its eigenvalues, equals the operator norm induced by the Euclidean vector norm, \|A\|_2 = \rho(A). This equality holds because the unitary diagonalization preserves the 2-norm, bounding \|Ax\|_2 \leq \rho(A) \|x\|_2 for \|x\|_2 = 1, with equality achieved for an eigenvector corresponding to the eigenvalue of largest modulus. A standard proof of the spectral theorem relies on Schur triangularization and the structure imposed by normality. By the Schur theorem, every matrix A admits a unitary U such that U^\dagger A U = T, where T is upper triangular with the eigenvalues of A on its diagonal. For normal A, compute A^\dagger A = A A^\dagger, leading to T^\dagger T = T T^\dagger; since T is upper triangular, this commutativity forces all superdiagonal entries of T to vanish, rendering T diagonal. Moreover, normal matrices have orthogonal eigenspaces: if v_1, v_2 are eigenvectors for distinct eigenvalues \lambda_1 \neq \lambda_2, then \langle v_1, v_2 \rangle = 0, as (\lambda_1 - \overline{\lambda_2}) \langle v_1, v_2 \rangle = \langle A v_1, v_2 \rangle - \langle v_1, A^\dagger v_2 \rangle = 0. Induction on the dimension, reducing to invariant subspaces, completes the diagonalization. In , the unitary diagonalizability of matrices corresponds to that of normal operators on Hilbert spaces, enabling efficient simulation of their actions, such as or measurements, by transforming to the eigenbasis where operations reduce to phase shifts or projections.

Characterizations

Algebraic Characterizations

A matrix A \in M_n(\mathbb{C}) is if and only if the [A, A^\dagger] = AA^\dagger - A^\dagger A = 0, where A^\dagger denotes the conjugate transpose of A. This condition is equivalent to the defining property AA^\dagger = A^\dagger A. Another algebraic characterization is that A is normal if and only if p(A) is normal for every polynomial p with complex coefficients. Equivalently, A is normal if and only if A^\dagger = p(A) for some polynomial p. In this case, the minimal polynomial of A^\dagger is obtained from the minimal polynomial of A by conjugating the coefficients. The Fuglede–Putnam theorem provides a related algebraic property: if A and B are normal operators on a Hilbert space and AB = BA, then A B^\dagger = B^\dagger A.

Analytic and Geometric Characterizations

A matrix A \in \mathbb{C}^{n \times n} is normal if and only if the matrices AA^\dagger and A^\dagger A have the same eigenvalues, counting algebraic multiplicities. These shared eigenvalues are the squares of the singular values of A, and for a normal matrix, they equal |\lambda_i|^2, where \lambda_i are the eigenvalues of A. The numerical range of a A, defined as W(A) = \left\{ \frac{x^\dagger A x}{x^\dagger x} \;\middle|\; x \in \mathbb{C}^n, x \neq 0 \right\}, provides a geometric characterization of . For a normal A, W(A) is the convex hull of its eigenvalues \{\lambda_1, \dots, \lambda_n\}. Conversely, if the numerical range of A equals the convex hull of its eigenvalues, then A is normal. This property arises from the unitary diagonalizability of normal matrices, which ensures that the Rayleigh quotients align precisely with the spectral points. The numerical range itself is always a convex set in the complex plane, but its equality to the eigenvalue convex hull distinguishes normal matrices geometrically. Singular values offer another analytic characterization tied to normality. For any normal matrix A, the singular values \sigma_i(A) are precisely the absolute values of its eigenvalues, \sigma_i(A) = |\lambda_i|, ordered non-increasingly. Moreover, A is normal if and only if its left and right singular vectors can be chosen to form orthonormal bases simultaneously, reflecting the alignment of the with the eigendecomposition via a common . In , the of an invertible normal matrix A simplifies significantly. The two-norm \kappa_2(A) = \|A\|_2 \|A^{-1}\|_2 equals the ratio of the largest to the smallest absolute eigenvalue, \kappa_2(A) = \max_i |\lambda_i| / \min_i |\lambda_i|. This follows from the equality of singular values and eigenvalue moduli, making normal matrices particularly amenable to stable computations when their eigenvalues are well-separated.

Special Cases

Hermitian Matrices

A Hermitian matrix is a square matrix over the complex numbers that equals its own , denoted A = A^\dagger, where A^\dagger = \overline{A}^T and the bar indicates complex conjugation. This self-adjoint property ensures that every is , as A A^\dagger = A^2 = A^\dagger A. Over the real numbers, coincide with symmetric matrices, where A^T = A. A key spectral property of Hermitian matrices is that all their eigenvalues are real numbers. Moreover, the eigenvectors corresponding to distinct eigenvalues are orthogonal with respect to the standard Hermitian inner product, and a complete orthonormal basis of eigenvectors exists for the entire space. This orthogonality facilitates the spectral decomposition A = U \Lambda U^\dagger, where U is unitary and \Lambda is diagonal with real entries. A Hermitian matrix A is positive semidefinite if \mathbf{x}^\dagger A \mathbf{x} \geq 0 for all nonzero complex vectors \mathbf{x}. Equivalently, all eigenvalues of A are nonnegative. Positive semidefiniteness is a cornerstone for matrices representing quadratic forms in optimization and physics. Real symmetric matrices serve as Hermitian examples in , while covariance matrices in statistics are symmetric , capturing data variability. The diagonalization of Hermitian covariance matrices underpins (), where orthogonal eigenvectors identify directions of maximum variance for .

Skew-Hermitian Matrices

A is a square over the complex numbers satisfying A^\dagger = -A. This property ensures that every skew-Hermitian matrix is , as A A^\dagger = A (-A) = -A^2 = (-A) A = A^\dagger A. Over the real numbers, skew-Hermitian matrices coincide with skew-symmetric matrices, where A^T = -A. A key spectral property of skew-Hermitian matrices is that all their eigenvalues are purely imaginary (including zero). Moreover, the eigenvectors corresponding to distinct eigenvalues are orthogonal with respect to the standard Hermitian inner product, and a complete orthonormal basis of eigenvectors exists for the entire space. This orthogonality facilitates the spectral decomposition A = U \Lambda U^\dagger, where U is unitary and \Lambda is diagonal with purely imaginary entries. Skew-Hermitian matrices arise in the study of Lie algebras and infinitesimal generators of unitary groups, with applications in quantum mechanics and differential equations.

Unitary Matrices

A unitary matrix is a square complex matrix U satisfying U^\dagger = U^{-1}, where U^\dagger denotes the conjugate transpose (Hermitian adjoint) of U. This condition implies U U^\dagger = U^\dagger U = I, the identity matrix, so every unitary matrix is normal. Unitary matrices preserve the Hermitian inner product: for vectors v, w \in \mathbb{C}^n, \langle U v, U w \rangle = \langle v, w \rangle. Consequently, the columns (and rows) of U form an for \mathbb{C}^n with respect to the Hermitian inner product. The eigenvalues of a all lie on the unit circle in the , i.e., have 1. More precisely, a normal matrix is if and only if all its eigenvalues have 1. Real unitary matrices are precisely the orthogonal matrices, which include rotation matrices; for example, the 2D rotation matrix \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix} is unitary for any real \theta. The normalized discrete Fourier transform matrix, with entries F_{jk} = n^{-1/2} \omega^{jk} where \omega = e^{2\pi i / n}, is also unitary. In quantum computing, the Hadamard gate H = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix} serves as a fundamental unitary matrix that creates superpositions.

Extensions and Analogies

Real and Quaternionic Normal Matrices

A real matrix A is defined to be normal if it commutes with its transpose, that is, A A^T = A^T A. This condition ensures that A preserves the Euclidean norm in a certain sense, analogous to the complex case with the Hermitian adjoint. However, unlike complex normal matrices, which are unitarily diagonalizable over \mathbb{C}, real normal matrices are not necessarily orthogonally diagonalizable over \mathbb{R}; orthogonal diagonalization occurs precisely when the matrix is symmetric. The of a real normal matrix reflects the presence of complex eigenvalues, which must occur in conjugate pairs. Specifically, every real normal matrix is orthogonally similar to a block-diagonal matrix whose diagonal blocks are either 1×1 blocks corresponding to real eigenvalues or blocks representing rotation-scaling matrices for eigenvalue pairs. For example, a block might take the form \begin{pmatrix} a & -b \\ b & a \end{pmatrix} for eigenvalues a \pm bi with b \neq 0, combining scaling by a and rotation by \theta = \tan^{-1}(b/a). This structure arises from the real Schur decomposition adapted to the normality condition. The study of real normal forms traces back to Issai Schur's 1909 work on matrix representations, which established foundational results for real matrices predating John von Neumann's 1933 generalization of and to the complex domain. Real normal matrices can be embedded into the complex setting by viewing them as complex matrices, where they retain with respect to the . Quaternionic normal matrices extend the concept to the division algebra of quaternions \mathbb{H}, where a matrix A \in M_n(\mathbb{H}) is normal if A A^\dagger = A^\dagger A, with A^\dagger denoting the conjugate transpose (entrywise quaternion conjugation followed by transpose). Quaternion conjugation involves negating the vector part of each entry, preserving the normed division ring structure. These matrices share spectral properties with their complex counterparts but operate in a non-commutative setting, leading to richer invariant subspace decompositions. Quaternionic normal matrices find significant applications in modeling 3D rotations, as unit quaternionic matrices represent elements of SU(2), the double cover of the rotation group SO(3). For instance, unitary quaternionic normal matrices (satisfying A^\dagger A = I) parameterize rotations in and , leveraging the between SU(2) and the unit quaternions. This framework avoids singularities in Euler angle representations and facilitates efficient composition of rotations.

Normal Operators in Hilbert Spaces

In the context of infinite-dimensional Hilbert spaces, the notion of a normal matrix extends to normal operators, providing a bridge between finite-dimensional linear algebra and abstract operator theory. A bounded linear operator T on a complex Hilbert space \mathcal{H} is defined to be normal if it commutes with its adjoint operator T^*, satisfying T T^* = T^* T. This condition ensures that the operator preserves the orthogonality of eigenspaces in a manner analogous to matrices. When \mathcal{H} is finite-dimensional, every normal operator corresponds exactly to a normal matrix under the choice of an orthonormal basis, recovering the classical case where the matrix satisfies A A^* = A^* A. The for bounded operators, originally proved by in 1932, states that any such operator on a separable admits a : there exists a spectral measure E on the Borel subsets of the such that T = \int_{\sigma(T)} \lambda \, dE(\lambda), where \sigma(T) is the of T. This representation allows T to be unitarily equivalent to multiplication by the identity function on L^2(\mu) for some measure \mu supported on \sigma(T), mirroring the diagonalization of matrices but using a continuous spectral measure instead of discrete eigenvalues. The theorem enables the development of a for operators, where functions of T are defined via f(T) = \int f(\lambda) \, dE(\lambda), facilitating computations and proofs in . For unbounded operators, the definition generalizes to densely defined, closed operators T whose domain equals that of T^*, with T T^* = T^* T holding on this common domain; such operators are normal if they are closed and satisfy the commutation relation where defined. The spectral theorem extends to these cases, provided the operator is normal in this sense, yielding a similar integral representation but now over the extended , often requiring careful handling of the resolvent to ensure domain issues are resolved. In quantum mechanics, prominent examples are the Q and P on L^2(\mathbb{R}), both unbounded self-adjoint (and thus normal) operators, though they fail to commute ([Q, P] = i \hbar I), preventing joint normality but allowing individual spectral decompositions via transforms. Normal operators play a central role in through their spectral decompositions, which underpin the study of C^*-algebras and von Neumann algebras generated by such operators. In applications to partial differential equations (PDEs), the intersects with : for instance, the Laplacian operator -\Delta on L^2(\mathbb{R}^n), which is and normal, is diagonalized by the , converting elliptic or parabolic PDEs into algebraic equations in the spectral ( for efficient solution via integration. This approach, rooted in the multiplier property of normal operators under unitary transformations like the , has broad utility in solving evolution equations and boundary value problems.

References

  1. [1]
    SCLA Normal Matrices - A First Course in Linear Algebra
    The square matrix A A is normal if A∗A=AA∗. ... So a normal matrix commutes with its adjoint. Part of the beauty of this definition is that it includes many other ...Missing: authoritative source
  2. [2]
    [PDF] Unit 17: Spectral theorem
    A real or complex. matrix is called normal if A∗A = AA∗. Examples of normal matrices are symmetric or anti-symmetric matrices. Normal matrices appear often in ...
  3. [3]
    [PDF] Lecture 3.26. Hermitian, unitary and normal matrices - Purdue Math
    A matrix is called normal if it satisfies. A∗A = AA∗. Evidently Hermitian and unitary matrices are normal. Exercise: give an example of a matrix which is ...Missing: authoritative source
  4. [4]
    Normal matrices - ScienceDirect.com
    In hopes that it will be useful to a wide audience, a long list of conditions on an n-by-n complex matrix A, equivalent to its being normal, is presented.Missing: definition | Show results with:definition
  5. [5]
    Matrix Analysis - Cambridge University Press & Assessment
    Roger A. Horn, The Johns Hopkins University, Charles R. Johnson, College of William and Mary, Virginia. Publisher: Cambridge University Press.
  6. [6]
    Normal matrices - Ximera - The Ohio State University
    One can check that this last condition implies must be a diagonal matrix (in other words, the only triangular matrix which is also normal is a diagonal one.Missing: authoritative source
  7. [7]
    conjugate transpose - PlanetMath.org
    Mar 22, 2013 · The conjugate transpose A* of a complex matrix A is A* = ¯AT, where ¯A is the complex conjugate of A, and AT is the transpose of A.
  8. [8]
    Conjugate transpose - StatLect
    Conjugate transpose is the result of transposing and taking the complex conjugate of a matrix, denoted as $A^{st}$ where the overline denotes complex ...Definition · Properties · Hermitian matrix
  9. [9]
    Operator Norm -- from Wolfram MathWorld
    When T is given by a matrix, say T(v)=Av , then ||T|| is the square root of the largest eigenvalue of the symmetric matrix A^(T)A , all of whose eigenvalues ...
  10. [10]
    Normal Matrix -- from Wolfram MathWorld
    The normal matrices are the matrices which are unitarily diagonalizable, i.e., is a normal matrix iff there exists a unitary matrix such that.Missing: authoritative source
  11. [11]
    Square Complex Matrix - an overview | ScienceDirect Topics
    A square complex matrix is defined as a matrix with an equal number of rows and columns, which plays a crucial role in modeling complex processes, ...
  12. [12]
    [PDF] 8.7 Complex Matrices
    Jul 8, 2020 · Hence every square complex matrix A has a (complex) eigenvalue. ... To this end, an n×n complex matrix N is called normal if NNH = NHN. It ...
  13. [13]
    Quaternions and matrices of quaternions - ScienceDirect.com
    We give a brief survey on quaternions and matrices of quaternions, present new proofs for certain known results, and discuss the quaternionic analogues of ...
  14. [14]
    theory of operators part i. single operators - f, j. murray - Project Euclid
    The work of von Neumann on Operator Theory as distinct from the later work on Rings of operators, extends from 1928 to 1932 with certain later additions.
  15. [15]
    [PDF] Matrix Theory, Math6304 Lecture Notes from September 18, 2012
    Sep 18, 2012 · In other words, every square matrix is unitarily equivalent to an upper triangular matrix. ... A matrix A ∈ Mn is normal iff ||Ax|| = ||A∗x||, for ...
  16. [16]
    Pauli matrices - Wikipedia
    In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices that are traceless, Hermitian, involutory and unitary.Pauli equation · Involutory matrix · Levi-Civita symbol
  17. [17]
    [PDF] Lecture 5 - Math 504: Advanced Linear Algebra
    Sep 29, 2021 · An n × n-matrix A ∈ cn×n is normal if and only if it is unitarily similar to a diagonal matrix. Proof. =⇒: This is just part (a) of the spectral ...
  18. [18]
    [PDF] 1 Matrix Norms
    This implies that the spectral norm of normal matrices are equal to the maximum eigenvalue in absolute value which is the spectral radius ρ(A): ρ(A) = kAk. 2 ...
  19. [19]
    The importance of being spectral - Quantum Journal
    Jul 9, 2019 · The spectral theorem, a major result in functional analysis, states that any normal (A linear operator is normal if it is closed, densely ...
  20. [20]
    [PDF] Chapter 4 Vector Norms and Matrix Norms - UPenn CIS
    If E is any real or complex vector space of finite dimension, then any two norms on E are equivalent. Next, we will consider norms on matrices. Page 10. 216.
  21. [21]
    [PDF] Normal Matrices* - Robert Grone - University of Waterloo
    In hopes that it will be useful to a wide audience, a long list of conditions on an n-by-n complex matrix A, equivalent to its being normal, is presented.
  22. [22]
    [PDF] A Characterization of Normal Matrices 1
    The result is that any polynomial of first degree is suitable, while not every polynomial of second degree is, although a large class among them is suitable. In ...
  23. [23]
    [PDF] Lecture notes on Numerical Range - Chi-Kwong Li
    ... normal matrix. 4 Location of the numerical range. A compact convex set in C can be written as the intersection of closed half-space and the convex hull of ...
  24. [24]
    Hermitian Matrix -- from Wolfram MathWorld
    Hermitian matrices have real eigenvalues whose eigenvectors form a unitary basis. For real matrices, Hermitian is the same as symmetric. Any matrix C which is ...
  25. [25]
    [PDF] PRINCIPAL COMPONENT ANALYSIS - UConn Health
    •The covariance matrix is square, symmetric with real values*. •The eigenvectors of a Hermitian matrix are orthogonal. As a consequence, the eigenvectors do ...
  26. [26]
    1.8 Positive Semi-Definite Matrices - A First Course in Linear Algebra
    Definition 1.8.​​ A square matrix A of size n is positive semi-definite if A is Hermitian and for all x∈Cn, x ∈ C n , ⟨x,Ax⟩≥0.
  27. [27]
    Positive Definite Matrix -- from Wolfram MathWorld
    A Hermitian (or symmetric) matrix is positive definite iff all its eigenvalues are positive. Therefore, a general complex (respectively, real) matrix is ...
  28. [28]
    [PDF] 1 The Covariance Matrix - TTIC
    A matrix satisfying this property for all u is called positive semi- definite. The covariance matrix is always both symmetric and positive semi- definite.
  29. [29]
    Unitary Matrix -- from Wolfram MathWorld
    A square matrix U is a unitary matrix if U^(H)=U^(-1), (1) where U^(H) denotes the conjugate transpose and U^(-1) is the matrix inverse.
  30. [30]
    [PDF] Lecture Notes for Math 623 Matrix Analysis - Michael O'Sullivan
    Apr 18, 2013 · (d) The set of normal matrices is closed under scalar multiplication, but is not closed under addition and is not closed under multiplication. ( ...
  31. [31]
    [PDF] Physics Math Linear Algebra, Fourier Analysis, Differential ...
    Use quaternions to show that the rotation matrix through an angle ✓ about ... The Fourier matrix is unitary. For example,. F4 F†4 = 1. 4. 1. 1. 1. 1. 1 i. 1 i. 1.
  32. [32]
    [PDF] Basic concepts in quantum computation - arXiv
    The most common quantum gate is the Hadamard gate, a single qubit gate. H performing the unitary transformation known as the Hadamard transform. It is ...
  33. [33]
    [PDF] CHAPTER 2
    Show that a normal matrix is unitary if and only if all its eigenvalues have absolute value 1. ... on an open arc of the unit circle of length ; such a matrix is ...<|control11|><|separator|>
  34. [34]
    [PDF] The QR Algorithm - Hannes Thiel
    Real normal matrices are orthogonally block-diagonalizable with blocks of size 1x1 and 2x2: ... Real symmetric matrices are orthogonally diagonalizable.
  35. [35]
    [PDF] Contributions of Issai Schur to Analysis - arXiv
    Jun 13, 2007 · The Schur product and Schur multipliers. Let A and B be matrices of the same size whose entries are either real or complex numbers (or even ...
  36. [36]
    [PDF] Commutative Quaternion Matrices - arXiv
    Jun 25, 2013 · the ith conjugate transpose of A. A square matrix A ∈ Mn (H) is said to be normal matrix by ith conjugate, if. AA†i = A†i A; Hermitian ...
  37. [37]
    [PDF] Spectral theory in Hilbert spaces (ETH Zürich, FS 09) E. Kowalski
    The goal of spectral theory, broadly defined, can be described as trying to “classify” all linear operators, and the restriction to Hilbert space occurs both ...<|control11|><|separator|>
  38. [38]
    [PDF] Lecture Notes on the Spectral Theorem - Dartmouth Mathematics
    Aug 10, 2018 · The object of these lectures was to give a reasonably self-contained proof of the Spectral Theorem for bounded normal operators on an infinite ...
  39. [39]
    [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).
  40. [40]
    [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.Missing: normal | Show results with:normal
  41. [41]
    [PDF] EE 261 - The Fourier Transform and its Applications
    In the modern formulation of partial differential equations, the Fourier transform has become the basis for defining the objects of study, while still ...