Fact-checked by Grok 2 weeks ago

Diagonalizable matrix

In linear algebra, a diagonalizable matrix is a square matrix A that is similar to a , meaning there exists an P such that P^{-1} A P = D, where D is a whose entries are the eigenvalues of A. This diagonalizes A, effectively representing it in a basis of its eigenvectors, with the diagonal entries of D corresponding to the scaling factors along those basis directions. A matrix A \in \mathbb{R}^{n \times n} (or over the complex numbers) is diagonalizable it has a full set of n linearly independent eigenvectors, which occurs when the geometric multiplicity equals the algebraic multiplicity for each eigenvalue. Matrices with distinct eigenvalues are always diagonalizable, as each eigenvalue has algebraic multiplicity one and thus a full eigenspace . Not all matrices are diagonalizable; for example, certain blocks with repeated eigenvalues and deficient eigenspaces are not. Diagonalization is crucial for simplifying matrix computations, such as raising a matrix to a power A^k = P D^k P^{-1}, where D^k is easily computed by raising each diagonal entry to the k-th power. It also facilitates the computation of matrix exponentials e^A = P e^D P^{-1}, which are essential in solving systems of linear differential equations and modeling continuous-time dynamical systems. In applications like Markov chains, , and , diagonalizable matrices enable efficient and stability analysis.

Fundamentals

Definition

In linear algebra, a square matrix A of size n \times n over a F is diagonalizable if there exists an P (also of size n \times n) and a D such that A = P D P^{-1}, where the diagonal entries of D are the eigenvalues of A. This relation expresses A as similar to a via a given by the columns of P. The P^{-1} A P = D preserves key spectral properties of A, including its eigenvalues, , , and , as these are invariant under similarity. For matrices with real entries, the field F is typically extended to the of complex numbers to guarantee the existence of all eigenvalues, even if they are non-real. The origins of the concept trace back to Joseph-Louis Lagrange's 18th-century investigations of forms, where he employed linear transformations to reduce them to diagonal form, and were formalized within modern linear algebra by David Hilbert's work on around 1900.

Characterization

A square matrix A over a F is diagonalizable if and only if there exists a basis of the underlying consisting of n linearly independent eigenvectors of A, where n is the of the . This condition ensures that A can be represented by a in some basis, as the eigenvectors form the columns of the P in the A = P D P^{-1}, where D is diagonal. An equivalent characterization involves the multiplicities of the eigenvalues of A. For each eigenvalue \lambda of A, the geometric multiplicity, defined as \dim(\ker(A - \lambda I)), must equal the algebraic multiplicity, which is the multiplicity of \lambda as a root of the \det(A - \lambda I). This equality holds across all eigenvalues if and only if the sum of the geometric multiplicities is n, guaranteeing a full basis of eigenvectors. Another criterion uses the minimal polynomial of A, the monic polynomial of least degree that annihilates A. The matrix A is diagonalizable over F if and only if its minimal polynomial factors into distinct linear factors over F, meaning it has no repeated roots. This condition implies that the minimal polynomial splits completely into linear terms without multiplicity greater than one. Over an algebraically closed field such as \mathbb{C}, the characteristic polynomial of any matrix always splits into linear factors by the fundamental theorem of algebra. In this setting, diagonalizability reduces to the minimal polynomial having distinct linear factors or, equivalently, the algebraic and geometric multiplicities matching for each eigenvalue, as the splitting is automatic. For matrices over general fields F, diagonalizability requires both that the characteristic polynomial splits into linear factors over F and that the geometric multiplicity equals the algebraic multiplicity for each root.

Diagonalization Techniques

Diagonalization Procedure

To diagonalize a square matrix A \in \mathbb{R}^{n \times n} (or over \mathbb{C}), the procedure involves computing its to determine if a basis of n linearly independent eigenvectors exists, enabling the A = PDP^{-1} where D is diagonal and P is invertible./07%3A_Spectral_Theory/7.02%3A_Diagonalization) The first step is to find the eigenvalues by solving the \det(A - \lambda I) = 0, where I is the ; the roots \lambda_1, \dots, \lambda_k (with possible multiplicities) are the eigenvalues of A. For each distinct eigenvalue \lambda_i, compute the corresponding eigenspace by solving the eigenvector equation (A - \lambda_i I)v = 0 to find a basis for the null space; the dimension of this eigenspace is the geometric multiplicity of \lambda_i./07%3A_Spectral_Theory/7.02%3A_Diagonalization) The matrix A is diagonalizable the geometric multiplicity equals the algebraic multiplicity (from the ) for every eigenvalue, ensuring the eigenspaces collectively \mathbb{R}^n (or \mathbb{C}^n) with n linearly eigenvectors. If n linearly independent eigenvectors v_1, \dots, v_n are obtained, form the matrix P with these as columns and the D = \operatorname{diag}(\lambda_1, \dots, \lambda_n); the is then A = PDP^{-1}, which can be verified by direct computation./07%3A_Spectral_Theory/7.02%3A_Diagonalization) If the geometric multiplicity is less than the algebraic multiplicity for any eigenvalue, the matrix is defective and not diagonalizable over the field; in such cases, the Jordan canonical form provides an alternative block-diagonal representation using generalized eigenvectors, though it requires additional computational steps beyond standard eigendecomposition. For large matrices, numerical implementations in software such as MATLAB's eig function or Python's SciPy linalg.eig are essential, as they employ algorithms like the QR method for eigenvalue computation. However, these are sensitive to floating-point errors, particularly for matrices with clustered or nearly degenerate eigenvalues, where small perturbations can lead to inaccurate eigenvectors or failure to detect linear independence.

Simultaneous Diagonalization

A set of matrices \{A_1, \dots, A_k\} is said to be simultaneously diagonalizable if there exists a single P such that P^{-1} A_i P is a for each i = 1, \dots, k. This extends the of from individual matrices to families, requiring a common basis of eigenvectors. A fundamental result states that if a set of diagonalizable matrices commute pairwise (i.e., [A_i, A_j] = A_i A_j - A_j A_i = 0 for all i, j), then they are simultaneously diagonalizable; conversely, if they are simultaneously diagonalizable, they . More generally, simultaneous diagonalizability holds if the matrices share a common basis of eigenvectors. For two commuting diagonalizable matrices A and B, there exists an invertible P such that P^{-1} A P = D, \quad P^{-1} B P = E, where D and E are diagonal. A proof sketch proceeds by on the . For the base case, assume A has a distinct eigenvalue \lambda with eigenspace V_\lambda. Since B commutes with A, B preserves V_\lambda, so V_\lambda decomposes into generalized eigenspaces of B. By diagonalizability of B, these are eigenspaces, allowing simultaneous on V_\lambda. The result extends to the full space by . One key application arises in the of quadratic forms, where a set of symmetric matrices can be simultaneously diagonalized via (i.e., P^T A_i P diagonal for nonsingular P) if they commute, facilitating the reduction of multiple quadratic forms to . For non-commuting matrices, simultaneous is not generally possible unless they share common eigenspaces, as commutativity ensures the preservation of eigenspaces under conjugation.

Examples

Diagonalizable Matrices

Diagonal matrices provide the simplest example of diagonalizable matrices. A diagonal matrix D is already in diagonal form, so it is trivially diagonalizable with the identity matrix P = I as the change-of-basis matrix, satisfying D = P D P^{-1}./07:_Spectral_Theory/7.02:_Diagonalization) Real symmetric matrices are always diagonalizable over the real numbers, as guaranteed by the spectral theorem. This theorem states that every real symmetric matrix has real eigenvalues and can be diagonalized using an orthogonal matrix P, where P^T = P^{-1}, yielding A = P D P^T with D diagonal containing the eigenvalues. For instance, the symmetric matrix \begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix} has eigenvalues 3 and 1, with corresponding orthonormal eigenvectors \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix} and \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -1 \end{pmatrix}, forming the columns of an orthogonal P. Projection matrices, which are idempotent (A^2 = A), offer another clear class of diagonalizable matrices. Such matrices have eigenvalues of only 0 or 1, and since the minimal divides x(x-1) (which has distinct roots), they are diagonalizable over the reals. A simple example is the orthogonal onto the x-axis in \mathbb{R}^2, given by A = \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix}, which is already diagonal with eigenvalues 1 and 0. To illustrate the diagonalization procedure explicitly, consider the $2 \times 2 matrix A = \begin{pmatrix} 1 & 1 \\ 0 & 2 \end{pmatrix}. The characteristic polynomial is \det(A - \lambda I) = (\lambda - 1)(\lambda - 2) = 0, yielding eigenvalues \lambda_1 = 1 and \lambda_2 = 2. For \lambda_1 = 1, solve (A - I)\mathbf{v} = 0, or \begin{pmatrix} 0 & 1 \\ 0 & 1 \end{pmatrix} \begin{pmatrix} v_1 \\ v_2 \end{pmatrix} = \begin{pmatrix} 0 \\ 0 \end{pmatrix}, giving eigenvector \mathbf{v}_1 = \begin{pmatrix} 1 \\ 0 \end{pmatrix}. For \lambda_2 = 2, solve (A - 2I)\mathbf{v} = 0, or \begin{pmatrix} -1 & 1 \\ 0 & 0 \end{pmatrix} \begin{pmatrix} v_1 \\ v_2 \end{pmatrix} = \begin{pmatrix} 0 \\ 0 \end{pmatrix}, giving eigenvector \mathbf{v}_2 = \begin{pmatrix} 1 \\ 1 \end{pmatrix}. The matrix P has these as columns: P = \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}, and D = \begin{pmatrix} 1 & 0 \\ 0 & 2 \end{pmatrix}, so A = P D P^{-1}. Over the complex numbers, normal matrices (those satisfying A^* A = A A^*, where A^* is the ) are unitarily diagonalizable. This means there exists a U (with U^* = U^{-1}) such that A = U D U^*, with D . Examples include Hermitian and .

Non-Diagonalizable Matrices

A is non-diagonalizable, or defective, if it does not possess a complete set of linearly independent eigenvectors that span the entire , which occurs when the geometric multiplicity of at least one eigenvalue is strictly less than its algebraic multiplicity. In such cases, the eigenspaces fail to provide a basis for the space, preventing similarity to a . A classic example of a non-diagonalizable matrix is the 2×2 Jordan block J = \begin{pmatrix} \lambda & 1 \\ 0 & \lambda \end{pmatrix}, where \lambda is an eigenvalue with algebraic multiplicity 2, as determined by the characteristic polynomial \det(J - \mu I) = (\lambda - \mu)^2. However, the geometric multiplicity is 1, since the eigenspace is spanned solely by vectors of the form \begin{pmatrix} 1 \\ 0 \end{pmatrix}, as \ker(J - \lambda I) has dimension 1. This is evident from the matrix J - \lambda I = \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix}, which has rank 1 and thus nullity 1, not 2. Another example arises over the real numbers: the rotation matrix R = \begin{pmatrix} 0 & -1 \\ 1 & 0 \end{pmatrix}, which represents a 90-degree counterclockwise rotation in \mathbb{R}^2. Its eigenvalues are the complex numbers i and -i, with no real eigenvectors, so the eigenspaces over \mathbb{R} are trivial and do not span \mathbb{R}^2, rendering R non-diagonalizable over the reals. In contrast to fully diagonalizable matrices, non-diagonalizable ones can still be analyzed via the Jordan canonical form, a block-diagonal matrix consisting of Jordan blocks that capture the structure of generalized eigenspaces for each eigenvalue. The theorem states that every square matrix over an algebraically closed field is similar to a unique Jordan canonical form (up to block ordering). This form, developed by Camille Jordan in his 1870 treatise Traité des substitutions et des équations algébriques, provides the closest analogue to diagonalization for defective matrices.

Applications

Matrix Functions

One key application of diagonalizability arises in the evaluation of functions applied to matrices. If a matrix A is diagonalizable, expressed as A = P D P^{-1} where D is the of eigenvalues \lambda_i and P consists of corresponding eigenvectors, then for a f defined on the complex numbers, the matrix f(A) can be computed as f(A) = P f(D) P^{-1}, where f(D) is the with entries f(\lambda_i). This formula holds for any f that is analytic in a neighborhood of the spectrum of A, leveraging the fact that functions of act entrywise on the diagonal. A prominent example is the computation of matrix powers. For a positive k, A^k = P D^k P^{-1}, where D^k has diagonal entries \lambda_i^k. This reduces the effort of repeated multiplications, which would otherwise require O(kn^3) operations via direct powering, to a single eigendecomposition followed by diagonal . Similarly, for matrix p(\lambda) = \sum_{k=0}^m c_k \lambda^k, the evaluation yields p(A) = P p(D) P^{-1}, with p(D) obtained by applying the polynomial to each eigenvalue, simplifying the process from on the full matrix to scalar operations. The matrix exponential e^A = \sum_{k=0}^\infty \frac{A^k}{k!} also benefits, computed as e^A = P e^D P^{-1}, where e^D has entries e^{\lambda_i}. This is particularly valuable in solving systems of linear equations, where e^{At} generates the solution operator. In general, the equation f(A) = P \operatorname{diag}(f(\lambda_1), \dots, f(\lambda_n)) P^{-1} encapsulates this approach. Computationally, while the eigendecomposition costs O(n^3), the subsequent diagonal function evaluation is O(n), offering efficiency for large n when A is well-conditioned, though numerical stability depends on the condition number of P.

Quantum Mechanics

In quantum mechanics, physical observables such as , , and are represented by Hermitian operators on a . These operators are always diagonalizable, admitting a complete set of orthonormal eigenvectors with real eigenvalues, as guaranteed by the for operators. The eigenvalues correspond to the possible measurable outcomes of the , while the eigenvectors form the basis in which the operator is diagonal. Upon measurement of an observable, the quantum state collapses to one of the eigenstates, with the probability of obtaining a specific eigenvalue \lambda_n given by the squared modulus of the coefficient in the expansion of the initial state over the eigenbasis: P(\lambda_n) = |\langle \psi_n | \psi \rangle|^2, where |\psi_n\rangle is the normalized eigenvector. This probabilistic interpretation arises directly from the diagonal form, ensuring that the outcomes are the eigenvalues and the likelihoods are determined by the overlaps with the eigenvectors. A key example is the operator H, which governs the of a quantum system; its yields the discrete levels as eigenvalues, essential for understanding bound states in atoms or molecules. In finite-dimensional models, such as discretized position or momentum operators on a basis, similarly reveals the of allowed values. For time-independent s, the operator U(t) = e^{-iHt/\hbar} simplifies dramatically in the eigenbasis, becoming diagonal with entries e^{-iE_n t / \hbar}, where E_n are the eigenvalues; this facilitates the computation of state evolution via the matrix exponential. The of a Hermitian takes the form H = U D U^{\dagger}, where U is a whose columns are the orthonormal eigenvectors, and D is diagonal with the real eigenvalues on the ; measurements project the state onto one of these columns, yielding the corresponding eigenvalue as the observed value. This mathematical framework for observables and their diagonalization originated in Erwin Schrödinger's development of wave mechanics in 1926, where the time-independent was posed as an eigenvalue problem for the . It was rigorously formalized by in 1932, who established the structure and underpinning quantum measurements and dynamics. In extensions beyond standard Hermitian quantum mechanics, PT-symmetric operators provide examples of non-Hermitian but diagonalizable Hamiltonians with real eigenvalues, enabling descriptions of systems with balanced gain and loss while preserving unitarity through a modified inner product.

Operator Theory

In , diagonalizability for bounded linear operators on infinite-dimensional generalizes the finite-dimensional case by requiring the existence of a complete of eigenvectors. Specifically, a bounded linear T: \mathcal{H} \to \mathcal{H} on a \mathcal{H} is diagonalizable if there exists a complete \{\psi_n\}_{n=1}^\infty such that T \psi_n = \lambda_n \psi_n for scalars \lambda_n \in \mathbb{C}, allowing T to be expressed in the form T = \sum_{n=1}^\infty \lambda_n |\psi_n\rangle \langle \psi_n|, where the series converges in the strong operator topology. This representation contrasts with finite matrices, as the infinite sum must account for potential accumulation of eigenvalues at zero, and not all operators admit such a basis even if they possess eigenvalues. The provides a key characterization: every on a separable is diagonalizable with respect to some of eigenvectors, with real eigenvalues. This result, independently established by Stone and , extends the finite-dimensional briefly referenced in matrix characterizations, ensuring that operators generate a by a real-valued function in the spectral representation. For compact operators, diagonalizability holds under additional conditions, such as self-adjointness, where the eigenvalues form an (except possibly for the kernel). The asserts that every compact on a separable admits such a basis, with eigenvalues \lambda_n satisfying \lambda_n \to 0 as n \to \infty. Examples include integral operators with square-integrable kernels, like the (Tf)(x) = \int_a^b K(x,y) f(y) \, dy where K is symmetric and L^2, which are compact and thus diagonalizable via their eigenfunctions. Non-diagonalizable operators abound in infinite dimensions; for instance, the unilateral S on \ell^2(\mathbb{N}), defined by S(e_n) = e_{n+1} for the \{e_n\}, has no eigenvalues and hence no eigenvector basis. Similarly, quasinilpotent operators, which have \{0\} but are not the zero , cannot be diagonalizable, as any such representation would imply T = 0. A canonical example is the Volterra V on L^2[0,1], given by (Vf)(x) = \int_0^x f(t) \, dt, which is compact, quasinilpotent, and lacks eigenvalues. In , diagonalizable operators play a crucial role in solving partial differential equations through expansions. For elliptic operators, the enables decomposition into series of s, such as for the Laplacian on bounded domains, facilitating and convergence in appropriate norms.

References

  1. [1]
    Diagonalization
    Definition. An n × n matrix A is diagonalizable if it is similar to a diagonal matrix: that is, if there exists an invertible n × n matrix C and a diagonal ...
  2. [2]
    ALAFF Diagonalizing a matrix - Texas Computer Science
    A matrix A∈Cm×m A ∈ C m × m is said to be diagonalizable if and only if there exists a nonsingular matrix X X and diagonal matrix D D such that X−1AX=D. X − 1 A ...
  3. [3]
    [PDF] 5.3 Diagonalization - UC Berkeley math
    A matrix is diagonalizable if it's similar to a diagonal matrix, which is a square matrix with 0s except possibly on the diagonal. A matrix with n distinct ...
  4. [4]
    [PDF] Unit 16: Diagonalization
    A matrix A is diagonalizable if it is similar to a diagonal matrix, meaning there exists an invertible matrix S such that S−1AS is diagonal.
  5. [5]
    EIG-0050: Diagonalizable Matrices and Multiplicity - Ximera
    A matrix is diagonalizable if it can be transformed into a diagonal matrix using an invertible matrix, and if for each eigenvalue, algebraic multiplicity ...
  6. [6]
    4.3 Diagonalization, similarity, and powers of a matrix
    A matrix is diagonalizable if it can be written as a diagonal matrix and an invertible matrix. It is diagonalizable if and only if it has a basis of ...
  7. [7]
    [PDF] 5.5 Similarity and Diagonalization - Emory Mathematics
    In Section 3.3 we studied diagonalization of a square matrix A, and found important applications (for example to linear dynamical systems).
  8. [8]
    Similarity and Diagonalization - A First Course in Linear Algebra
    Good things happen when a matrix is similar to a diagonal matrix. For example, the eigenvalues of the matrix are the entries on the diagonal of the diagonal ...
  9. [9]
    Diagonalizable Matrix -- from Wolfram MathWorld
    The diagonalization theorem states that an n×n matrix A is diagonalizable if and only if A has n linearly independent eigenvectors.
  10. [10]
    [PDF] Highlights in the History of Spectral Theory
    The central fact of modern spectral theory is that certain linear operators on infinite dimensional spaces can also be presented in "diagonal" form. Thus the ...
  11. [11]
    [PDF] EIGENVALUES AND EIGENVECTORS 1. Diagonalizable linear ...
    Theorem 5.1. A matrix A ∈ Rn×n is diagonalizable if and only if the sum of the geometric multiplicities of all of the eigenvalues of A is n ...
  12. [12]
    [PDF] n × n matrix. A is diagonalizable if and only if it has eigenvectors
    Thm. [C] A : n × n matrix. A is diagonalizable if and only if dimEλ(A) is equal to the multiplicity of λ for every eigenvalue λ of A. (λI − A)X = O.Missing: theorem | Show results with:theorem
  13. [13]
    [PDF] algebraic and geometric multiplicities of eigenvalues, generalized ...
    A matrix A admits a basis of eigenvectors if and only of for every its eigenvalue λ the geometric multiplicity of λ is equal to the algebraic multiplicity of λ.
  14. [14]
    [PDF] the minimal polynomial and some applications - Penn Math
    Theorem 4.11. Let A: V → V be a linear operator. Then A is diagonalizable if and only if its minimal polynomial in F[T] splits in F[T] and has distinct roots.
  15. [15]
    [PDF] Math 4571 – Lecture 24
    A matrix A ∈ Mn×n(F) is diagonalizable (over F) if and only if the characteristic polynomial factors into a product of linear terms in. F[x] and, for each ...
  16. [16]
    Lecture 22: Diagonalization and powers of A | Linear Algebra
    In this lecture we learn to diagonalize any matrix that has n independent eigenvectors and see how diagonalization simplifies calculations.Missing: procedure authoritative sources
  17. [17]
    [PDF] Math 2331 – Linear Algebra - 5.3 Diagonalization
    An n × n matrix A is diagonalizable if and only if A has n linearly independent eigenvectors. In fact, A = PDP-1, with D a diagonal matrix, if and only if the ...
  18. [18]
    [PDF] A useful basis for defective matrices: Jordan vectors and the ... - MIT
    For a diagonalizable matrix, the fundamen- tal vectors are the eigenvectors, which are useful in their own right and give the diagonalization of the matrix as a.
  19. [19]
    [PDF] NUMERICAL METHODS FOR LARGE EIGENVALUE PROBLEMS ...
    matrix A is diagonalizable and let X be the matrix that transforms it into diagonal ... The total number of operations required for computing a Gram matrix with.
  20. [20]
    [PDF] Numerical Algorithms - People | MIT CSAIL
    A programmer using floating-point data types and operations must be vigilant when it comes to detecting and preventing poor numerical operations. For ...
  21. [21]
    [PDF] Matrix Theory, Math6304 Lecture Notes from September 6, 2012
    Sep 6, 2012 · 32 Theorem. Let A, B ∈ Mn be diagonalizable. Then AB = BA if and only if A and B are simultaneously diagonalizable. Proof.
  22. [22]
    [PDF] Simultaneous commutativity of operators - Keith Conrad
    For linear operators to be simultaneously diagonalizable, they at least have to be indi- vidually diagonalizable, but more is needed (see Example 1). A further ...
  23. [23]
    Simultaneous Diagonalization via Congruence of Hermitian Matrices
    This paper aims at solving the Hermitian SDC problem, which is the simultaneous diagonalization via ∗ -congruence of a finite collection of Hermitian matrices.
  24. [24]
    [PDF] Simultaneous Diagonalization and SVD of Commuting Matrices - arXiv
    Jun 29, 2020 · It is well known that if two diagonalizable matrices have the same eigenvectors, then they commute. The converse also is true and a construction ...
  25. [25]
    [PDF] Symmetric Matrices and the Spectral Theorem - Purdue Math
    May 4, 2025 · A symmetric matrix has entries symmetric across the main diagonal. If A is symmetric, then ⃗v · (A⃗w)=(A⃗v) · ⃗w for all vectors ⃗v and ⃗w.
  26. [26]
    [PDF] Projectors - Purdue Math
    Mar 18, 2023 · (3). So for every idempotent P we have. V = NP ⊕ RP . 1. Page 2. It follows that all idempotents are diagonalizable, with eigenvalues 0 and 1.
  27. [27]
    [PDF] Unitary Diagonalization of Matrices - UMD MATH
    Theorem 3. A matrix A is diagonalizable with a unitary matrix if and only if A is normal. In other words: a) If A is normal there is a unitary matrix S so that ...
  28. [28]
    ALAFF Jordan Canonical Form - Texas Computer Science
    We conclude that there are matrices that are not diagonalizable. We call such matrices defective. 🔗. Definition 9.2.6.1. Defective matrix. A matrix A∈Cm×m A ∈ ...<|control11|><|separator|>
  29. [29]
    [PDF] Jordan Normal form of 2 × 2 matrices - UC Berkeley math
    Corollary: Let A be a 2 × 2 matrix which is not diagonalizable. Then there exist matrices D and N, where D is diagonal and N is nilpotent, with A = D + N. To ...Missing: non- | Show results with:non-
  30. [30]
    [PDF] Lecture 4.2. Jordan form - Purdue Math
    Mar 31, 2020 · Today we finally address deficient matrices, that is those which are non- diagonalizable, which do not have enough eigenvectors to span the ...Missing: defective | Show results with:defective
  31. [31]
    [PDF] lecture april 3, 6, 8: eigenvectors and eigenvalues
    Apr 3, 2020 · How could a matrix not be diagonalizable? Here's another example. Let. A “. ˆ0 ´1. 1. 0. ˙ .
  32. [32]
    [PDF] The Jordan Canonical Form
    honour Camille Jordan (1838 – 1922), who in 1870 published the influential book “Traité des substitutions et des équations algébriques” (667 pages).
  33. [33]
    [PDF] Jordan Normal Form Revisited - Auburn University
    Oct 3, 2007 · A little History: In 1870 the Jordan canonical form appeared in Treatise on substitutions and algebraic equations by Camille Jordan (1838-1922).
  34. [34]
    [PDF] Computing Matrix Functions Higham, Nicholas J. and Al-Mohy ...
    Feb 8, 2010 · The survey is organized by classes of methods, which are broadly those based on similarity transformations, those employing approximation by ...
  35. [35]
    [PDF] Quantum Theory I, Recitation 1 Notes - MIT OpenCourseWare
    As we stated, two diagonalizable operators A and B are simultaneously diagonalizable if and only if they commute, [A, B] = 0. Note that this is equivalent to ...
  36. [36]
    [PDF] Spectral theorems for Hermitian and unitary operators - Purdue Math
    c) there exists an orthogonal basis of the whole space, consisting of eigen- vectors. Thus all Hermitian matrices are diagonalizable. Moreover, for every Her-.
  37. [37]
    [PDF] Lecture 1: Review of Quantum Mechanics
    Jan 1, 2020 · If T is a Hermitian unbounded operator, then there is a spectral theorem. First assume that the spectrum is discrete. Let λi be.
  38. [38]
    [PDF] 2. Introduction to Quantum Mechanics - MIT OpenCourseWare
    3) The probability of obtaining the given eigenvalue in the measurement is the probability amplitude modulus. |2 square. E.g. p(om) = |cm . 2.3.1 ...
  39. [39]
    [PDF] arXiv:1202.2817v1 [quant-ph] 13 Feb 2012
    Feb 13, 2012 · In quantum mechanics, for example, the energy levels of a quantum system are obtained by diagonalization of the system's Hamiltonian.
  40. [40]
    Quantum matrix diagonalization visualized - AIP Publishing
    Nov 1, 2019 · We show how to visualize the process of diagonalizing the Hamiltonian matrix to find the energy eigenvalues and eigenvectors of a generic one-dimensional ...
  41. [41]
    [PDF] 6. Time Evolution in Quantum Mechanics - MIT OpenCourseWare
    Notice that if the operator A is time independent and it commutes with the Hamiltonian E then the operator is conserved, it is a constant of the motion (not ...
  42. [42]
    [PDF] 1926-Schrodinger.pdf
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The analogy is to ...
  43. [43]
    Mathematical Foundations of Quantum Mechanics: New Edition - jstor
    Quantum mechanics was still in its infancy in 1932 when the young John von Neumann ... 4: to define Hilbert space, which furnishes the mathematical basis ...
  44. [44]
    -symmetric quantum mechanics | Rev. Mod. Phys.
    Oct 28, 2024 · However, a P T -symmetric Hamiltonian can also define a physically acceptable quantum-mechanical system even if the Hamiltonian is not Hermitian ...
  45. [45]
    The Volterra operator is a compact universal quasinilpotent
    Every compact quasinilpotent Hilbert space operator can be uniformly approximated by operators that are similar to the Volterra operator in L2([0,1],dx).