Fact-checked by Grok 2 weeks ago

Orthogonal diagonalization

Orthogonal diagonalization is a fundamental concept in linear algebra referring to the decomposition of a real symmetric matrix A as A = PDP^T, where P is an orthogonal matrix (satisfying P^T P = I) whose columns form an orthonormal basis of eigenvectors of A, and D is a diagonal matrix with the corresponding eigenvalues of A on its main diagonal. This process is possible precisely when A is symmetric, as guaranteed by the spectral theorem, which states that every real symmetric matrix has real eigenvalues and is orthogonally diagonalizable via an orthonormal set of n eigenvectors for an n \times n matrix. Eigenvectors corresponding to distinct eigenvalues are automatically orthogonal, and for repeated eigenvalues, the eigenspaces can be orthonormalized using processes like Gram-Schmidt. The significance of orthogonal diagonalization lies in its ability to simplify matrix computations and provide geometric insights, such as representing linear transformations as scalings along perpendicular axes in an orthonormal . It is particularly useful in applications including the analysis of quadratic forms, where it identifies principal axes; problems; and for projections onto eigenspaces. In broader contexts, such as physics and , this facilitates efficient numerical methods and interpretations of phenomena involving symmetric operators, like vibrations or quantum mechanical observables.

Prerequisites

Symmetric matrices

A is a square matrix A that equals its own , denoted A = A^T, where the A^T has entries swapped across the . In the context of real vector spaces, symmetric matrices typically consist of real entries and play a central role in representing bilinear forms that are symmetric. One key property of symmetric matrices is their association with s, which are homogeneous polynomials of degree two expressed as q(\mathbf{x}) = \mathbf{x}^T A \mathbf{x} for a \mathbf{x} \in \mathbb{R}^n. This arises naturally from the inner product \langle T\mathbf{v}, \mathbf{v} \rangle where T is a linear operator corresponding to the A. Additionally, symmetric matrices are invariant under by orthogonal matrices, meaning if P is orthogonal, then P^T A P is also symmetric, preserving the structure in orthonormal bases. Diagonal matrices provide trivial examples of symmetric matrices, as their off-diagonal entries are zero, satisfying A = A^T immediately. A simple 2×2 symmetric matrix can be constructed as A = \begin{pmatrix} a & b \\ b & c \end{pmatrix}, where a, b, c \in \mathbb{R}, ensuring symmetry by matching the off-diagonal elements. A fundamental characteristic of real symmetric matrices is that all their eigenvalues are real numbers.

Orthogonal matrices and transformations

An is a square Q satisfying Q^T Q = I, where Q^T denotes the of Q and I is the ; equivalently, the inverse of an orthogonal matrix is its , Q^{-1} = Q^T. This condition ensures that the columns (and rows) of Q form an orthonormal set of vectors, meaning each has unit Euclidean norm and they are pairwise orthogonal. Consequently, the columns of Q constitute an for the underlying . Orthogonal matrices preserve key geometric structures under linear transformations. Specifically, for any x, the norm is : \| Qx \| = \| x \|, as this follows from the preservation of the (Qx)^T (Qx) = x^T Q^T Q x = x^T x. Similarly, inner products are preserved: x^T y = (Qx)^T (Qy), which implies that angles between vectors remain unchanged, since the cosine of \theta between x and y is \cos \theta = \frac{x^T y}{\|x\| \|y\|}, and all components are under Q. These properties ensure that orthogonal transformations do not distort lengths, shapes, or orientations in the , making them isometries. In terms of geometric interpretations, orthogonal matrices represent rotations and reflections in . Rotation matrices, which preserve , have +1, while reflection matrices, which reverse , have -1; in general, the of any real orthogonal matrix is \pm 1. For example, in two dimensions, a rotation by \theta is given by the matrix \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix}, which satisfies the orthogonality condition and preserves distances. Reflections, such as across the x-axis in \mathbb{R}^2, are represented by \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}, with -1. Orthogonal conjugation by such matrices preserves the of symmetric matrices.

Main Theorem

Statement of the spectral theorem

The spectral theorem for real symmetric matrices asserts that every real symmetric n \times n matrix A (satisfying A^T = A) is orthogonally diagonalizable. Specifically, there exists an Q (with Q^T Q = I) whose columns form an of \mathbb{R}^n consisting of eigenvectors of A, and a D = \operatorname{diag}(\lambda_1, \lambda_2, \dots, \lambda_n) with real eigenvalues \lambda_i on the diagonal, such that A = Q D Q^T. This decomposition implies that symmetric matrices possess a complete set of orthonormal eigenvectors, enabling the representation of forms associated with A as sums of squares weighted by the eigenvalues \lambda_i. The reality of the eigenvalues and the of the eigenvectors follow directly from the of A. The theorem's development in the is attributed to key contributions from , who in 1829 proved the diagonalizability of symmetric matrices via orthogonal transformations; , who advanced related determinant and transformation techniques around 1830 (published 1841); and , who in the 1850s–1860s explored eigenvalue properties in linear transformations, laying groundwork for the full spectral framework.

Conditions for orthogonal diagonalizability

A A is orthogonally diagonalizable it is symmetric, meaning A = A^T. This condition ensures the existence of an P such that P^T A P = D, where D is a containing the real eigenvalues of A. Symmetric matrices always possess real eigenvalues and a complete set of orthonormal eigenvectors, which form the columns of P. More generally, a real matrix is if it commutes with its , satisfying A A^T = A^T A. However, for orthogonal diagonalizability over the reals, the matrix must also have exclusively real eigenvalues; non-symmetric matrices, such as skew-symmetric ones where A^T = -A, typically have purely imaginary eigenvalues and thus cannot be orthogonally diagonalized unless they are the zero matrix. In practice, the sufficient and necessary condition for real matrices reduces to symmetry, as real matrices with all real eigenvalues are symmetric. Over the complex numbers, the analogous condition involves Hermitian matrices, where A = A^* with A^* denoting the . Hermitian matrices are unitarily diagonalizable, generalizing the real symmetric case, with real eigenvalues and an orthonormal eigenbasis under the complex inner product. Not all diagonalizable matrices are orthogonally diagonalizable; for instance, the matrix \begin{pmatrix} 1 & 1 \\ 0 & 2 \end{pmatrix} is diagonalizable over the reals but not (hence not symmetric), as its eigenvectors are not orthogonal. This highlights that diagonalizability requires only linearly independent eigenvectors, whereas orthogonal diagonalizability demands an orthonormal eigenbasis.

Proof and Derivation

Outline of the proof

The proof of the spectral theorem, which asserts that every real symmetric matrix is orthogonally diagonalizable, relies on mathematical induction on the dimension n of the matrix. For the base case where n=1, the matrix is a scalar and thus already diagonal with a real eigenvalue, satisfying the theorem trivially. Assuming the theorem holds for all symmetric matrices of dimension less than n, the characteristic polynomial of the n \times n symmetric matrix A is considered first; since A has real entries, this polynomial has real coefficients, and the reality of its roots (eigenvalues) for symmetric matrices ensures at least one real eigenvalue exists. To identify one such eigenvalue explicitly, the Rayleigh quotient is maximized over the unit sphere in \mathbb{R}^n, which, by the compactness of the sphere and continuity of the quotient, attains a maximum value that serves as a real eigenvalue \lambda, with the maximizer providing a corresponding eigenvector. This eigenvector spans a one-dimensional eigenspace, and due to the symmetry of A, the to this eigenspace is an under A, reducing the problem to an (n-1) \times (n-1) via onto this (or equivalently, a in block form). By the induction hypothesis, this reduced admits an of eigenvectors spanning the complement. The full space is thus spanned by the initial eigenvector together with the orthonormal eigenbasis of the complement, forming an orthonormal eigenbasis for \mathbb{R}^n. The spectral theorem follows from iteratively applying this eigenvalue extraction process across dimensions.

Detailed steps using eigenvectors

The proof of the spectral theorem for real symmetric matrices proceeds by establishing the existence of real eigenvalues and an orthonormal basis of eigenvectors. The first step involves solving the characteristic equation \det(A - \lambda I) = 0 to find the eigenvalues \lambda. For a real symmetric matrix A, all eigenvalues are real, which follows from the properties of quadratic forms: if Av = \lambda v for a nonzero eigenvector v (possibly complex), then v^* A v = \lambda v^* v, where v^* denotes the conjugate transpose. Since A is symmetric and real, v^* A v is real, and v^* v > 0, implying \lambda is real. The second step confirms the existence of eigenvectors corresponding to each eigenvalue and their orthogonality when eigenvalues are distinct. For each real eigenvalue \lambda_i, the eigenspace is nontrivial, allowing selection of eigenvectors v_i such that A v_i = \lambda_i v_i. Specifically, for distinct eigenvalues \lambda and \mu with corresponding eigenvectors v and w, arises because v^T A w = \mu v^T w and, by symmetry of A, w^T A v = \lambda w^T v, leading to (\lambda - \mu) v^T w = 0. Thus, since \lambda \neq \mu, it follows that v^T w = 0. When eigenvalues have multiplicity greater than one, the eigenvectors within the same eigenspace may not be orthogonal. To address this, apply the Gram-Schmidt orthogonalization process to a basis of the eigenspace, yielding an orthonormal set of eigenvectors for that eigenvalue. This ensures the entire collection of eigenvectors across all eigenspaces can be orthogonalized. Finally, normalize these orthogonal eigenvectors to form an \{v_1, \dots, v_n\} for \mathbb{R}^n. The matrix Q has these normalized vectors as its columns, satisfying Q^T Q = I, and the D = \diag(\lambda_1, \dots, \lambda_n) contains the eigenvalues. This completes the decomposition A = Q D Q^T.

Computational Methods

Finding eigenvalues and eigenvectors

For symmetric matrices, the eigenvalues are always real numbers, and they can be ordered in non-decreasing sequence as \lambda_1 \leq \lambda_2 \leq \dots \leq \lambda_n. The algebraic method to find eigenvalues begins by forming the characteristic polynomial p(\lambda) = \det(A - \lambda I), where A is the n \times n symmetric matrix and I is the identity matrix; the eigenvalues are the roots of this polynomial equation p(\lambda) = 0. For small matrices, such as $2 \times 2, the characteristic polynomial is quadratic, and the eigenvalues are given explicitly by the quadratic formula: \lambda = \frac{\operatorname{trace}(A) \pm \sqrt{[\operatorname{trace}(A)]^2 - 4 \det(A)}}{2}, where \operatorname{trace}(A) is the sum of the diagonal elements and \det(A) is the determinant. This approach is exact but becomes computationally intensive for larger n, as solving high-degree polynomials requires numerical root-finding techniques. Numerical methods are essential for larger symmetric matrices. The , after reducing the matrix to tridiagonal form via transformations, iteratively applies QR decompositions to converge to the eigenvalues on the diagonal. This method is stable and efficient for symmetric matrices, typically requiring O(n^3) operations overall, with rapid convergence for well-separated eigenvalues. For approximating the dominant eigenvalue (the one with largest magnitude), the power iteration method starts with a random v_0 and iteratively computes v_{k+1} = A v_k / \|A v_k\|, converging to the corresponding eigenvector as k increases, with the eigenvalue estimated by the v_k^T A v_k. Once an eigenvalue \lambda is known, the corresponding eigenvectors are found by solving the linear system (A - \lambda I)v = 0 for non-trivial solutions v \neq 0, typically via or null space computation. These eigenvectors are then normalized to unit length, \|v\| = 1, to ensure they form an when assembled into the for .

Constructing the orthogonal matrix

Once the eigenvalues and corresponding eigenvectors of a symmetric matrix A have been computed, the orthogonal matrix Q is constructed by arranging the normalized eigenvectors as its columns. Specifically, for an n \times n symmetric matrix A with eigenvalues \lambda_1, \dots, \lambda_n and associated eigenvectors v_1, \dots, v_n, each eigenvector v_i is normalized by dividing it by its Euclidean norm \|v_i\|, yielding unit vectors u_i = v_i / \|v_i\|. The matrix Q is then formed as Q = [u_1 \, | \, u_2 \, | \, \dots \, | \, u_n], ensuring that the columns of Q form an for \mathbb{R}^n. In the case of repeated eigenvalues, where the geometric multiplicity exceeds one, an initial basis for the eigenspace may not consist of vectors. To obtain an for such eigenspaces, the Gram-Schmidt orthogonalization process is applied to the basis vectors, followed by normalization. This step preserves the between eigenspaces inherent to symmetric matrices while ensuring intra-eigenspace . The resulting orthonormal eigenvectors are then used to assemble Q as described. To verify the construction, the of Q is confirmed by computing Q^T Q, which should equal the I due to the of the columns. Additionally, the full is validated by reconstructing A as A = Q D Q^T, where D is the of eigenvalues, and checking that this equals the original A up to numerical precision. These checks ensure the accuracy of the . For numerical stability in large-scale computations, direct assembly from eigenvectors can be sensitive to rounding errors; instead, transformations such as reflections or Givens rotations are employed to build Q iteratively, minimizing loss of during the process. These methods are integral to algorithms like the QR iteration for eigendecomposition.

Examples and Illustrations

Diagonalization of a 2x2 symmetric matrix

To illustrate the process of orthogonal diagonalization, consider the symmetric 2×2 matrix A = \begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}. The eigenvalues are found by solving the characteristic equation \det(A - \lambda I) = 0, which yields the polynomial \lambda^2 - 4\lambda + 3 = 0. The roots are \lambda_1 = 3 and \lambda_2 = 1. For \lambda_1 = 3, solve (A - 3I)\mathbf{v}_1 = \mathbf{0}, giving the eigenvector \mathbf{v}_1 = \begin{pmatrix} 1 \\ 1 \end{pmatrix}. Normalizing to unit length produces \hat{\mathbf{v}}_1 = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix}. For \lambda_2 = 1, solve (A - I)\mathbf{v}_2 = \mathbf{0}, yielding \mathbf{v}_2 = \begin{pmatrix} -1 \\ 1 \end{pmatrix}, normalized to \hat{\mathbf{v}}_2 = \frac{1}{\sqrt{2}} \begin{pmatrix} -1 \\ 1 \end{pmatrix}. These eigenvectors are orthogonal, as their dot product is zero. The orthogonal matrix Q has these normalized eigenvectors as columns: Q = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & -1 \\ 1 & 1 \end{pmatrix}, and the is D = \begin{pmatrix} 3 & 0 \\ 0 & 1 \end{pmatrix}. To verify, compute Q D Q^T: First, D Q^T = \frac{1}{\sqrt{2}} \begin{pmatrix} 3 & 3 \\ -1 & 1 \end{pmatrix}. Then, Q D Q^T = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & -1 \\ 1 & 1 \end{pmatrix} \cdot \frac{1}{\sqrt{2}} \begin{pmatrix} 3 & 3 \\ -1 & 1 \end{pmatrix} = \frac{1}{2} \begin{pmatrix} 3 + 1 & 3 - 1 \\ 3 - 1 & 3 + 1 \end{pmatrix} = \begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix} = A. This confirms the decomposition A = Q D Q^T. Geometrically, the columns of Q correspond to a by 45 degrees, aligning the with the principal axes of A, where the matrix acts as independent scalings by the eigenvalues along these directions.

Application to a

Orthogonal diagonalization plays a key role in (), where the of multivariate data is decomposed to reveal underlying structures of variance. Consider a derived from data points, such as measurements of two correlated variables: A = \begin{pmatrix} 5 & 2 \\ 2 & 3 \end{pmatrix} Here, the diagonal elements represent the variances of each variable (5 and 3), while the off-diagonal elements capture their covariance (2). Performing orthogonal diagonalization on this symmetric positive semi-definite matrix yields A = Q D Q^T, where D is the diagonal matrix of eigenvalues \lambda_1 = 4 + \sqrt{5} \approx 6.24 and \lambda_2 = 4 - \sqrt{5} \approx 1.76, ordered decreasingly. The columns of the orthogonal matrix Q are the corresponding normalized eigenvectors, providing an orthonormal basis for the principal directions. For \lambda_1, one eigenvector is approximately \begin{pmatrix} 0.851 \\ 0.525 \end{pmatrix} (normalized), and for \lambda_2, approximately \begin{pmatrix} -0.525 \\ 0.851 \end{pmatrix}. The eigenvalues quantify the spread (variance) of the data along these principal components: \lambda_1 along the first (major) direction and \lambda_2 along the second (minor) direction. The transformation via Q rotates the original to align with these uncorrelated axes, simplifying representation by decorrelating the variables while preserving total variance (trace of A equals 8, the sum of eigenvalues). In this case, the first principal component explains approximately 78% of the total variance (\lambda_1 / 8), a common scenario where the dominant component captures the bulk of variability for . This aids interpretation by highlighting directions of greatest spread; for instance, the first eigenvector points along the elongated of the . In visualizations of the original , the eigenvectors indicate the orientations of the enclosing the points (e.g., a 95% ), with semi-major and semi-minor lengths proportional to \sqrt{\lambda_1} and \sqrt{\lambda_2}, respectively.

Properties and Extensions

Uniqueness of the diagonalization

The orthogonal diagonalization of a real A = Q D Q^T, where Q is an and D is a containing the eigenvalues of A, exhibits specific uniqueness properties. The matrix D is unique up to of its diagonal entries, as the eigenvalues of A are the roots of its and thus uniquely determined, counting multiplicities. In contrast, the orthogonal matrix Q, whose columns form an of eigenvectors, is not unique. It can vary by of its columns (corresponding to reordering the eigenvalues in D), by sign flips in any individual column (since if \mathbf{v} is an eigenvector, so is -\mathbf{v}), and, more generally, by the choice of within each eigenspace. When all eigenvalues of A are distinct, each eigenspace is one-dimensional, so Q is up to column and individual sign changes. In this case, the only freedoms are the discrete choices of ordering and signs, preserving the overall decomposition's structure. If an eigenvalue \lambda has multiplicity greater than one, its eigenspace has dimension equal to that multiplicity, allowing any for the corresponding columns of Q. This introduces continuous , such as arbitrary rotations within the eigenspace via orthogonal transformations, while maintaining . For instance, if multiplicity is two, the two columns can be replaced by any pair of orthonormal vectors spanning the same . For a of pairwise commuting symmetric matrices, they share a common set of eigenspaces and thus admit simultaneous orthogonal diagonalization A_i = Q D_i Q^T for a single Q, with each D_i diagonal. This common Q inherits the same uniqueness properties, determined by the joint eigenspaces, and the decomposition is unique up to the permutations, sign flips, and basis choices within those shared subspaces.

Relation to positive definite matrices

A symmetric matrix A is defined as positive definite if it satisfies \mathbf{x}^T A \mathbf{x} > 0 for every nonzero \mathbf{x}. This condition is equivalent to all eigenvalues of A being positive. For a positive definite matrix A, orthogonal diagonalization yields A = Q D Q^T, where Q is orthogonal and D is a with all positive entries corresponding to the eigenvalues of A. An alternative factorization for positive definite matrices is the A = L L^T, where L is lower triangular with positive diagonal entries; this provides a numerically stable way to represent A without explicitly computing eigenvalues, though it is closely related to the . One key property arising from this diagonalization is that the trace of A, which equals the sum of its eigenvalues, is positive. In optimization, if the of a twice-differentiable is positive definite at a critical point, that point is a strict local minimum, leveraging the positive eigenvalues to confirm convexity in all directions. Additionally, the principal square root A^{1/2} can be constructed via the diagonalization as A^{1/2} = Q D^{1/2} Q^T, where D^{1/2} has the positive square roots of the eigenvalues on its diagonal; this A^{1/2} is itself symmetric and positive definite.

Applications

In statistics and data analysis

In principal component analysis (PCA), a key statistical technique for dimensionality reduction and data exploration, orthogonal diagonalization of the sample covariance matrix yields the principal components as the eigenvectors and the explained variances as the eigenvalues. This decomposition, possible because the covariance matrix is symmetric and positive semi-definite, identifies orthogonal directions in the data space that capture the maximum successive amounts of variance. The resulting orthogonal matrix rotates the original correlated variables into a new set of uncorrelated principal components, facilitating visualization and noise reduction in high-dimensional datasets. By selecting the top principal components corresponding to the largest eigenvalues, reduces the dataset's dimensionality while retaining a specified proportion of the total variance, often 80-95% in practice for real-world applications. This variance-preserving transformation is particularly valuable in multivariate analysis for handling and improving computational efficiency in subsequent modeling tasks. Orthogonal diagonalization also plays a role in , where it decomposes the correlation or during factor extraction to isolate underlying latent s, with eigenvalues indicating factor strengths. Orthogonal rotations, such as varimax, are then applied to the factor loadings to enhance interpretability while maintaining uncorrelated factors. In the context of distance metrics, orthogonal diagonalization simplifies the computation of the by transforming the into diagonal form, converting the into a sum of squared standardized principal component scores. This approach accounts for variable correlations and scales, providing a scale-invariant measure of multivariate dissimilarity. A classic illustration appears in the analysis of the Iris dataset, where orthogonal diagonalization of the among sepal length, sepal width, length, and width measurements reveals principal components that effectively separate the three (setosa, versicolor, and virginica), with the first two components explaining over 95% of the variance. As seen in the earlier example, this process decorrelates features like length and width, highlighting dominant patterns in .

In physics and quantum mechanics

In quantum mechanics, physical observables are represented by Hermitian operators, which are analogous to real symmetric matrices in finite-dimensional spaces. These operators can always be diagonalized in an , with the eigenvalues corresponding to the possible outcomes of measurements and the eigenvectors forming the basis in which the operator is diagonal. The time-independent , H \psi = E \psi, is solved by diagonalizing the operator H, yielding H = Q E Q^\dagger in the real symmetric case, where Q is an , E is diagonal with energy eigenvalues, and the columns of Q are the orthonormal eigenvectors representing stationary states. In classical mechanics, orthogonal diagonalization is applied to symmetric matrices arising in the analysis of coupled systems, such as mass-spring networks, to identify normal modes of vibration where each mode oscillates independently at a characteristic frequency. For example, in a system of two coupled harmonic oscillators with equal masses m and spring constants k (coupling constant k'), the equations of motion lead to a symmetric matrix whose orthogonal diagonalization produces eigenvalues proportional to the squared normal frequencies \omega_1^2 = k/m (in-phase mode) and \omega_2^2 = (k + 2k')/m (out-of-phase mode), decoupling the motion into independent oscillators.

References

  1. [1]
    [PDF] 8.2 Orthogonal Diagonalization
    In particular, if a matrix A has n orthogonal eigenvectors, they can (by normalizing) be taken to be orthonormal. The corresponding diagonalizing matrix P has.
  2. [2]
    [PDF] Orthogonally Diagonalizable Matrices
    Orthogonal. T diagonalization gives a new coordinate system in terms of which we can orthogonal. “picture” the transformation clearly. Page 7. How to ...
  3. [3]
    Orthonormal Diagonalization - A First Course in Linear Algebra
    Since D D is diagonal, it just multiplies each entry of a vector by a scalar. Diagonal entries that are positive or negative, with absolute values bigger or ...
  4. [4]
    [PDF] LADR4e.pdf - Linear Algebra Done Right - Sheldon Axler
    ... linear algebra textbook, the proof given here uses linear algebra techniques and makes nice use of a basis of 𝒫𝑛(𝐅), which is the (𝑛 + 1)-dimensional ...
  5. [5]
    [PDF] RES.18-011 (Fall 2021) Lecture 12: Orthogonal Matrices
    Orthogonal matrices are those preserving the dot product. Defnition 12.3 A matrix A ∈ GLn(R) is orthogonal if Av · Aw = v · w for all vectors v and w. In ...
  6. [6]
    Orthogonal Transformations and Orthogonal Matrices - UTSA
    Jan 29, 2022 · Properties. Matrix properties. A real square matrix is orthogonal if and only if its columns form an orthonormal basis of the Euclidean space ...
  7. [7]
    [PDF] orthogonal matrices - UTK Math
    Basic properties. (1) A matrix is orthogonal exactly when its column vectors have length one, and are pairwise orthogonal; likewise for the row vectors. In ...
  8. [8]
    [PDF] orthogonal matrices - UTK Math
    An orthogonal transformation of Rn is a rotation if it has determinant 1, a reflection if it has determinant −1. Examples. 1. Reflection on a given plane in R3.
  9. [9]
    [PDF] Unit 8: The orthogonal group
    Feb 20, 2019 · Examples of orthogonal matrices are rotation matrices and reflection matrices. These two types are the only 2 × 2 matrices which are ...
  10. [10]
    [PDF] Unit 17: Spectral theorem
    17.2. Theorem: Symmetric matrices have only real eigenvalues. 17.3. Theorem: If A is symmetric, then eigenvectors to different eigenvalues are perpendicular.
  11. [11]
    Matrices and determinants - MacTutor History of Mathematics
    Cauchy also introduced the idea of similar matrices (but not the term) and showed that if two matrices are similar they have the same characteristic equation.
  12. [12]
    Cauchy and the spectral theory of matrices - ScienceDirect.com
    View PDF; Download full issue. Search ScienceDirect. Elsevier ... Cauchy and the spectral theory of matrices. Author links open overlay panel. Thomas Hawkins.
  13. [13]
    [PDF] SPECTRAL THEOREM Orthogonal Diagonalizable A diagonal ...
    (Spectral theorem) A ∈ Rn×n is orthogonally diagonalizable if and only if it is symmetric. An important consequence of this is that a symmetric n×n matrix has ( ...<|control11|><|separator|>
  14. [14]
    [PDF] Theorem 3.8.1
    Jun 28, 2020 · A (real) n × n matrix A is orthogonally diagonalizable if and only if A is symmetric. Proof. First, suppose A is orthogonally diagonalizable.<|control11|><|separator|>
  15. [15]
    [PDF] Orthogonally diagonalizable matrix
    An orthogonally diagonalizable matrix is symmetric, and if A is orthogonally diagonalizable, then A has real eigenvalues.
  16. [16]
    Normal matrices and diagonalizability
    Theorem: The product of two unitary matrices is unitary. are simultaneously diagonalizable if and only if they commute. Theorem: A matrix is normal if and only ...
  17. [17]
    Complex Matrices - Ximera - The Ohio State University
    ... orthogonally diagonalizable (that is, is diagonal for some real orthogonal matrix ). ... Show that a real normal matrix is either symmetric or has the form . If ...
  18. [18]
    [PDF] Chapter 13 Applications of SVD and Pseudo-inverses - CIS UPenn
    If A is a (real) normal matrix, then we know from Theo- rem 10.15 that A can be block diagonalized with respect to an orthogonal matrix U as. A = UΛU>, where ...
  19. [19]
    [PDF] Lecture 3.26. Hermitian, unitary and normal matrices - Purdue Math
    Now suppose that we have an orthogonal matrix Q. The eigenvalues are no longer guaranteed to be real, so in general, one cannot diagonalize Q using only real ...
  20. [20]
    [PDF] Eigendecompositions of Hermitian matrices
    Sep 4, 2020 · If A P nˆn is Hermitian, then it is unitarily diagonalizable with real eigenvalues. Hermitian matrices are also called self-adjoint. MATH ...
  21. [21]
    [PDF] Linear Algebra (Math 2890) Solution to Final Review Problems
    (b) Give an example of a 2 × 2 matrix which is diagonalizable but not orthogonally diagonalizable? ... So A is not orthogonally diagonalizable. 7. Page 8. 4 ...
  22. [22]
    [PDF] Proof of the spectral theorem
    Nov 5, 2013 · Proof of Spectral Theorem. Recall that we are proving only that a selfad- joint operator has the orthogonal eigenspace decomposition described.Missing: textbook citation
  23. [23]
    [PDF] Spectral Theorem and Applications - UChicago Math
    Sep 23, 2016 · The Spectral Theorem states that every symmetric transformation in a finite-dimensional Euclidean space has an orthonormal eigenbasis.
  24. [24]
    [PDF] Spectral theorems, SVD, and Quadratic forms
    Theorem. If A is a real symmetric matrix, then its eigenvalues are real, and eigenvectors with different eigenvalues are orthogonal. Proof. Suppose λ is a ...
  25. [25]
    [PDF] Math 2940: Symmetric matrices have real eigenvalues
    The first step of the proof is to show that all the roots of the characteristic polynomial of A (i.e. the eigenvalues of A) are real numbers. Recall that if z = ...
  26. [26]
    [PDF] Note 14: Symmetric Matrices and Spectral Theorem - EECS16B
    Apr 12, 2024 · In this note, we will discuss the Spectral Theorem, which provides us with useful properties of symmetric matrices. 2 Spectral Theorem. Now it ...
  27. [27]
    [PDF] 1 Some Facts on Symmetric Matrices
    Definition: Matrix A is symmetric if A = AT . Theorem: Any symmetric matrix. 1) has only real eigenvalues;. 2) is always diagonalizable;.
  28. [28]
    The Characteristic Polynomial
    Theorem(Eigenvalues are roots of the characteristic polynomial) Let A be an n × n matrix, and let f ( λ )= det ( A − λ I n ) be its characteristic polynomial. ...
  29. [29]
    7.1: Eigenvalues and Eigenvectors of a Matrix
    Mar 27, 2023 · The eigenvectors of a matrix A are those vectors X for which multiplication by A results in a vector in the same direction or opposite direction to X.
  30. [30]
    [PDF] The QR Algorithm - Ethz
    The QR algorithm computes a Schur decomposition of a matrix. It is certainly one of the most important algorithm in eigenvalue computations [9].
  31. [31]
    [PDF] 9. QR algorithm
    algorithms for computing eigenvalues of matrices of order 𝑛 ≥ 5 must be iterative ... • we already noted that matrices 𝐴𝑘 are symmetric if 𝐴 is symmetric (page ...
  32. [32]
    [PDF] 1 Power iteration - CS@Cornell
    Oct 17, 2016 · 5.1 Basic power iteration. 1. % [v,lambda] = power(A, v, maxiter, rtol). 2. %. 3. % Run power iteration to compute the dominant eigenvalue of A ...
  33. [33]
    Power Method - an overview | ScienceDirect Topics
    The Power Method is used to find a dominant eigenvalue (one having the largest absolute value), if one exists, and a corresponding eigenvector. •. To apply the ...
  34. [34]
    [PDF] Chapter 6 Eigenvalues and Eigenvectors
    Multiply an eigenvector by A, and the vector Ax is a number λ times the original x. The basic equation is Ax = λx. The number λ is an eigenvalue of A. The ...
  35. [35]
    [PDF] 1. Orthogonal Diagonalization - KSU Math
    With Fact 1 in hand, start with a symmetric n × n matrix A, pick one eigenvalue λ1 and a λ1-eigenvector x (for A), normalize it, thus producing a unit ...
  36. [36]
    [PDF] Eigenvectors and Diagonalization
    In other words, A can be diagonalized by an orthogonal matrix U: A = UDU−1,. D diagonal and real. (The columns of U are the elements of the orthonormal ...
  37. [37]
    [PDF] 1 Spectral theorem - 1.1 Diagonalization
    We can take the eigenvectors to be of length 1 and the matrix P having the xi as columns will be an orthogonal matrix. The general case where some eigenvalues ...
  38. [38]
    7.1 Symmetric matrices and variance - Understanding Linear Algebra
    Each of the following matrices is symmetric so the Spectral Theorem tells us that each is orthogonally diagonalizable. The point of this activity is to find an ...<|control11|><|separator|>
  39. [39]
    [PDF] Linear Algebra - Chapter 5: Norms, Inner Products and Orthogonality
    Givens rotations can be used as an alternative to Householder reflections to construct a QR factorization. Householder reflections are in general more efficient ...
  40. [40]
    [PDF] Diagonalization of a 2 χ 2 real symmetric matrix
    Diagonalization involves computing eigenvalues and eigenvectors of a 2x2 matrix, then finding a real orthogonal matrix to diagonalize it. The matrix form is S ...
  41. [41]
    [PDF] Statistics-and-PCA | MIT
    Sep 7, 2017 · This process of diagonalizing the covariance matrix is called principal component analysis, or PCA. Let's try it: In [19]: σ2, Q = eig(S).<|control11|><|separator|>
  42. [42]
    [PDF] 18-660: Numerical Methods for Engineering Design and Optimization
    Eigen decomposition. Page 24. Slide 24. Dimension Reduction by PCA. □ Example (continued):. In this case, the 3x3 covariance matrix has a rank of 2. Only 2 ...
  43. [43]
    [PDF] 7 Symmetric Matrices - Berkeley Math
    Theorem 3 (The Spectral Theorem for Symmetric Matrices). An n×n symmetric matrix A has the following properties: 1. A has n real eigenvalues, counting ...
  44. [44]
    Eigenvalues of real symmetric matrices
    Orthogonal matrix. Real symmetric matrices not only have real eigenvalues, they are always diagonalizable. In fact, more can be said about the diagonalization.
  45. [45]
  46. [46]
    [PDF] A note on simultaneously diagonalizable matrices - Ele-Math
    Recall that a set of real symmetric matrices can be simultaneously diagonalized by an orthogonal matrix if and only if they are pairwise commutative, e.g., [5], ...Missing: uniqueness | Show results with:uniqueness
  47. [47]
    [PDF] 8.3 Positive Definite Matrices
    A square matrix is called positive definite if it is symmetric and all its eigenvalues λ are positive, that is λ > 0. Because these matrices are symmetric, the ...
  48. [48]
    [PDF] Test for Positive and Negative Definiteness
    for A is positive definite if and only if all the eigenvalues are positive. Since, det(A) = λ1λ2, it is necessary that the determinant of A be positive.
  49. [49]
    [PDF] Analyzing the Hessian
    If the Hessian at a given point has all positive eigenvalues, it is said to be a positive-definite matrix. This is the multivariable equivalent of “concave up”.
  50. [50]
    [PDF] 7.2 Positive Definite Matrices and the SVD - MIT Mathematics
    The entries in the diagonal matrix † are the square roots of the eigenvalues. The matrices AAT and ATA have the same nonzero eigenvalues.
  51. [51]
    [PDF] A TUTORIAL ON PRINCIPAL COMPONENT ANALYSIS
    Diagonalizing a covariance matrix might not pro- duce satisfactory results. The most rigorous form of removing redundancy is statistical independence. P(y1 ...
  52. [52]
    [PDF] Quantum Theory I, Lecture 3 Notes - MIT OpenCourseWare
    We leave the proof as an exercise. 3.1.3 Diagonalization of Hermitian Operators. Theorem 2. A Hermitian matrix Hij = <φi|H|φj> can always be diagonalized by a ...
  53. [53]
    [PDF] Spectral Theorems for Hermitian and unitary matrices - Purdue Math
    For an Hermitian matrix: a) all eigenvalues are real, b) eigenvectors corresponding to distinct eigenvalues are orthogonal, c) there exists an orthogonal basis ...
  54. [54]
    [PDF] Numerical Solutions of the Schrödinger Equation - Physics
    This is the Schrödinger equation in the k-basis, written more compactly as. HC = EC. (42). Solving it corresponds to diagonalizing the Hamiltonian matrix H. 9 ...
  55. [55]
    [PDF] Lecture 10: Solving the Time-Independent Schrödinger Equation
    Lecture 10: Solving the Time-Independent Schrödinger Equation ... where we displayed the form of the Hamiltonian operator H with the time independent potential.
  56. [56]
    [PDF] Lecture 8: Normal Modes - LIGO-Labcit Home
    Normal mode analysis makes it easy to understand the complicated motion of individual oscillators when they are coupled to others. (Although vibrations of ...
  57. [57]
    [PDF] On the normal modes of coupled harmonic oscillators - arXiv
    Nov 20, 2023 · It is clear that we can always diagonalize the kinetic and potential energies for a system of coupled harmonic oscillators as shown in equation ...
  58. [58]
    [PDF] 3 Coupled Harmonic Oscillators, Normal Modes - Xie Chen
    As both M1/2 and K are symmetric real matrices, M−1/2KM−1/2 is real symmetric and there are n solutions with orthogonal eigenvectors.Missing: diagonalization | Show results with:diagonalization