Fact-checked by Grok 2 weeks ago

Polar decomposition

In linear algebra, the polar decomposition is a fundamental theorem that factorizes any complex matrix A \in \mathbb{C}^{m \times n} as A = UH, where U is a partial (unitary in the case of square invertible matrices) and H is a positive semi-definite uniquely determined as the of A^H A. This decomposition separates the "rotational" or component U from the "stretching" or positive scaling component H, providing insight into the geometric action of the linear transformation represented by A. For real matrices, the decomposition takes the form A = QP, with Q orthogonal and P positive semi-definite symmetric. The guarantees the existence and of this under appropriate conditions, such as when the of A equals the dimension of the domain for of U. It is closely related to the (), from which the polar decomposition can be derived by setting U = P Q^H and H = Q \Sigma Q^H for A = P \Sigma Q^H, where \Sigma is diagonal with non-negative entries. Originally introduced by Léon Autonne in for matrices and later generalized, the polar decomposition has roots in early 20th-century work on matrix canonical forms and has been extensively developed in for Hilbert spaces. Beyond , the polar decomposition finds applications in for orthogonalization and matrix sign functions, in for quantum algorithms and , and in for decomposing deformation tensors into stretch and rotation parts. Computationally, it can be obtained via iterative methods like the Newton-Schulz or directly from the , with ensured in finite-precision arithmetic.

Core Concepts

Definition

In linear algebra, for square matrices, the polar decomposition factorizes a matrix into a product involving a unitary matrix and a positive semi-definite matrix, analogous to separating rotational and scaling components of a linear transformation. For any A \in \mathbb{C}^{n \times n}, the right polar decomposition expresses A as A = U P, where U is a satisfying U^* U = I_n (with U^* denoting the and I_n the ), and P is a positive semi-definite . Here, positive semi-definite means P = P^* (Hermitian) with all eigenvalues non-negative, and P is explicitly the unique positive semi-definite of A^* A, denoted P = (A^* A)^{1/2}, satisfying P^2 = A^* A. The left polar decomposition takes the form A = P' U', where P' is the unique positive semi-definite Hermitian square root of A A^*, so P' = (A A^*)^{1/2} with (P')^2 = A A^*, and U' is unitary. If A is invertible, both the right and left decompositions are unique.

Geometric Interpretation

The polar decomposition of a linear transformation represented by a matrix A \in \mathbb{R}^{n \times n} separates it into a product A = UP, where U is an orthogonal matrix encoding a rotation or reflection, and P is a positive semi-definite symmetric matrix representing a stretching or compression along principal axes. In the two-dimensional case, for a $2 \times 2 matrix, this decomposition intuitively captures how A distorts shapes by first applying an elliptical stretch via P and then a rigid rotation or reflection via U. A key arises when considering on the unit circle in \mathbb{R}^2. The positive part P transforms the unit circle into an , with the aligned to the eigenvectors of P and lengths given by its eigenvalues (the principal stretches). The unitary part U then rotates or reflects this to its final position, matching the of the unit circle under A. This process highlights the decomposition's role in isolating distortive and rigid components of the transformation. For a concrete example, consider the matrix A = \begin{pmatrix} 2 & 0 \\ 0 & 1 \end{pmatrix}, which stretches vectors along the x-axis by a factor of 2 while leaving the y-axis unchanged. Its polar decomposition is A = I \cdot P, where U = I (, implying no ) and P = A itself, as P is already positive semi-definite. Geometrically, this simply elongates the unit circle into an with axes of lengths 2 and 1, without further reorientation. In three dimensions, the interpretation extends analogously: U performs an , such as a around an or a , preserving volumes and angles, while P applies a positive semi-definite scaling that deforms a into an (degenerate if singular) along its principal axes. The unitary component U acts as an , maintaining lengths of vectors, whereas P encodes the singular values of A as the stretch factors along orthogonal directions, quantifying the transformation's distortive effect independent of orientation.

Algebraic Properties

General Properties

In the polar decomposition A = U P of a square complex matrix A, where U is unitary and P is positive semi-definite Hermitian, the factor U is isometric and thus has operator norm \|U\| = 1. Since \|A x\| = \|U (P x)\| = \|P x\| for all x, it follows that the operator norm satisfies \|A\| = \|P\|. The eigenvalues of P coincide with the singular values of A, as P = \sqrt{A^* A} and the singular values are the square roots of the eigenvalues of A^* A. The eigenvalues of the unitary factor U lie on the unit circle in the complex plane. The Frobenius norm of A relates directly to P via \|A\|_F^2 = \operatorname{trace}(A^* A) = \operatorname{trace}(P^2), since A^* A = P^2. A matrix A is invertible if and only if its positive factor P is invertible, as U is always invertible; in this case, the inverse is given by A^{-1} = P^{-1} U^*. If A = U P and B = V Q are polar decompositions, the polar decomposition of the product AB does not take the simple form U V \cdot P Q, but instead has positive factor \sqrt{B^* P^2 B}.

Relation to SVD

The (SVD) provides an explicit means to construct the polar decomposition of a A \in \mathbb{C}^{m \times n}, expressed as A = W \Sigma V^*, where W \in \mathbb{C}^{m \times m} and V \in \mathbb{C}^{n \times n} are unitary matrices, and \Sigma \in \mathbb{R}^{m \times n} is a rectangular with non-negative real singular values \sigma_1 \geq \sigma_2 \geq \cdots \geq \sigma_r > 0 on the diagonal (with r = \operatorname{rank}(A)) and zeros elsewhere. From this SVD, the right polar decomposition A = U P follows by setting U = W V^* (a partial with U^* U being the orthogonal projection onto the range of V) and P = V \Sigma V^* (a positive semi-definite ), yielding U P = W V^* V \Sigma V^* = W \Sigma V^* = A. The singular values of A are precisely the eigenvalues of P. Similarly, the left polar decomposition A = P' U' is obtained with P' = W \Sigma W^* (positive semi-definite ) and U' = W V^* (partial ), satisfying P' U' = W \Sigma W^* W V^* = W \Sigma V^* = A. For real matrices A \in \mathbb{R}^{m \times n}, the SVD takes the form A = W \Sigma V^T with orthogonal W and V, leading to analogous real polar forms where U and U' are orthogonal (or partial isometries) and P, P' are symmetric positive semi-definite; the inherent sign ambiguities in the columns of W and V allow choices that align phases to ensure properties like positive in the orthogonal factors when desired. Computationally, these relations imply that the polar decomposition arises as a direct byproduct of the via simple multiplications, whereas extracting the from a known polar form necessitates the eigendecomposition of the positive semi-definite factor, which is comparably intensive to itself.

Relation to Normal Matrices

A normal matrix A satisfies A A^* = A^* A, where A^* denotes the () of A. In the polar decomposition A = U P, with U unitary and P positive semidefinite, A is if and only if U and P commute, i.e., U P = P U. This commutativity implies that there exists a unitary matrix Q that simultaneously diagonalizes both U and P, so A is unitarily similar to a diagonal matrix via the same basis. Since P^2 = A^* A is positive semidefinite (hence ) and A is , the polar factors share a common eigenbasis. By the for matrices, A = Q \Lambda Q^*, where Q is unitary and \Lambda = \operatorname{diag}(\lambda_1, \dots, \lambda_n) contains the eigenvalues of A. The positive part is then P = Q |\Lambda| Q^*, with |\Lambda| = \operatorname{diag}(|\lambda_1|, \dots, |\lambda_n|), and the unitary part is U = Q (\Lambda / |\Lambda|) Q^*, where \Lambda / |\Lambda| = \operatorname{diag}(e^{i \arg(\lambda_1)}, \dots, e^{i \arg(\lambda_n)}). For example, consider the matrix with eigenvalues $2e^{i\pi/4} and $3e^{i\pi/3}; then P has diagonal entries 2 and 3, while U has e^{i\pi/4} and e^{i\pi/3} in the shared eigenbasis. This structure aligns the polar decomposition directly with the , as the eigenvalues of A determine both the singular values (of P) and the phases (of U).

Existence and Construction

For Normal Operators

For normal operators on a , the facilitates both the existence and explicit construction of the polar decomposition. Specifically, a bounded A on a \mathcal{H} admits a given by A = \int_{\sigma(A)} \lambda \, dE(\lambda), where \sigma(A) denotes the of A and E is the unique spectral measure (resolution of the identity) such that E(\cdot) is a on the Borel subsets of \mathbb{C} with support in \sigma(A). The positive part P = |A| of the decomposition is constructed via as P = \int_{\sigma(A)} |\lambda| \, dE(\lambda). To see that P serves as the , note that the A^* A has A^* A = \int_{\sigma(A)} |\lambda|^2 \, dE(\lambda), and since the function f(t) = \sqrt{t} (for t \geq 0) is continuous and positive, the yields P as the unique positive of A^* A, ensuring P is positive and . The partial isometry U is then defined by U = \int_{\sigma(A) \setminus \{0\}} \frac{\lambda}{|\lambda|} \, dE(\lambda) on the closure of the range of P, with U extended by zero on the kernel of P (which corresponds to the support of E(\{0\})). This ensures U^* U = I on the range of P and U U^* = I on the range of A, verifying that U is a partial isometry with initial space \overline{\operatorname{ran} P} and final space \overline{\operatorname{ran} A}. The product A = U P follows directly from the multiplicativity of the functional calculus, as U P = \left( \int_{\sigma(A) \setminus \{0\}} \frac{\lambda}{|\lambda|} \, dE(\lambda) \right) \left( \int_{\sigma(A)} |\lambda| \, dE(\lambda) \right) = \int_{\sigma(A)} \lambda \, dE(\lambda) = A on the range, with both sides vanishing on the kernel. The uniqueness of this polar decomposition for normal operators stems from the uniqueness of the spectral measure E in the , which determines both P and U uniquely via the above integrals. Any other decomposition A = V Q with Q positive and V a partial satisfying the range conditions would yield the same Q = |A| (as the unique positive ) and thus the same V = A Q^{-1} (where defined), aligning with the construction. In the finite-dimensional case, where A is a on \mathbb{C}^n, the reduces to unitary A = W D W^* with D diagonal containing the eigenvalues \lambda_i. The polar decomposition then aligns with P = W \operatorname{diag}(|\lambda_1|, \dots, |\lambda_n|) W^* and U = W \operatorname{diag}(e^{i \arg \lambda_1}, \dots, e^{i \arg \lambda_k}, 0, \dots, 0) W^*, where the first k entries correspond to the non-zero eigenvalues.

For Invertible Matrices

For an A \in \mathbb{C}^{n \times n}, the A^* A is positive definite Hermitian, ensuring that the principal matrix logarithm \log(A^* A) exists and is Hermitian, as the eigenvalues of A^* A are . This setup allows a constructive approach to the polar decomposition via the matrix exponential, which maps Hermitian matrices to positive definite ones. The positive definite factor P is constructed as P = \exp\left(\frac{1}{2} \log(A^* A)\right), which coincides with the unique positive definite (A^* A)^{1/2}. The unitary factor is then U = A P^{-1}. This yields the right polar decomposition A = U P, where P is positive definite and U is unitary. To verify unitarity, note that U^* = P^{-1} A^*, so U^* U = P^{-1} A^* A P^{-1}. Substituting A^* A = P^2 gives U^* U = P^{-1} P^2 P^{-1} = I. Similarly, U U^* = I follows by symmetry or direct computation. In the real case, for invertible A \in \mathbb{R}^{n \times n}, A^T A is positive definite symmetric, so \log(A^T A) is real symmetric and P = \exp\left(\frac{1}{2} \log(A^T A)\right) is positive definite symmetric, yielding an orthogonal U = A P^{-1} with U^T U = I. The construction ensures orthogonality. The sign of \det U is determined by \det A, as \det P > 0. An alternative construction in the real case parametrizes the orthogonal factor U via the , where skew-symmetric matrices are mapped to orthogonal ones without eigenvalue -1, providing a rational useful for avoiding logarithmic computations in certain numerical contexts.

General Proof

One approach to proving the existence of the polar decomposition for an arbitrary square complex A \in \mathbb{C}^{n \times n}, possibly singular, relies on arguments from the invertible case. Approximate A by the invertible A_\epsilon = A + \epsilon I for \epsilon > 0, which admits a polar decomposition A_\epsilon = U_\epsilon P_\epsilon where U_\epsilon is unitary and P_\epsilon is positive definite. As \epsilon \to 0^+, the sequence \{P_\epsilon\} converges in the to a positive semi-definite P = \lim_{\epsilon \to 0^+} P_\epsilon = (A^* A)^{1/2}, the unique positive semi-definite of A^* A. Similarly, \{U_\epsilon\} converges to a partial U such that U is unitary on the range of A (equivalently, the range of P) and U P = A, with U acting as zero on the of A. This limit yields the polar decomposition A = U P, as the ensures the factors remain bounded and the decomposition is continuous in the space of matrices. An alternative proof constructs the decomposition directly using the () of A, which exists for any . The states that A = V \Sigma W^* where V and W are unitary matrices and \Sigma is diagonal with non-negative real entries (the singular values of A). Define P = W \Sigma W^*, which is positive semi-definite, and U = V W^*, which is unitary. Then A = U P, establishing the polar decomposition without relying on limits. This construction highlights that the proof is non-constructive in the sense that it invokes the existence of the , whose proof itself requires for Hermitian matrices. The polar decomposition A = U P is unique in the following sense: the positive semi-definite factor P is uniquely determined as the positive of A^* A. The unitary factor U is unique on the range of P (or equivalently, the range of A), where it acts as an mapping the range of P onto the range of A; on the of the range of P, U may be chosen arbitrarily as long as it remains unitary overall. The construction preserves rank, with \operatorname{rank}(A) = \operatorname{rank}(P), since the non-zero singular values of A correspond exactly to the positive eigenvalues of P, and the kernel dimensions match via the partial isometry property of U.

Generalizations

Bounded Operators on Hilbert Spaces

In the context of bounded linear operators on a H, the polar decomposition generalizes the finite-dimensional matrix case by factoring a bounded operator A \in B(H) as A = U |A|, where |A| = \sqrt{A^* A} is the unique positive square root of the positive self-adjoint operator A^* A, and U is a partial isometry with initial projection onto the closure of the range of |A|. The partial isometry U satisfies U^* U = P, the support projection of |A|, and is unitary on the subspace \overline{\operatorname{ran}(|A|)}. The existence of this decomposition follows from the structure of B(H), where A^* A is positive , and the guarantees a unique positive |A|. The partial isometry U is then constructed by extending an from \overline{\operatorname{ran}(|A|)} to \overline{\operatorname{ran}(A)} and setting it to zero on the . This ensures U preserves norms on its initial space and aligns the ranges appropriately. Uniqueness holds for |A| due to the spectral theorem, and for U on \operatorname{ran}(A), with U^* U equaling the support projection of |A|. Key properties include the equality of operator norms \|A\| = \|U\| = \||A\|\|, since U is an isometry on the relevant subspace. Spectrum relations from the matrix case extend, with the spectrum of A related to that of |A| via the partial isometry action, adjusted for possible kernel and cokernel effects in infinite dimensions. A representative example is the unilateral right S on \ell^2(\mathbb{N}), defined by S(e_n) = e_{n+1} for the \{e_n\}, which is a bounded with S^* S = I but S S^* \neq I. Here, |S| = I, so the polar decomposition is S = S \cdot I, where S itself serves as the partial isometry, with initial projection I and final projection I - |e_1\rangle\langle e_1|.

Unbounded Operators

In the context of unbounded s on a \mathcal{H}, the polar decomposition applies to closed densely defined linear s A: \mathcal{D}(A) \subseteq \mathcal{H} \to \mathcal{H}, where \mathcal{D}(A) is dense in \mathcal{H}. Such an admits a unique A = U |A|, with |A| = (A^* A)^{1/2} denoting the positive self-adjoint of the of A^* A, defined on the domain \mathcal{D}(|A|) = \mathcal{D}(A). Here, U is a partial (bounded linear ) satisfying \ker U = \ker A, with initial space \overline{\mathrm{ran}(|A|)} and final space \overline{\mathrm{ran}(A)}, such that U^* U is the orthogonal projection onto \overline{\mathrm{ran}(|A|)} and U U^* is the orthogonal projection onto \overline{\mathrm{ran}(A)}. The construction relies on the spectral theorem for the self-adjoint operator |A|, which ensures the existence of a unique positive self-adjoint extension. Specifically, since A is closed and densely defined, A^* exists and is closed, and the symmetric operator A^* A (initially defined on \mathcal{D}(A^* A) = \mathcal{D}(A) \cap \mathcal{D}(A^*)) has a self-adjoint closure, allowing |A| to be well-defined via functional calculus. The partial isometry U is then constructed by extending a densely defined operator from \mathrm{ran}(|A|^{1/2}) to the whole space: for x \in \mathcal{D}(A), set U |A|^{1/2} y = A (|A|^{-1/2} y) where y \in \mathcal{D}(|A|^{1/2}) with \mathcal{D}(A) \subseteq \mathcal{D}(|A|^{1/2}), and this extends uniquely to a bounded partial isometry on \mathcal{H}. The equality A = U |A| holds precisely on \mathcal{D}(A). Existence follows directly from the closedness of A, which implies that the closure of A^* A is and positive, guaranteeing the for |A|; the partial isometry U then extends uniquely due to the closed graph theorem applied to the graph of A. However, significant challenges arise with domains: while \mathcal{D}(|A|) = \mathcal{D}(A), the domain of U is the entire \mathcal{H}, so the product U |A| is only equal to A on \mathcal{D}(A) and may differ elsewhere, reflecting the unbounded growth of |A| outside this domain. In , this manifests in operators like the P = -i \frac{d}{dx} on L^2(\mathbb{R}) with domain the H^1(\mathbb{R}); its polar decomposition is P = U |P|, where in the Fourier basis, |P| multiplies by |k| and U by \mathrm{sign}(k), both preserving the domain H^1(\mathbb{R}), but illustrating how spectral measures handle unbounded spectra while respecting domain restrictions for physical observables.

Quaternion Polar Decomposition

The quaternion algebra \mathbb{H} is a four-dimensional associative division algebra over the real numbers, spanned by the basis \{1, i, j, k\} with multiplication rules i^2 = j^2 = k^2 = -1, ij = k, jk = i, and ki = j. Matrices over \mathbb{H}, denoted M_n(\mathbb{H}), are n \times n arrays with entries in \mathbb{H}, equipped with the standard Hermitian inner product \langle x, y \rangle = y^* x, where ^* denotes the conjugate transpose and the quaternion conjugate of q = a + bi + cj + dk is \overline{q} = a - bi - cj - dk. The polar decomposition for a matrix A \in M_n(\mathbb{H}) expresses A = UP, where U \in M_n(\mathbb{H}) is unitary (U^* U = I) and P \in M_n(\mathbb{H}) is Hermitian (P = P^* and x^* P x \geq 0 for all x \in \mathbb{H}^n, with equality to zero only if x = 0 in the invertible case). This decomposition exists for all A \in M_n(\mathbb{H}), including non-invertible matrices, and generalizes the complex case while accounting for the non-commutativity of \mathbb{H}. The factor is constructed as P = \sqrt{A^* A}, where the is defined via the for Hermitian quaternion matrices, which admit a A^* A = V D V^* with V unitary and D real diagonal with non-negative entries, yielding \sqrt{A^* A} = V \sqrt{D} V^* using the non-negative real on the . The unitary factor is then U = A P^\dagger for the Moore-Penrose pseudoinverse P^\dagger in the general case, or U = A P^{-1} when A is invertible; this ensures U is a partial with initial space matching the range of P. Uniqueness holds for the factor P in the minimal polar decomposition (where the kernel of U is minimized), analogous to the complex setting, but non-commutativity introduces distinctions between right polar decompositions (A = UP) and left variants (A = PU'), with the standard form being right-sided. For invertible A, both U and P are unique. A simple example is the $1 \times 1 case, where any nonzero quaternion q \in \mathbb{H} decomposes as q = |q| \cdot u with |q| = \sqrt{q \overline{q}} \geq 0 the real modulus (positive semidefinite) and u = q / |q| a unit quaternion (\overline{u} u = 1), representing a scaling by a positive real followed by a rotation in 3D space.

Alternative Planar Decompositions

In the planar case, where A \in \mathbb{R}^{2 \times 2} is an , the polar decomposition A = UP separates A into an orthogonal factor U \in O(2) (with \det(U) = \pm 1) and a symmetric positive definite factor P, capturing or followed by a stretch along principal axes. Alternative factorizations exist that attempt similar separations into a rotational component and a or component, but they differ in structure and geometric interpretation. For instance, the expresses A = QR, where Q \in O(2) is orthogonal and R is upper triangular with positive diagonal entries, effectively combining anisotropic scaling along coordinate axes with shear. This contrasts with polar decomposition, as R is generally not symmetric, mixing shear and scaling effects rather than isolating pure stretch. A key distinction arises when reflections are involved: the polar decomposition accommodates \det(U) = -1 naturally, preserving the symmetric positive structure of P, whereas alternatives restricted to proper rotations in SO(2) (with \det = 1) would require adjusting the second factor, often resulting in a non-symmetric or non-positive component that loses the clean separation of orthogonal and stretch parts. The polar decomposition's uniqueness stems from its reliance on the symmetric positive square root of A^\top A, ensuring P aligns stretches with orthogonal principal directions invariant to coordinate rotations, a property not shared by basis-dependent alternatives like QR. For example, consider the simple shear matrix A = \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}. Its polar decomposition yields P = \sqrt{A^\top A} = \begin{pmatrix} \frac{2}{\sqrt{5}} & \frac{1}{\sqrt{5}} \\ \frac{1}{\sqrt{5}} & \frac{3}{\sqrt{5}} \end{pmatrix} (symmetric with eigenvalues \frac{1 + \sqrt{5}}{2} and \frac{\sqrt{5} - 1}{2}) and U = A P^{-1}, a rotation by about 26.565 degrees; in contrast, the QR form is A = I \cdot A, where the identity is orthogonal but the upper triangular A embeds both shear and uniform scaling without distinguishing them. This highlights polar's preference for geometric applications, as its factors remain invariant under orthogonal transformations of the basis, providing a coordinate-free interpretation of deformation into rigid motion and strain.

Connections to Other Factorizations

The polar decomposition of a square invertible matrix A ∈ ℝ^{n×n} is closely related to the QR decomposition A = Q R, where Q is orthogonal and R is upper triangular with positive diagonal entries. The positive factor P in the polar decomposition A = U P is the unique symmetric positive definite square root of A^T A, while R is the upper Cholesky factor of A^T A (i.e., A^T A = R^T R). Thus, P and R are linked through the symmetric structure of A^T A, but P is symmetric whereas R is triangular; the unitary factor U = A P^{-1} differs from Q = A R^{-1} by the transformation R P^{-1}, which is generally not the identity unless A is orthogonal. This connection allows QR decompositions to be used in iterative algorithms for computing the polar decomposition, such as Newton's method, where each step orthogonalizes an approximate unitary factor via QR to ensure stability. For normal matrices, the polar decomposition aligns directly with the eigendecomposition. A normal matrix A satisfies A A^* = A^* A and admits an eigendecomposition A = V D V^, where V is unitary and D is diagonal with complex entries. The polar factors are then P = V |D| V^ (positive semidefinite, with |D| the diagonal of absolute values) and U = V (D / |D|) V^* (unitary, capturing the phases or signs in D / |D|). This separation isolates the magnitude (|λ_i|) in P and the directional component (e^{i arg(λ_i)}) in U, providing a canonical form that extends the eigendecomposition to non-normal cases via the more general polar structure; for non-normal matrices, the Jordan canonical form complicates direct alignment, as it involves non-diagonalizable blocks not separable into unitary and positive parts in the same way. Compared to the Schur decomposition A = Q T Q^*, where Q is unitary and T is upper triangular (revealing eigenvalues on the diagonal of T), the polar decomposition offers advantages in preserving for the positive component. The Schur "stretch" T is triangular and generally non-symmetric, which can obscure applications requiring symmetric positive operators (e.g., modeling or definitions); in contrast, the polar P is explicitly symmetric positive definite, facilitating better structure exploitation in optimization and approximation tasks without additional symmetrization steps.

Numerical Computation

Algorithms for Matrices

One standard direct method for computing the polar decomposition of a matrix relies on its singular value decomposition (SVD). For a complex matrix A \in \mathbb{C}^{m \times n} with m \geq n, compute the thin SVD A = U \Sigma V^*, where U \in \mathbb{C}^{m \times n} has orthonormal columns, \Sigma \in \mathbb{R}^{n \times n} is diagonal with non-negative singular values, and V \in \mathbb{C}^{n \times n} is unitary. The right polar decomposition is then given by A = UP, where the positive semidefinite factor is P = V \Sigma V^* and the partial isometry (unitary if full rank) is U = U V^*. For the general rectangular case or m < n, the construction uses the appropriate thin SVD and pseudoinverse of P to ensure U = A P^\dagger. For the 2×2 case, closed-form expressions exist for nonsingular real matrices, leveraging traces and determinants to avoid full SVD computation. The positive semidefinite factor P = \sqrt{A^T A} can be obtained explicitly: let \tau = \operatorname{tr}(A^T A) and \delta = \det(A^T A); then P = \frac{A^T A + \sqrt{\delta} \, I}{\sqrt{\tau + 2 \sqrt{\delta}}}, assuming \tau^2 \neq 4\delta and positive definiteness. The orthogonal factor follows as U = A P^{-1}. These formulas arise from the explicit square root computation for 2×2 positive definite matrices via the . When requiring the orthogonal factor U to be a proper rotation (i.e., \det(U) = 1 for real matrices with \det(A) > 0), adjust the signs of columns in the SVD singular vectors: specifically, if \det(U V^T) = -1, negate one column of V and the corresponding column of U to flip the determinant without altering the product U \Sigma V^*. This ensures compatibility with the special orthogonal group SO(n). The of the SVD-based approach is O(n^3) for an n \times n using standard algorithms like bidiagonalization followed by QR iterations, making it suitable for moderate-sized problems. This method is backward stable, preserving relative accuracy even for highly ill-conditioned matrices where condition numbers exceed machine precision. For singular matrices, the extends naturally via : P has rank equal to that of A, with zero singular values set to zero, and U acts as a partial mapping the range of P unitarily onto the of A. Compute U using the pseudoinverse P^\dagger restricted to the range of P, ensuring U = A P^\dagger on that .

Iterative Methods and Convergence

Iterative methods for computing the polar decomposition of a matrix A typically approximate the unitary factor U and positive semidefinite factor P through successive refinements, avoiding direct computation of square roots or singular value decompositions, which can be expensive for large or ill-conditioned matrices. These methods are particularly useful for large-scale problems where direct methods are impractical, and they often exhibit near the solution under suitable scaling. One prominent approach is the method applied to the matrix function, which targets the unitary polar factor U = \sign(A), defined via the polar decomposition A = U P where P = (A^* A)^{1/2}. The unscaled Newton iteration for the is given by X_{k+1} = \frac{1}{2} (X_k + X_k^{-1}), \quad X_0 = A, which converges to U for full-rank A, with convergence in a neighborhood of the solution provided the condition is satisfied. To ensure global and numerical , scaling is introduced, yielding the scaled variant X_{k+1} = \frac{1}{2} (\mu_k X_k + \mu_k^{-1} X_k^{-*}), where \mu_k is chosen based on norms of X_k and its inverse, such as \mu_k = (\|X_k\|_2 \|X_k^{-1}\|_2)^{1/4}; this requires at most n iterations for an n \times n with optimal scaling and exhibits backward when inverses are computed accurately via bidiagonalization or QR . is locally for full-rank matrices and linear globally, with the radius depending on \|A\|_2 < \sqrt{3} for the unscaled case to avoid divergence. A fixed-point iteration variant focuses on the positive polar factor P, derived from Newton's method for solving P^2 = A^* A. The iteration is P_{k+1} = \frac{1}{2} (P_k + (A^* A) P_k^{-1}), with P_0 an initial positive approximation (e.g., scaled identity); this converges linearly to P with rate depending on the condition number \kappa_2(A), and quadratic near P if started sufficiently close. Scaling by a factor \alpha > \|A^* A\|_2 / 2 ensures convergence within the unit disk, but the method can suffer instability without reformulation, motivating coupled iterations. Once P is obtained, U = A P^{-1} follows directly. The Denman-Beavers iteration provides a practical, alternative for simultaneous computation of U and P, reformulating the square root iteration as a coupled pair to avoid explicit inverses in unstable forms: Y_{k+1} = \frac{1}{2} (Y_k + Z_k^{-1}), \quad Z_{k+1} = \frac{1}{2} (Z_k + Y_k^{-1}), initialized with Y_0 = [A^* A](/page/A^* A), Z_0 = I, converging to P = Y_\infty and P^{-1} = Z_\infty; then U = A Z_\infty. This achieves quadratic convergence for \|Y_0\|_2 < 1 after scaling and is numerically , preserving , particularly for normal matrices where the factors align closely with the . Error analysis shows backward stability with relative error bounded by O(\epsilon \kappa_2(A)), where \epsilon is machine precision, avoiding direct computations that can amplify rounding errors. Recent advancements include block-coordinate methods, such as Riemannian block , which optimize over blocks of the unitary factor on the to compute sparse or structured polar decompositions efficiently for large-scale data; these exhibit linear under restricted assumptions and are suitable for high-dimensional applications like tensor approximation. Overall, these iterations offer backward stable approximations with rates scaling favorably for well-conditioned problems, though ill-conditioning requires careful scaling to maintain behavior.

References

  1. [1]
    [PDF] Polar decomposition - MIT Mathematics
    Nov 25, 2019 · Here is the polar decomposition I stated in class. Theorem 1.3. In the setting of (1.1), there is a factorization. T = SP. (S ∈ L(V ...
  2. [2]
    [PDF] The polar decomposition— properties, applications and algorithms
    The polar decomposition was introduced by Autonne [1] in 1902. A thor- ougłi cliscussion of the history of it is given in Horn and Johnson [29, Sect. 3.0] ...Missing: theorem origin
  3. [3]
    [PDF] Lecture 7.4: Polar decomposition
    Theorem. Every linear map A: X → U can be written as A = UP where P ≥ 0 and U is unitary. This is called the (left) polar decomposition of A.
  4. [4]
    [PDF] The Matrix Sign Decomposition and Its Relation to the Polar ...
    Most results for the matrix sign decomposition have a counterpart for the polar decomposition A = UH, and vice versa. To illustrate this, we derive best ...<|control11|><|separator|>
  5. [5]
  6. [6]
    [PDF] Polar Decomposition and the Closest Rotation - Julian Panetta
    Feb 23, 2015 · A square n × n complex matrix A has two polar decompositions: the standard right polar decomposition. A = UP and the left polar decomposition ...
  7. [7]
    [PDF] Introduction to Continuum Mechanics
    Next, a geometric interpretation is obtained for the polar decomposition decomposition, starting with the right polar decomposition (3.86)1. To this end ...
  8. [8]
    [PDF] Polar Decomposition of a Matrix
    Apr 28, 2014 · We know from Theorem OD in a First Course in Linear Algebra that our matrices P and H are both orthonormally diagonalizable.
  9. [9]
    [PDF] 2.4.1 Stretch Vector and Stretch - ME338A CONTINUUM MECHANICS
    Apr 22, 2008 · The polar decomposition theorem states that any non-singular, ... This makes the geometric interpretation of the polar decom- position ...
  10. [10]
    [PDF] 3·Singular Value Decomposition
    Apr 1, 2017 · The singular value decomposition also gives an immediate formula for the 2-norm of a matrix. kAxk kxk = s1.
  11. [11]
    [PDF] Chapter 12 Singular Value Decomposition and Polar Form
    Given an SVD decomposition A = VDU>, let. R = V U> and S = UDU>. It is clear that R is orthogonal and that S is positive semidefinite symmetric, and. RS = V U> ...
  12. [12]
  13. [13]
    [PDF] Functional Analysis: Spectral Theory
    Feb 5, 2012 · Functional Analysis: Spectral Theory. V.S. Sunder. Institute of Mathematical Sciences. Madras 600113. INDIA. July 31, 2000. Page 2. i ha ha ...Missing: textbook | Show results with:textbook
  14. [14]
    [PDF] Functional Analysis Princeton University MAT520 Lecture Notes
    Aug 18, 2023 · 10 The spectral theorem for bounded normal operators. 110. 10.1 ... Functional analysis is the branch of mathematics that is obtained ...
  15. [15]
    [PDF] Chapter 1 Hilbert space and linear operators
    Obviously, the following definitions and results are also valid for bounded linear operators. ... Theorem 1.6.5 (Polar decomposition). For any B ∈ B(H) there ...
  16. [16]
    Perturbation Theory for Linear Operators - SpringerLink
    Free delivery 14-day returnsPerturbation Theory for Linear Operators · Reviews. "The monograph by T. Kato is an excellent textbook in the theory of linear operators in Banach and Hilbert ...
  17. [17]
    Linear Operators in Hilbert Spaces - SpringerLink
    The purpose of this book is to give an introduction to the theory of linear operators on Hilbert spaces and then to proceed to the interesting applica tions of ...
  18. [18]
  19. [19]
    [PDF] Matrix Animation and Polar Decomposition - Graphics Interface
    This paper presents a method for decomposing matrices, focusing on rotation extraction using Polar Decomposition, which is useful for animation and ...
  20. [20]
    Illustration of polar decomposition | University of Utah CSM Group
    Feb 14, 2013 · This posting explains the meaning of a polar decomposition, and it gives two numerical methods for computing it.
  21. [21]
    Computing the Polar Decomposition—with Applications - SIAM.org
    Computing the Singular Value Decomposition of a Product of Two Matrices ... A quadratically convergent Newton method for computing the polar decomposition of a ...
  22. [22]
    [PDF] Stable and Efficient Computation of Generalized Polar ... - arXiv
    Apr 14, 2021 · Abstract: We present methods for computing the generalized polar decomposition of a matrix based on the dynamically weighted Halley (DWH) ...
  23. [23]
    [PDF] arXiv:2011.12449v1 [math.NA] 25 Nov 2020
    Nov 25, 2020 · ric eigendecomposition, singular value decomposition, polar decomposition, and CS ... Let B ∈ Cm×m be a nonsingular normal matrix. Let ...
  24. [24]
    The Canonical Generalized Polar Decomposition - SIAM.org
    We present methods for computing the generalized polar decomposition of a matrix based on the dynamically weighted Halley iteration. This method is well ...
  25. [25]
  26. [26]
    [PDF] A Riemannian Block Coordinate Descent Method for Computing the ...
    Sep 27, 2021 · the the polar decomposition as the retraction, L1 =1+. √. 2/2,L2 = √. 10. 3 A Riemannian Block Coordinate Descent Algorithm for Com- puting ...