Fact-checked by Grok 2 weeks ago

Cayley transform

The Cayley transform is a family of mathematical mappings named after the 19th-century British mathematician , with key formulations in linear algebra and that preserve important geometric or algebraic structures. In linear algebra, it provides a bijective correspondence between skew-symmetric matrices S (satisfying S^T = -S) and proper orthogonal matrices R \in SO(n) (without the eigenvalue -1), defined by the formula R = (I - S)(I + S)^{-1}, where I is the and I + S is invertible; this parametrization avoids singularities associated with other rotation representations like . In complex analysis, it denotes the Möbius transformation \phi(z) = \frac{z - i}{z + i}, which conformally maps the upper half-plane \{z \in \mathbb{C} : \Im(z) > 0\} bijectively onto the open unit disk \{w \in \mathbb{C} : |w| < 1\}, sending the real axis to the unit circle (excluding the point 1) and \infty to 1. These transforms, introduced in their matrix form by Cayley in 1846, facilitate computations in diverse fields by converting problems between equivalent domains. The matrix Cayley transform originates from Cayley's early work on linear transformations and was later generalized by mathematicians like in the context of and classical matrix ensembles. It maps the of skew-symmetric matrices (infinitesimal rotations) to the of rotations, ensuring that the resulting R satisfies R^T R = I and \det(R) = 1. Key properties include its invertibility—recovering S = (I - R)(I + R)^{-1}—and its extension to handle the full via compositions or diagonal sign adjustments, though it excludes matrices with -1 as an eigenvalue to avoid singularities. This formulation is particularly valuable in numerical stability, as skew-symmetric parameters can be exponentiated directly to yield rotations via the in limiting cases. In complex analysis, the Cayley transform exemplifies the power of linear fractional transformations (automorphisms of the Riemann sphere) for solving boundary value problems and studying modular forms, as it interchanges the hyperbolic geometry of the half-plane with that of the disk. Its inverse \phi^{-1}(w) = i \frac{1 + w}{1 - w} similarly maps the unit disk to the upper half-plane, preserving angles and facilitating the transfer of analytic properties across domains; for instance, it sends the origin to i and the unit circle (minus 1) to the real line. This transform is fundamental in the proof of the and in applications like the , where functions on the disk are analyzed via half-plane counterparts. Beyond these core areas, Cayley transforms find applications in physics and engineering, such as parameterizing attitude matrices in spacecraft control (higher-order variants extend to full SO(3) coverage) and representing unitary operators in quantum mechanics via skew-Hermitian generators. In signal processing, generalized forms parameterize paraunitary filter banks for perfect reconstruction. These uses highlight the transform's role in bridging algebraic structures with geometric intuitions, influencing modern topics from random matrix theory to K-theory.

Introduction and Properties

Definition

The Cayley transform is a Möbius transformation that maps the upper half-plane of the complex plane conformally onto the unit disk. Specifically, for a complex number z with positive imaginary part, the transform is given by w = \frac{z - i}{z + i}, which sends the real axis to the unit circle (excluding the point 1) and the point i to the origin. In the context of linear algebra over the complex numbers, the Cayley transform provides a birational map from skew-Hermitian matrices to unitary matrices. For an n \times n skew-Hermitian matrix A (satisfying A^* = -A), the transform is defined as Q = (I - A)(I + A)^{-1}, provided that -1 is not an eigenvalue of A (ensuring I + A is invertible); the resulting Q satisfies Q^* Q = I. For the real case, when A is a real skew-symmetric matrix (A^T = -A), the same formula Q = (I - A)(I + A)^{-1} maps to the special orthogonal group, yielding matrices Q with Q^T Q = I and \det Q = 1, again under the condition that -1 is not an eigenvalue of A. This transform, originally introduced by in 1846, preserves key structures such as conformality in the geometric setting (mapping angles to angles) and unitarity or orthogonality in the algebraic setting.

History

The Cayley transform was introduced by in his 1846 paper "On linear transformations," published in the Cambridge and Dublin Mathematical Journal. In this work, Cayley explored mappings between skew-symmetric matrices and orthogonal matrices in the context of linear transformations, marking an early contribution to what would become . Cayley's contribution formed part of his broader investigations into linear transformations during the mid-1840s, predating his formal development of in the 1858 memoir "A memoir on the theory of matrices," where he systematically treated matrices as algebraic objects. The transform initially appeared in studies of real linear transformations, reflecting the 19th-century British mathematical school's emphasis on algebraic structures. During the latter half of the 19th century, the Cayley transform found early applications in invariant theory and algebraic geometry, areas central to Cayley's research and the British algebraic tradition exemplified by figures like . These uses highlighted its role in preserving geometric properties under group actions. Over time, the transform evolved from its real matrix origins to generalizations in complex and quaternionic settings, building on 's 1843 invention of quaternions, which Cayley studied in prior works.

Basic Properties

The Cayley transform is invertible, with the inverse given by A = (I - Q)(I + Q)^{-1} in the real matrix case. In the complex case, the transform Q = (I - A)(I + A)^{-1} for skew-Hermitian A has inverse A = (I - Q)(I + Q)^{-1}, assuming I + A is invertible. This mapping preserves key algebraic structures: in the real setting, it sends skew-symmetric matrices to proper orthogonal matrices with determinant 1, while in the complex setting, it maps skew-Hermitian matrices to unitary matrices. Moreover, the Cayley transform maintains the Lie algebra structure, providing a diffeomorphism between open subsets of the Lie algebra and the corresponding Lie group, such as from \mathfrak{so}(n) to SO(n). In the scalar complex form \phi(z) = \frac{z - i}{z + i}, the transform is birational, serving as a conformal map from the upper half-plane to the unit disk, analytic everywhere except at z = -i, and bijective onto its image. For the matrix case with traceless skew-symmetric A (which all skew-symmetric matrices are), the resulting orthogonal Q satisfies \det Q = 1. The Cayley transform interacts with the exponential map by approximating it for small elements in the Lie algebra, as Q \approx \exp(2A) when \|A\| is small, and it commutes with certain conjugations in the group.

Geometric Interpretations

Real Homography

The Cayley transform provides a geometric interpretation as a projective transformation, or homography, that maps the extended real line \mathbb{R} \cup \{\infty\} bijectively onto the unit circle in the complex plane. In its standard form for the real case, the map is given by \phi(t) = \frac{t - i}{t + i} for t \in \mathbb{R}, where i is the imaginary unit. This formula yields points on the unit circle because for real t, |\phi(t)| = 1. The point at infinity is mapped to \phi(\infty) = 1, while specific points such as t = 0 map to -1. The transformation is conformal, preserving angles in the sense of 1-dimensional projective geometry, and arises as a restriction of the full to the real line. As a homography, the Cayley transform can be represented by the $2 \times 2 complex matrix \begin{pmatrix} 1 & -i \\ 1 & i \end{pmatrix} acting on homogeneous coordinates in the projective line \mathbb{RP}^1, up to scalar multiple; this induces the fractional linear transformation while preserving cross-ratios, a key property of projective transformations over the reals extended to this embedding. Although the coefficients are complex, the restriction to real inputs ensures the image lies on the real-projective structure of the unit circle, which is projectively equivalent to \mathbb{RP}^1 via stereographic projection. Fixed points of the map occur where \phi(t) = t, leading to the quadratic equation t^2 + (i-1)t + i = 0, with no real solutions, indicating no fixed points on the domain; the singularity (pole) at t = -i lies off the real line, ensuring the map is well-defined and smooth on \mathbb{R} \cup \{\infty}\}. This structure highlights its role as an involution up to composition with reflection, analogous to inversions in projective geometry. Geometrically, the transform offers analogies to stereographic projection, converting the linear structure of the extended real line into the circular topology of the unit circle, facilitating visualizations of projective equivalences. For instance, the interval (-\infty, 0) maps to the upper semicircle arc approaching 1 to -1, while (0, \infty) maps to the lower semicircle arc from -1 approaching 1; the point t=1 maps to -i, sending the origin-adjacent interval to a quarter-arc near the negative imaginary axis. In real analysis, this mapping aids in solving boundary value problems by conformally transporting conditions from the unbounded line to the compact circle, where periodic or circular boundary behaviors simplify integral equations or harmonic functions, similar to how stereographic projection resolves spherical problems onto the plane.

Complex Homography

The Cayley transform in the context of complex homography refers to the Möbius transformation that establishes a biholomorphic correspondence between the open upper half-plane and the open unit disk in the complex plane. Defined by the formula w = \frac{z - i}{z + i}, this map sends the domain \{ z \in \mathbb{C} \mid \Im(z) > 0 \} bijectively onto \{ w \in \mathbb{C} \mid |w| < 1 \}. The transformation arises naturally as a specific instance of a linear fractional transformation and serves as a fundamental tool in complex analysis for domain mappings. As a boundary case of the real homography restricted to the real line, the complex version extends analytically to the interior, enabling mappings of regions rather than just boundaries. The inverse map, which recovers the upper half-plane from the unit disk, is given by z = i \frac{1 + w}{1 - w}. This inverse is also a Möbius transformation and confirms the bijectivity of the original map. On the boundary, the extended real axis (including infinity) is sent to the unit circle: specifically, the real axis maps to the unit circle excluding the point w = 1, while the point at infinity maps to w = 1. These properties ensure that the transformation aligns boundaries appropriately, facilitating the study of functions across equivalent domains. Being holomorphic and nowhere zero in the upper half-plane (with a simple pole at z = -i outside the domain), the Cayley transform preserves angles and orientation, rendering it conformal. This conformal invariance is central to its role in the Riemann mapping theorem, where the upper half-plane's equivalence to the unit disk exemplifies how simply connected domains in the complex plane can be normalized to a standard form for analysis. In applications, the Cayley transform bridges the Poincaré upper half-plane model and the Poincaré disk model of two-dimensional hyperbolic geometry, providing an isometry that preserves geodesic distances and the hyperbolic metric. This equivalence allows computations in one model to be transferred to the other, aiding in the visualization and study of hyperbolic structures. Furthermore, the map is instrumental in solving boundary value problems for , such as Dirichlet problems, by conformally transporting complicated boundaries in the upper half-plane to the unit disk, where harmonic functions with circular symmetry can be more readily determined.

Quaternion Extension

The quaternion extension of the Cayley transform generalizes the mapping to the non-commutative algebra of quaternions \mathbb{H}, providing a bijection between pure imaginary quaternions and unit quaternions excluding the point -1. For a pure imaginary quaternion q \in \mathbb{H} with q^* = -q, the transform is defined as u = (1 - q)(1 + q)^{-1}, where u is a unit quaternion satisfying |u| = 1. This formula arises from the representation of quaternions as $2 \times 2 complex matrices, where the transform aligns with the standard Cayley map on the Lie algebra \mathfrak{su}(2) to the group \mathrm{SU}(2). The mapping covers the space of pure imaginary quaternions (isomorphic to \mathbb{R}^3) onto the 3-sphere of unit quaternions minus -1, preserving the hyperbolic geometry of the domain in a manner analogous to the complex case but accounting for non-commutativity. As q approaches infinity in norm, u approaches -1, which is excluded to ensure invertibility. The inverse transform recovers the pure imaginary quaternion via q = (1 - u)(1 + u)^{-1}, which is well-defined for u \neq -1 and yields a pure imaginary result due to the unit norm of u. An equivalent form, up to sign convention, is q = i(u + 1)(u - 1)^{-1} when adjusting for specific basis choices in the imaginary part, ensuring q remains pure. This parameterization relates directly to 3D rotations, as the unit quaternions form the group \mathrm{SU}(2) \cong \mathrm{Spin}(3), the double cover of the rotation group \mathrm{SO}(3). A pure imaginary q = \tan(\theta/2) \, \mathbf{n} (with unit vector \mathbf{n} along the rotation axis and angle \theta) maps to the unit quaternion u = \cos(\theta/2) + \sin(\theta/2) \, \mathbf{n}, enabling efficient representation of rotations without singularities except at \theta = 2\pi. Rotations act on vectors (as pure quaternions \mathbf{v}) via conjugation: \mathbf{v}' = u \mathbf{v} u^{-1}. Key properties highlight the non-commutative structure: the transform does not preserve multiplication directly but relates products via conjugation, as u_1 u_2 corresponds to a composed rotation, while the images under the map satisfy u_1 u_2 u_1^* \approx transformed product up to Lie algebra elements. This non-commutativity distinguishes it from the complex homography (a scalar commutative case) and ensures the map is a local diffeomorphism near the identity, facilitating numerical stability in rotation interpolation despite the exclusion of -1.

Algebraic Applications

Matrix Mappings

The Cayley transform provides a mapping from the Lie algebra \mathfrak{so}(n) of skew-symmetric n \times n real matrices to the special orthogonal Lie group SO(n). Specifically, for A \in \mathfrak{so}(n), the transform is given by Q = (I + A)(I - A)^{-1}, where I is the n \times n identity matrix, provided that I - A is invertible. This formula ensures that Q \in \mathrm{SO}(n), as Q^\top Q = I and \det Q = 1. An analogous construction applies to the special unitary group SU(n). For K \in \mathfrak{su}(n), the Lie algebra of traceless skew-Hermitian n \times n complex matrices, the Cayley transform is U = (I - K)(I + K)^{-1}, where the Hermitian adjoint replaces the transpose, yielding U \in \mathrm{SU}(n) with U^\dagger U = I and \det U = 1, assuming I + K is invertible. The Cayley transform relates closely to the exponential map on the Lie group. For small elements A \in \mathfrak{so}(n), Q \approx \exp(2A), serving as a rational approximation (specifically, the [1/1] Padé approximant) to the exponential, though it defines an exact birational map between the respective varieties. This approximation facilitates numerical computations and parameterizations near the identity. The transform preserves the dimension of the parameter space: \dim \mathfrak{so}(n) = n(n-1)/2 = \dim \mathrm{SO}(n), and similarly \dim \mathfrak{su}(n) = n^2 - 1 = \dim \mathrm{SU}(n), ensuring a faithful local parameterization. The mapping is bijective, with every Q \in \mathrm{SO}(n) except those with eigenvalue -1 attained exactly once from a unique A \in \mathfrak{so}(n). The inverse transform recovers A = (Q - I)(Q + I)^{-1}, valid wherever I + Q is invertible.

Examples

One illustrative example of the Cayley transform arises in the two-dimensional real case, mapping a 2×2 skew-symmetric matrix to a matrix in SO(2). Consider the skew-symmetric matrix A = \begin{pmatrix} 0 & -\theta \\ \theta & 0 \end{pmatrix} for real \theta > 0. The Cayley transform is given by Q = (I + A)(I - A)^{-1}. First, compute I - A = \begin{pmatrix} 1 & \theta \\ -\theta & 1 \end{pmatrix}, with determinant $1 + \theta^2 and inverse \frac{1}{1 + \theta^2} \begin{pmatrix} 1 & -\theta \\ \theta & 1 \end{pmatrix}. Then, I + A = \begin{pmatrix} 1 & -\theta \\ \theta & 1 \end{pmatrix}, yielding Q = \frac{1}{1 + \theta^2} \begin{pmatrix} 1 - \theta^2 & -2\theta \\ 2\theta & 1 - \theta^2 \end{pmatrix}. This matrix represents a by angle \phi = 2 \arctan \theta, as the entries match the standard form \begin{pmatrix} \cos \phi & -\sin \phi \\ \sin \phi & \cos \phi \end{pmatrix}. To verify , direct computation confirms Q^T Q = I. In three dimensions, the Cayley transform maps 3×3 skew-symmetric matrices in the so(3) to matrices in SO(3), with a direct relation to . For a around a unit axis \mathbf{u} = (u_1, u_2, u_3)^T by angle \phi, the corresponding is A = \phi \hat{u}, where \hat{u} = \begin{pmatrix} 0 & -u_3 & u_2 \\ u_3 & 0 & -u_1 \\ -u_2 & u_1 & 0 \end{pmatrix}. Parameterizing via A = \tan(\phi/2) \hat{u}, the Cayley transform Q = (I + A)(I - A)^{-1} produces the Q = I + \frac{2}{1 + \tan^2(\phi/2)} \left( \tan(\phi/2) \hat{u} + \tan^2(\phi/2) \hat{u}^2 \right), which aligns with the Rodrigues formula Q = I + \sin \phi \, \hat{u} + (1 - \cos \phi) \hat{u}^2 upon . For a specific case, take \mathbf{u} = (0,0,1)^T and \phi = \pi/2, so \theta = \tan(\pi/4) = 1 and A = \begin{pmatrix} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}. Then I - A = \begin{pmatrix} 1 & 1 & 0 \\ -1 & 1 & 0 \\ 0 & 0 & 1 \end{pmatrix}, with inverse \frac{1}{2} \begin{pmatrix} 1 & -1 & 0 \\ 1 & 1 & 0 \\ 0 & 0 & 2 \end{pmatrix}, and Q = (I + A)(I - A)^{-1} = \begin{pmatrix} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 1 \end{pmatrix}, a 90-degree around the z-axis. holds as Q^T Q = I. This parameterization covers rotations with |\phi| < \pi. For the complex case in SU(2), relevant to quantum mechanics and qubit states, the Cayley transform maps anti-Hermitian traceless matrices in su(2) to special unitary matrices. The Lie algebra su(2) is spanned by i times the Pauli matrices \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix} , \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix} , \sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} . Consider a general element K = i \frac{\theta}{2} \mathbf{n} \cdot \vec{\sigma} for unit vector \mathbf{n} and real \theta, which is skew-Hermitian. Equivalently, let H = \frac{\theta}{2} \mathbf{n} \cdot \vec{\sigma} be Hermitian, then the transform is U = (I - i H)(I + i H)^{-1}, yielding a unitary matrix representing a rotation in the qubit Bloch sphere by angle \theta around \mathbf{n}. For the specific case along the z-axis, take H = \frac{\theta}{2} \sigma_z = \frac{\theta}{2} \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}. Then I + i H = \begin{pmatrix} 1 + i \frac{\theta}{2} & 0 \\ 0 & 1 - i \frac{\theta}{2} \end{pmatrix}, with inverse \frac{1}{1 + \frac{\theta^2}{4}} \begin{pmatrix} 1 - i \frac{\theta}{2} & 0 \\ 0 & 1 + i \frac{\theta}{2} \end{pmatrix}. Thus, U = (I - i H)(I + i H)^{-1} = \begin{pmatrix} \cos\frac{\theta}{2} - i \sin\frac{\theta}{2} & 0 \\ 0 & \cos\frac{\theta}{2} + i \sin\frac{\theta}{2} \end{pmatrix}, which acts on qubit states as a rotation operator. Unitarity is verified by U^\dagger U = I, with \det U = 1. This construction links directly to qubit state evolution under SU(2) transformations. An important edge case occurs when the Cayley transform is undefined: for the orthogonal group, matrices Q with eigenvalue -1 (e.g., Q = -I) lie outside the image of the transform, as I - A becomes singular if A has eigenvalue 1, which corresponds to rotations by odd multiples of \pi. A scaled variant of the transform can mitigate this limitation in applications.

Generalizations to Other Groups

The Cayley transform extends to the symplectic Lie group \mathrm{Sp}(2n, \mathbb{R}), the group of $2n \times 2n real matrices preserving a symplectic form, by providing a retraction from the Lie algebra \mathfrak{sp}(2n, \mathbb{R}) (consisting of matrices A satisfying A^\top J + J A = 0, where J is the standard symplectic matrix) to the group itself. This generalization, known as the symplectic Cayley transform, is defined for a tangent vector Z at a point X \in \mathrm{Sp}(2p, 2n) as R_{\mathrm{cay}, X}(Z) = \left(I - \frac{1}{2} S_{X,Z} J \right)^{-1} \left(I + \frac{1}{2} S_{X,Z} J \right) X, where S_{X,Z} = G_X Z (X^\top J)^\top + X^\top J (G_X Z)^\top and G_X = I - \frac{1}{2} X J X^\top J^\top, ensuring preservation of the symplectic structure. It arises naturally in the formulation of diagonally implicit for isospectral systems on quadratic Lie groups, including \mathrm{Sp}(2N, \mathbb{R}), where the transform \mathrm{cay}(\xi) = (I - \xi/2)^{-1} (I + \xi/2) maintains eigenvalues and symplectic properties during numerical integration. A complex analogue exists for \mathrm{Sp}(n, \mathbb{C}), adapting the transform to preserve the Hermitian symplectic form in representations. In the context of conformal groups, the Cayley transform generalizes to rational mappings on Jordan algebras associated with pseudo-Euclidean spaces, facilitating embeddings of Lorentz groups O(p,q) into larger conformal structures. For instance, the transform R(x) = (x - e)(x + e)^{-1}, where e is the identity element, conjugates inversion to negation in the algebra, enabling a conformal compactification of the Lorentz group via graph embeddings into asymmetric matrices \mathrm{Asym}(I_{p,q}, \mathbb{R})^c. This construction links O(p,q) to the conformal group \mathrm{Co}(V)_0 for V = \mathrm{Herm}(n, \mathbb{C}), isomorphic to \mathrm{SU}(n,n), and supports causal structures in higher-dimensional spacetime models. Higher-order Cayley transforms iterate the classical form to enhance approximations of the matrix exponential \exp(Q) for skew-symmetric Q, particularly in numerical schemes for Lie group differential equations. Defined as C = (I - Q)^n (I + Q)^{-n} for integer n > 1, these transforms expand the parameter domain (e.g., beyond 180° rotations in \mathrm{SO}(3)) while avoiding singularities near the origin, offering super-exponential convergence rates in Hilbert spaces. They are employed in attitude determination and geometric integration, such as commutator-free methods combining Cayley–Magnus expansions for high-order accuracy in simulations. For non-compact groups like \mathrm{SL}(n, \mathbb{R}), the generalized Cayley map \Phi projects the group onto its \mathfrak{sl}(n, \mathbb{R}) via orthogonal projection in a representation space, establishing a between hyperbolic elements in the group (those with positive eigenvalues) and hyperbolic elements in the algebra. This mapping is conjugation-equivariant and preserves regular orbits, facilitating analysis of hyperbolic dynamics in non-compact semisimple groups. Despite these extensions, the Cayley transform does not parametrize all groups uniformly; its birational nature depends on the representation's power span property and weight diagram geometry, failing for certain non-compact forms like \rho(\mathrm{SL}_2(\mathbb{C})) in symmetric powers due to incomplete coverage of Cartan subalgebras. Compactness issues arise in non-compact cases, where singularities (e.g., at \det(I - u) = 0) limit global applicability, restricting the transform to local charts or specific conjugacy classes.

Functional Analysis

Operator Mappings

In the context of functional analysis, the Cayley transform provides a bijection between bounded self-adjoint operators H and unitary operators U without the eigenvalue -1 on a Hilbert space \mathcal{H}. For a bounded self-adjoint operator H on \mathcal{H}, satisfying H^* = H, the transform is defined as U = (I - iH)(I + iH)^{-1}, where I denotes the identity operator. This mapping is well-defined because H is self-adjoint with real spectrum, ensuring I + iH is invertible as its spectrum has real part 1. The operator U is unitary, meaning U^* U = I and U U^* = I on \mathcal{H}. The adjoint is U^* = (I - iH)^{-1}(I + iH), and direct computation confirms unitarity since (I - iH)(I + iH) = (I + iH)(I - iH) = I + H^2, a positive , allowing simplification to the identity. This unitarity follows from the self-adjointness of H. The construction relates closely to the for self-adjoint operators, which decomposes H via a spectral measure, allowing the Cayley transform to be understood through on the spectrum. The inverse recovers H = -i (I - U)(I + U)^{-1}, defined for unitaries avoiding -1. This operator-theoretic Cayley transform extends the finite-dimensional matrix case to infinite-dimensional Hilbert spaces via the for bounded operators, where functions of H are defined through its . In the finite-dimensional setting, it maps matrices to unitary matrices avoiding -1, and the infinite-dimensional version preserves this structure on separable Hilbert spaces. In , the transform plays a key role in generating unitary evolution operators from Hamiltonians, facilitating the analysis of e^{-iHt} by mapping the Hamiltonian to a whose spectral properties encode the dynamics.

Infinite-Dimensional Operators

In , the Cayley transform provides a between unbounded operators on a and unitary operators without the eigenvalue 1. For a densely defined, closed, unbounded A: D(A) \subset \mathcal{H} \to \mathcal{H} on a complex \mathcal{H}, where the spectrum \sigma(A) \subset \mathbb{R} ensures that i \notin \sigma(A), the Cayley transform is defined as U = (A - iI)(A + iI)^{-1}, with the resolvent (A + iI)^{-1} being a bounded operator on all of \mathcal{H}. This U is a unitary operator on \mathcal{H}, and its domain is the entire space \mathcal{H}, while the inverse transform recovers A = i(I + U)(I - U)^{-1} on the domain D(A) = \{ x \in \mathcal{H} : (I - U)x \in R(I + U) \}. The condition that \sigma(A) avoids the imaginary axis (specifically, \pm i) guarantees the existence and boundedness of the resolvent, enabling this mapping. A key application arises in semigroup theory, where the Cayley transform translates stability properties from continuous-time evolution to discrete-time iterations. For a generator A of a strongly continuous contraction semigroup \{e^{-tA}\}_{t \geq 0} on \mathcal{H}, the Cayley transform V = (A - I)(A + I)^{-1} (or variants with i for Hilbert spaces) maps to a contraction operator whose powers V^n correspond to discrete approximations of the semigroup. Strong stability of the semigroup—meaning \|e^{-tA} x\| \to 0 as t \to \infty for all x \in \mathcal{H}—implies strong stability of \{V^n\}_{n \in \mathbb{N}}, with \|V^n x\| \to 0 as n \to \infty, under conditions like the absence of eigenvalues on the imaginary axis. This equivalence facilitates the analysis of long-time behavior in infinite-dimensional systems, preserving properties such as polynomial decay rates; for instance, if \|e^{-tA} A^{-1}\| = O(t^{-\alpha}) for \alpha > 0, then \|V^n A^{-1}\| = O(n^{-\alpha}). In partial differential equations (PDEs), the Cayley transform maps generators of contraction to , aiding in existence proofs and stability analysis. For the \partial_t x = \nabla \cdot (\alpha \nabla x) on a \Omega with $0 < mI \leq \alpha \leq MI, the extended operator A_{\text{ext}} = \begin{pmatrix} 0 & \nabla \\ -\nabla^* & 0 \end{pmatrix} generates a contraction semigroup on L^2(\Omega) \oplus L^2(\Omega), and the Cayley transform with multiplier S = (\alpha - I)(\alpha + I)^{-1} yields a unitary preserving energy estimates. Similarly, for damped wave equations like \partial_{tt} u + k_v \partial_t u - \Delta u = 0, the generator A_{\text{ext},v} = \begin{pmatrix} 0 & I \\ \Delta & -k_v \end{pmatrix} produces a contraction semigroup, transformed via Cayley to a unitary group on an extended space, ensuring well-posedness even for degenerate damping. These mappings highlight how unitary extensions simplify for dissipative PDEs. Challenges in applying the Cayley transform to unbounded operators include ensuring maximality and handling domain restrictions. While U acts on all of \mathcal{H} for self-adjoint A, the domain D(U) effectively incorporates D((A + iI)^{-1}) = \mathcal{H}, but for non-self-adjoint extensions of symmetric operators, U is only a partial isometry with domain R(A + iI) and range R(A - iI), requiring deficiency indices n_+(A) = n_-(A) for self-adjoint extensions to exist. Maximality conditions, such as closed range and closure, must be verified to guarantee unitarity, particularly in applications where the approaches the imaginary axis, potentially leading to ill-posed resolvents. In modern , the Cayley transform underpins time-stepping methods for stiff evolution equations, such as integrators. The extrapolated Cayley transform, approximating e^{-tA} via rational functions like (I - (t/2)A)(I + (t/2)A)^{-1} with higher-order corrections, enables efficient of semigroups in dimensions, preserving for large time steps in PDE simulations. For instance, sixth-order schemes based on generalized Cayley transforms, using Padé approximations and fast methods, solve nonlinear PDEs like the Zakharov-Shabat system with unitarity conservation and reduced computational cost. These methods are particularly valuable for high-fidelity simulations in and , where unbounded operators arise naturally.

References

  1. [1]
    [PDF] Remarks on the Cayley Representation of Orthogonal Matrices and ...
    R = (I − S)(I + S)−1. This is a classical result of Cayley [3] (1846) and R is called the Cayley transform of S. Among other sources ...
  2. [2]
    Cayley Transform -- from Wolfram MathWorld
    The linear fractional transformation z|->(iz)/(i+z) that maps the upper half-plane {z:I[z]>0} conformally onto the unit disk {z:|z|<}.
  3. [3]
    [PDF] Complex Analysis with Applications Princeton University MAT330 ...
    Jan 27, 2023 · (this is called the Cayley transform) and its inverse is d−1 : S1. → R eiθ. 7→ i. 1 − eiθ. 1+eiθ. = tan θ. 2. Hence g ◦ d−1 : ∂B1 (0) → R ...
  4. [4]
    [PDF] Higher Order Cayley Transforms with Applications to Attitude ...
    In this paper we generalize previous results on attitude representations using Cayley transforms. First, we show that proper orthogonal matrices, that naturally ...
  5. [5]
    [PDF] Special Paraunitary Matrices, Cayley Transform, and ... - Minh N. Do
    The Cayley transform of paraunitary matrices are PSH matrices. Now we consider the Cayley transform of SPU matrices. We define a special PSH (SPSH) matrix as ...
  6. [6]
    Fall 2014 - Problem 7
    Finally, we can let φ 8 ( z ) = z − i z + i \phi_8\p{z} = \frac{z - i}{z + i} φ8​(z)=z+iz−i​, the Cayley transform, which maps the upper half-plane to the unit ...
  7. [7]
    [PDF] On the Cayley transform of positivity classes of matrices
    Feb 22, 2002 · We begin by making precise the term Cayley transform. Definition 1.1. Let A ∈ Mn(C) such that I + A is invertible. The Cayley transform of ...
  8. [8]
    Cayley's parameterization of orthogonal matrices - PlanetMath.org
    Mar 22, 2013 · Hence, the Cayley transform is defined for all matrices such that −1 - 1 is not an eigenvalue of O O . (Recall that this condition is ...Missing: mathematics | Show results with:mathematics
  9. [9]
    XIII. On certain results relating to quaternions - Taylor & Francis Online
    (1845). XIII. On certain results relating to quaternions . The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science: Vol. 26, No.
  10. [10]
    Arthur Cayley - Biography - MacTutor - University of St Andrews
    Cayley's introductory paper in matrix theory was written in French and published in a German periodical [in 1855]. ... [He] introduces, although quite sketchily, ...
  11. [11]
    [PDF] Abstract Introduction Higher Order Cayley Transforms
    Cayley transforms can be used to generate higher-order. Cayley parameters defined by pm = ê tan. ( ф. 2m. ) m = 3,4,... (13). Moreover, it should be obvious why ...
  12. [12]
    [PDF] Random orthogonal matrices and the Cayley transform - arXiv
    Oct 5, 2018 · In this version, the Cayley transform of X is the p × k orthogonal matrix. C(X)=(Ip + X)(Ip − X)−1. Ip×k where Ip×k denotes the p×k matrix ...
  13. [13]
    [PDF] HIGHER ORDER CAYLEY TRANSFORMS WITH APPLICATIONS ...
    Viewing the Cayley transform as a bilinear transformation which maps the space of skew-symmetric matrices onto the space of proper orthogonal matrices (and vice ...
  14. [14]
  15. [15]
    [PDF] Reflections on the Lemniscate of Bernoulli: The Forty-Eight Faces of ...
    Jul 15, 2010 · We note that the Cayley map k(z) restricts to a conformal map of the upper half-plane H = {z = x + iy : y > 0} onto the unit disc D = {z = x + ...
  16. [16]
    [PDF] A Guide to Complex Variables
    Oct 14, 2007 · 6.2.5 The Cayley Transform. Theorem (The Cayley Transform): The linear fractional transformation z +→ (i − z)/(i + z) maps the upper half ...
  17. [17]
    Quaternionic Cayley transform revisited - ScienceDirect
    The classical Cayley transform κ ( t ) = t − i t + i = t 2 − 1 − 2 i t 1 + t 2 is a bijective map between the real line R and the set T ∖ { 1 } , where T is the ...
  18. [18]
    Cayley Transform
    The image below shows how the function g(z) = (zi)/(z+i) maps the upper-half plane into the unit disk.
  19. [19]
    Upper Half Plane - an overview | ScienceDirect Topics
    The inverse Cayley transform K - 1 ( z ) = i 1 - z 1 + z takes the unit circle in the real axis and the unit disk in the upper half-plane. Consider the image of ...
  20. [20]
  21. [21]
    Review of the exponential and Cayley map on SE(3) as relevant for ...
    Sep 1, 2021 · Using unit quaternions (Euler parameters) is a common choice, which implies using SU(2) to represent rotations [55,63]. However, this ...
  22. [22]
  23. [23]
    [PDF] Orthogonal Recurrent Neural Networks with Scaled Cayley Transform
    To construct the orthogonal weight matrix, we parametrize it with a skew-symmetric matrix through a scaled Cayley transform. This scaling allows us to avoid the ...
  24. [24]
    [PDF] RIEMANNIAN OPTIMIZATION ON THE SYMPLECTIC STIEFEL ...
    Jun 26, 2020 · By exploiting the low-rank structure of the tangent vectors, we construct a numerically efficient update for the symplectic Cayley transform. In ...
  25. [25]
    Variational symplectic diagonally implicit Runge-Kutta methods for ...
    Jul 9, 2022 · Our derivation relies on a formulation of diagonally implicit isospectral Runge-Kutta methods in terms of the Cayley transform.
  26. [26]
    [PDF] Cayley transform, Lie groups, symmetric spaces, Stiefel manifolds
    The Cayley transform for orthogonal groups is a well known construction with applications going from analysis and linear algebra to computer science, ...
  27. [27]
    [PDF] On some Causal and Conformal Groups - Wolfgang Bertram
    Generalizing the example of the group U(p, q), we will find and describe a fairly large class of spaces having a conformal or causal structure which is ...
  28. [28]
    [PDF] the generalized cayley map from an algebraic group to its lie algebra
    For the spin representation of Spin(V ) the map Φ essentially coincides with the classical. Cayley transform. ... Lie algebra is reductive. For abelian Lie ...
  29. [29]
    [PDF] The Cayley Transform on Representations - arXiv
    The classical Cayley transform is a birational map between a quadratic matrix group and its Lie algebra, which was first discovered by Cayley in 1846. Because ...
  30. [30]
    Allgemeine Eigenwerttheorie Hermitescher Funktionaloperatoren
    Die vollstetigen und die beschränkten Formen (bzw. Operatoren) wurden von Hilbert entdeckt, und er löste ihre Eigenwertproblem (Göttinger Nachr. 1906, S. 157– ...
  31. [31]
    [PDF] The Spectral Theorem for Self-Adjoint and Unitary Operators
    generally, if H is another Hilbert space, we say Φ ∈ L(H,H) is unitary provided Φ is one-to-one and onto, and (Φu,Φv)H = (u, v)H, for all u, v ∈ H.
  32. [32]
    [PDF] Chen,Aden.pdf - UChicago Math
    We will first develop the spectral theorem for unitary operators and then pass it to self-adjoint operators using the Cayley transform. Along the way, we also ...
  33. [33]
    [PDF] The Cayley Transform and Self–adjoint Extensions - UBC Math
    Nov 27, 2018 · Throughout these notes H is a Hilbert space. We shall use D(B) and R(B) to denote the domain and range, respectively, of the linear operator B.
  34. [34]
    [PDF] 3 Self-adjoint operators (unbounded)
    Continuous functions of unitary operators are defined. In combination with the Cayley transform they provide some functions of self-adjoint operators. The ...
  35. [35]
    Stability Analysis in Continuous and Discrete Time, using the Cayley ...
    Jun 22, 2010 · Systems with a finite Bergman distance share the same stability properties, and the Bergman distance is preserved under the Cayley transform.
  36. [36]
  37. [37]
    None
    ### Summary of Sections on Cayley Transform, Decay Estimates, and Stability Preservation for Infinite-Dimensional Polynomially Stable Semigroups
  38. [38]
    [PDF] Feedback theory extended for proving generation of contraction ...
    Abstract. Recently, the following novel method for proving the existence of solu- tions for certain linear time-invariant PDEs was introduced: The operator.
  39. [39]
    [PDF] Fast sixth-order algorithm based on the generalized Cayley ... - arXiv
    Nov 23, 2020 · The exponential integrator is a special case. Schemes based on rational approximation allow the use of fast algorithms to solve the initial ...