Fact-checked by Grok 2 weeks ago

Wigner rotation

In special relativity, the Wigner rotation is the spatial rotation that emerges from the composition of two successive non-collinear Lorentz boosts, resulting in a Lorentz transformation equivalent to a single boost combined with a rotation rather than a pure boost. This effect highlights the non-commutativity of boosts in different directions, a feature absent in Galilean transformations of classical mechanics. The magnitude of the rotation, known as the Wigner angle, depends on the velocities and directions of the boosts and can be derived using the Lorentz factors \gamma_1 and \gamma_2, as well as the velocity components \beta_1 and \beta_2. Although named after , who analyzed it in the context of unitary representations of the inhomogeneous in his seminal 1939 paper, the phenomenon was first discovered by in 1913 and independently derived by Ludwik Silberstein in 1914, with further developments by Llewellyn Thomas in the 1920s. Wigner's contribution framed it within the "little group" of the , where rotations appear in the irreducible representations for particles with positive energy and momentum, essential for classifying elementary particles by . The rotation is order-dependent: reversing the sequence of boosts inverts the direction of the rotation. The Wigner rotation has profound implications in relativistic physics, particularly in explaining the , where a continuously accelerating particle with experiences a precession in its rest frame due to the cumulative effect of infinitesimal Wigner rotations along its worldline. This precession corrects the spin-orbit coupling in atomic spectra and is crucial for understanding the behavior of spinning particles in electromagnetic fields or gravitational contexts. In , it affects the evolution of spin states and entanglement correlations, such as in relativistic extensions of the Einstein-Podolsky-Rosen paradox, where the rotation reduces violations of Bell inequalities unless compensated by adjusted measurement bases. Applications extend to particle accelerators, where it influences beam polarization, and to general relativity analogs in curved spacetimes like the .

Physical Setup and Velocity Composition

Frame Configurations and Boosts

In , the Wigner rotation arises in scenarios involving three inertial frames, denoted as Σ, Σ', and Σ'', where Σ' moves with velocity \mathbf{u} relative to Σ, and Σ'' moves with velocity \mathbf{v} relative to Σ', with \mathbf{u} and \mathbf{v} being non-collinear. This configuration models the composition of successive Lorentz transformations between frames with misaligned relative motions, such as in the motion of particles or rigid bodies undergoing velocity changes in different directions. Lorentz boosts, which are elements of the , describe these relative motions by altering the coordinates of events while preserving the interval. A pure is a that relates two inertial frames moving at constant along a single direction, resulting solely in a change of without any spatial . When the relative velocities are collinear, the of two such boosts yields another pure boost. However, for non-collinear velocities, the introduces an additional spatial , known as the Wigner rotation, alongside the net boost effect. The coordinate systems in these frames are typically aligned at the origins at a chosen event, using standard Cartesian spatial axes and time synchronized via Einstein's convention. The Wigner rotation manifests as a misalignment of the spatial axes between the initial and final frames, effectively rotating the orientation of objects or reference directions in Σ'' relative to Σ. This axis misalignment highlights the non-commutative nature of non-parallel boosts in the . In the reversed configuration, where the order of boosts is swapped—first \mathbf{v} relative to Σ to reach an intermediate frame, then \mathbf{u} to Σ''—the resulting Wigner rotation is the of the original, changing the sense of while preserving the . This property underscores the dependence of the rotation on the sequence of velocity compositions.

Relativistic Velocity Addition

In , the composition of velocities differs fundamentally from , where velocities simply add as vectors. Classical vector addition can yield a resultant speed exceeding the c, violating the principle that no object can reach or surpass c. Relativistic velocity addition accounts for the invariance of c and the structure of , ensuring the composed remains below c. This adjustment arises from the applied to the coordinates and times used to define velocity, leading to a non-linear combination that preserves and the structure. The general formula for relativistic velocity addition applies when the velocities \mathbf{u} and \mathbf{v} are non-collinear. Consider a reference frame S' moving with velocity \mathbf{u} relative to frame S, where an object has velocity \mathbf{v} in S'. To derive the velocity \mathbf{w} in S, decompose \mathbf{v} into components parallel (\mathbf{v}_\parallel) and perpendicular (\mathbf{v}_\perp) to \mathbf{u}, with v_\parallel = (\mathbf{v} \cdot \mathbf{u}) / u and u = |\mathbf{u}|. The parallel component is w_\parallel = \frac{u + v_\parallel}{1 + \frac{u v_\parallel}{c^2}}, while the perpendicular component is \mathbf{w}_\perp = \frac{\mathbf{v}_\perp}{\gamma_u \left(1 + \frac{u v_\parallel}{c^2}\right)}, where \gamma_u = \frac{1}{\sqrt{1 - \frac{u^2}{c^2}}} is the Lorentz factor for \mathbf{u}. The full velocity is \mathbf{w} = w_\parallel \hat{\mathbf{u}} + \mathbf{w}_\perp, where \hat{\mathbf{u}} = \mathbf{u}/u. This formulation shows that the perpendicular component is suppressed by the factor $1/\gamma_u > 1, reflecting time dilation and length contraction effects in the moving frame. For arbitrary directions of \mathbf{u}, the parallel and perpendicular directions are defined relative to \mathbf{u}. The Lorentz factor for the composed velocity \mathbf{w} is given by \gamma_w = \gamma_u \gamma_v \left(1 + \frac{\mathbf{u} \cdot \mathbf{v}}{c^2}\right), where \gamma_v = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} and v = |\mathbf{v}|. This relation follows from the normalization of the four-velocity under Lorentz transformations and highlights the multiplicative nature of relativistic energies and momenta in velocity composition. In the collinear case, where \mathbf{v} is parallel to \mathbf{u} (so \mathbf{v}_\perp = 0 and \mathbf{u} \cdot \mathbf{v} = u v), the formula simplifies to the one-dimensional addition w = \frac{u + v}{1 + \frac{u v}{c^2}}, with no perpendicular adjustment and \gamma_w = \gamma_u \gamma_v \left(1 + \frac{u v}{c^2}\right). For example, if u = 0.8c and v = 0.8c in the same direction, classical addition gives $1.6c, but the relativistic result is w \approx 0.976c < c, demonstrating the formula's role in maintaining physical consistency. This collinear simplification produces no rotational misalignment between frames, unlike the non-collinear case, where the altered direction of \mathbf{w} relative to a naive sum necessitates a compensatory rotation in subsequent frame alignments.

Mathematical Formulation

Composition of Lorentz Boosts

In special relativity, a Lorentz boost represents a change of inertial frame moving at constant velocity \vec{v} relative to the original frame. In 3+1 dimensional Minkowski spacetime with metric signature (+,-,-,-) and 4-vectors ordered as (ct, \vec{x}), the general form of the boost matrix B(\vec{v}) is B(\vec{v}) = \begin{pmatrix} \gamma_v & -\gamma_v \vec{\beta}_v^T \\ -\gamma_v \vec{\beta}_v & \mathbf{I} + (\gamma_v - 1) \frac{\vec{\beta}_v \vec{\beta}_v^T}{\beta_v^2} \end{pmatrix}, where \gamma_v = (1 - \beta_v^2)^{-1/2}, \vec{\beta}_v = \vec{v}/c, \beta_v = |\vec{\beta}_v|, \mathbf{I} is the 3×3 , and the superscript T denotes the (treating \vec{\beta}_v as a column vector). This matrix preserves the interval and mixes time and space components according to the direction and of \vec{v}. The of two successive Lorentz boosts, \Lambda = B(\vec{v}) B(\vec{u}), where \vec{u} and \vec{v} are the respective velocities, does not generally yield a pure boost unless \vec{u} and \vec{v} are collinear. Instead, the product decomposes as \Lambda = B(\vec{w}) R(\hat{\epsilon}, \epsilon), where B(\vec{w}) is a pure boost with net velocity \vec{w} and R(\hat{\epsilon}, \epsilon) is a spatial rotation by angle \epsilon around unit axis \hat{\epsilon}. This decomposition arises because the Lorentz group is non-abelian, leading to a rotational component in the transformation when the boosts are non-parallel. A pure Lorentz matrix is characterized by in its spatial block and specific antisymmetric time-space mixing in the off-diagonal blocks. Computing the explicit product \Lambda for non-collinear \vec{u} and \vec{v} introduces asymmetric off-diagonal elements in the spatial block, such as terms proportional to cross products of the directions, which cannot be eliminated without incorporating a . For instance, assuming \vec{u} along the x-axis and \vec{v} in the xy-plane, the (2,3) and (3,2) elements of the spatial block differ, violating the of a . The net boost velocity \vec{w} in the decomposition is uniquely determined by the time-space mixing terms in the first row and column of \Lambda (excluding the (0,0) element), which match those of B(\vec{w}) and correspond to the relativistic velocity addition of \vec{u} and \vec{v}. Specifically, the components satisfy -\gamma_w \vec{\beta}_w = the relevant block of \Lambda, ensuring consistency with the non-rotational part of the transformation.

Emergence of the Rotation Component

When two non-collinear pure Lorentz boosts are composed, the resulting transformation \Lambda is generally not a pure boost but includes an additional rotational component, known as the Wigner rotation. This emergence occurs because the is non-abelian, meaning that the order of applying boosts in different directions matters, leading to a spatial that cannot be eliminated by a further boost. The transformation \Lambda can be decomposed uniquely into \Lambda = B(\mathbf{w}) R(\mathbf{n}, \theta), where B(\mathbf{w}) is a pure boost corresponding to the net velocity \mathbf{w} obtained from relativistic velocity addition, and R(\mathbf{n}, \theta) is a spatial by \theta around the \mathbf{n} perpendicular to the plane formed by the two boost directions. This decomposition isolates the rotational part, with the B(\mathbf{w}) aligning the time axis appropriately while the rotation acts on the spatial coordinates. In terms of effect on coordinate axes, the boosted frame's spatial axes are rotated relative to the naive expectation from classical vector addition of velocities. For instance, if one boost is along the x-axis and the second along the y-axis, the composite transformation rotates the xy-plane axes by the Wigner angle, altering the orientation of rods or measurement devices in the final frame without changing lengths or the overall boost velocity magnitude. This rotation manifests as a misalignment between the instantaneous rest frame's orientation and the direction of motion. Geometrically, the Wigner rotation compensates for the non-commutativity of non-parallel boosts, reflecting the of velocity space (or rapidity space). In rapidity space, successive boosts correspond to hyperbolic translations along non-parallel geodesics, and the enclosed area of the formed by these paths determines the rotation angle, ensuring consistency in the Lorentz group's structure. To identify the rotation matrix explicitly, one computes the spatial block of the transformation after removing the pure component: applying the B(-\mathbf{w}) to \Lambda yields B(-\mathbf{w}) \Lambda = R, where the bottom-right 3×3 submatrix of R is precisely the R_3(\mathbf{n}, \theta), satisfying R_3^T R_3 = I and \det R_3 = 1. This spatial block encapsulates the pure rotational effect on position vectors in the .

Derivation of the Wigner Rotation

Axis and Angle Calculation

The axis of the Wigner rotation arising from the composition of two non-collinear Lorentz boosts with velocities \vec{u} and \vec{v} is directed along the unit vector \hat{n} = \frac{\vec{u} \times \vec{v}}{|\vec{u} \times \vec{v}|}. This direction is perpendicular to the spanned by \vec{u} and \vec{v}, reflecting the geometric of the rotation induced by the non-commutativity of boosts in that plane. To derive the rotation angle \theta, consider the Lorentz transformation matrices for the two boosts. A boost with velocity \vec{\beta} = \vec{v}/c (where c is the speed of light) and Lorentz factor \gamma = (1 - \beta^2)^{-1/2} has the spatial part involving contractions and the time-space mixing terms. The composition B_{\vec{v}} B_{\vec{u}} yields a general Lorentz transformation that decomposes as a pure boost B_{\vec{w}} followed by a rotation R(\theta, \hat{n}), where \vec{w} is the composed velocity given by the relativistic velocity addition formula: \vec{w} = \frac{1}{\gamma_u (1 + \vec{\beta_u} \cdot \vec{\beta_v})} \left( \vec{u} + \vec{v}_\parallel + \frac{\vec{v}_\perp}{\gamma_u} \right), with \vec{v}_\parallel and \vec{v}_\perp the components of \vec{v} parallel and perpendicular to \vec{u}, and \gamma_w = \gamma_u \gamma_v (1 + \vec{\beta_u} \cdot \vec{\beta_v}). The rotation emerges because the product matrix is asymmetric in its spatial block, and extracting the symmetric boost part leaves the antisymmetric rotation component. The angle \theta can be found by computing the of the rotation matrix, which satisfies \operatorname{Tr} R = 1 + 2 \cos \theta. For the general case, with angle \phi between \vec{u} and \vec{v}, the explicit formula is \cos \theta = \frac{(\gamma_u + \gamma_v + \gamma_w + 1)^2}{(\gamma_w + 1)(\gamma_u + 1)(\gamma_v + 1)} - 1, where \gamma_w = \gamma_u \gamma_v (1 + \beta_u \beta_v \cos \phi) and \beta_u = u/c, \beta_v = v/c. This expression is obtained by substituting the velocity addition into the condition after decomposing the composed transformation. An equivalent half-angle form, useful for or further derivations, is \tan \frac{\theta}{2} = \frac{\gamma_u \gamma_v \beta_u \beta_v \sin \phi}{\gamma_u + \gamma_v + \gamma_w (1 + \beta_u \beta_v \cos \phi)}, derived from in rapidity space or direct matrix elements. In the limit of infinitesimal boosts (small u, v \ll c), the gammas approach 1 and \gamma_w \approx 1, yielding \theta \approx |\vec{u} \times \vec{v}| / c^2. This approximation links the finite Wigner rotation to the differential rate in continuous acceleration scenarios.

Alternative Derivations

One alternative approach to deriving the Wigner rotation employs parameters to parameterize Lorentz boosts, providing a geometric interpretation in . A boost in direction \mathbf{u} with speed parameter \beta_u = v_u/c is characterized by the \zeta_u, where \beta_u = \tanh \zeta_u and the \gamma_u = \cosh \zeta_u. For two successive collinear boosts with rapidities \zeta_u and \zeta_v, the composite rapidity is simply \zeta_w = \zeta_u + \zeta_v, yielding \tanh \zeta_w = \tanh(\zeta_u + \zeta_v). However, for non-collinear boosts, the composition introduces a rotation in the plane perpendicular to the bisector of the rapidities, with the Wigner angle \theta given by \tan(\theta/2) = \frac{\sinh \zeta_u \sinh \zeta_v \sin \phi}{ \cosh \zeta_u + \cosh \zeta_v + (\cosh \zeta_u \cosh \zeta_v - 1)(1 - \cos \phi) }, where \phi is the angle between \mathbf{u} and \mathbf{v}; this arises from the non-commutativity of boosts in rapidity space, analogous to anholonomy in . Another method utilizes representations of the , which compactly encode boosts and rotations. In this framework, a pure boost along direction \mathbf{n} with \zeta is represented by the B = \cosh(\zeta/2) + \mathbf{n} \sinh(\zeta/2), where \mathbf{n} is the vector. The of two non-collinear boosts B_u B_v yields a product that decomposes into a boost followed by a spatial R = \cos(\theta/2) + \hat{\mathbf{k}} \sin(\theta/2), where \hat{\mathbf{k}} is the perpendicular to the plane of \mathbf{u} and \mathbf{v}, and the angle \theta matches the Wigner rotation; this multiplication directly reveals the rotational component without explicit matrix inversion. A related derivation leverages the Euler angle parametrization of the Lorentz group SO(3,1), decomposing a general proper orthochronous transformation into a product of boosts and rotations analogous to the ZYZ Euler angles for SO(3). Specifically, any Lorentz transformation can be expressed as a boost in the z-direction, followed by a rotation around z, another z-boost, and further rotations, with the Wigner rotation emerging from the misalignment in the non-commuting boost sequence; the angles are solved from the trace and determinant conditions of the transformation matrix, yielding the rotation parameters explicitly. At the Lie algebra level, the Wigner rotation originates from the non-vanishing of boost generators, such as [K_x, K_y] = -i J_z, where K_x, K_y generate boosts along the x- and y-axes, and J_z generates rotations around z; this relation implies that finite non-collinear boosts, approximated via the Baker-Campbell-Hausdorff formula, include a rotational term proportional to the commutator. In the infinitesimal limit of successive small boosts, corresponding to accelerated motion under the Lorentz force, the Wigner rotation reduces to the , a continuous rotation of the spin frame at angular velocity \boldsymbol{\omega}_T = -\frac{\gamma^2}{(\gamma + 1) c^2} \mathbf{v} \times \mathbf{a}, where \mathbf{v} is velocity and \mathbf{a} is acceleration, providing the relativistic correction to spin-orbit coupling.

Group-Theoretic Foundations

Lorentz Group Structure

The , denoted SO(3,1), consists of linear transformations that preserve the Minkowski metric in four-dimensional spacetime, forming a six-dimensional . It can be understood as a semi-direct product SO(3,1) ≅ SO(3) ⋊ B, where SO(3) is the of spatial rotations and B is the of pure boosts (hyperbolic rotations in space-time planes), with the non-abelian structure arising from the action of rotations on boosts by conjugation. This semi-direct product captures the essential asymmetry between rotations, which form a compact , and boosts, which do not, leading to the group's overall non-compact nature. For applications in , the relevant component is the proper orthochronous Lorentz group SO⁺(3,1), which includes transformations with positive and that preserve the orientation of time (Λ⁰₀ ≥ 1), ensuring continuity with the identity and excluding or time-reversal operations. This subgroup is connected and simply covers the physically realizable symmetries of . Elements of the group are parametrized via the from its : a general is given by \Lambda = \exp\left( \frac{i}{2} \omega_{\mu\nu} M^{\mu\nu} \right), where ω_{μν} = -ω_{νμ} is the antisymmetric parameter tensor, and M^{μν} are the generators satisfying the Lorentz algebra commutation relations. The generators decompose into spatial rotations and boosts, with M^{0i} = K_i for i=1,2,3 denoting the boost generators in the spatial directions, and M^{ij} = \epsilon^{ijk} J_k for the rotation generators around the k-axis. This parametrization highlights the six independent parameters: three for rotations and three for boosts. A key feature of the group's non-abelian structure is that the boost generators K_i do not commute unless they are , as their yields a : non-collinear boosts compose to produce an additional component, which is central to the Wigner rotation phenomenon.

Boost Generators and Commutators

The of the SO(3,1) is spanned by the rotation generators J_i and boost generators K_i (i=1,2,3), which satisfy the commutation relations \begin{align*} [J_i, J_j] &= i \epsilon_{ijk} J_k, \ [J_i, K_j] &= i \epsilon_{ijk} K_k, \ [K_i, K_j] &= -i \epsilon_{ijk} J_k. \end{align*} These relations encode the structure of the , where the non-vanishing [K_i, K_j] demonstrates that boosts along non-parallel directions do not commute, generating a via the right-hand side. A finite boost along direction \hat{u} with \zeta_u is represented as \exp(-i \zeta_u \hat{u} \cdot \vec{K}), where \vec{K} = (K_1, K_2, K_3). The composition of two such non-collinear infinitesimal boosts, \exp(-i \zeta_v \hat{v} \cdot \vec{K}) \exp(-i \zeta_u \hat{u} \cdot \vec{K}), can be analyzed using the Baker-Campbell-Hausdorff formula. To leading order, this yields \exp(-i \zeta_w \hat{w} \cdot \vec{K} - i \theta \hat{n} \cdot \vec{J}), where the angle \theta arises from the term [ -i \zeta_v \hat{v} \cdot \vec{K}, -i \zeta_u \hat{u} \cdot \vec{K} ] / 2. In vector notation, the component is \theta \hat{n} = -\frac{1}{2} (\zeta_u \hat{u} \times \zeta_v \hat{v}), or equivalently, \theta = - (\zeta_u \times \zeta_v) \cdot \hat{n} with \hat{n} the unit vector along the . For boosts confined to a , the relevant subgroup is SO(2,1)^+, the proper orthochronous in (2+1)-dimensional . This subgroup admits an Euler-like parametrization, decomposing any element into a -rotation- sequence: B(\xi_2) R([\phi](/page/Phi)) B(\xi_1), where B(\xi) denotes a with \xi and R([\phi](/page/Phi)) a by [\phi](/page/Phi). The central [\phi](/page/Phi) corresponds precisely to the Wigner rotation emerging from the non-commutativity of the bounding . This decomposition provides a way to isolate the rotational component in planar Lorentz transformations.

Historical Development

Early Discoveries

The phenomenon of rotation arising from the composition of non-collinear Lorentz boosts, later termed the Wigner rotation, was first theoretically predicted in the early years of as researchers explored the of . In 1913, examined the relativistic composition of velocities using a geometric approach based on spaces, where he identified a kinematic rotation for a body in curvilinear or orbital motion, arising from the non-Euclidean nature of velocity space. This insight appeared in his note "La théorie de la relativité et la cinématique," marking the initial recognition of the effect in the context of resolving paradoxes in velocity superposition. Building on this, Ludwik Silberstein in 1914 rigorously derived the rotational component in his book The Theory of Relativity, demonstrating that the product of two successive boosts in different directions yields not only a net boost but also an intrinsic rotation of the spatial frame. Silberstein's analysis emphasized the mathematical structure of Lorentz transformations, showing how this rotation emerges inevitably from the group's properties, though without a full group-theoretic interpretation at the time. The physical implications became clear in 1926 with Llewellyn Thomas's application to dynamics. Thomas showed that the rotation, which he called , must be accounted for in the spin-orbit interaction of an orbiting a , introducing a factor of approximately 1/2 that reconciles the relativistic calculation of the g-factor with experimental observations of atomic . This work resolved a key paradox in by incorporating the into the electron's rest frame. These early findings stemmed from efforts to align with , particularly in addressing discrepancies like the electron's . In 1939, generalized the rotation in through his classification of unitary representations of the , revealing its role in the transformation properties of particle states.

Key Contributions and Rediscoveries

The Wigner rotation gained prominence in the quantum mechanical context through 's 1939 analysis of unitary representations of the , where it emerged as the spatial rotation component arising from the composition of non-collinear boosts in the transformation of particle states. Although the phenomenon was earlier identified in classical by Llewellyn Thomas in 1926 for infinitesimal boosts, Wigner's work extended its application to finite transformations in , leading to the convention of attributing the general finite case to Wigner while reserving "Thomas precession" specifically for the infinitesimal limit or the precessional rate in accelerated frames. Post-1930s literature saw ongoing debates regarding sign conventions in the formulas for the angle and axis, particularly in distinguishing active versus passive interpretations and resolving inconsistencies between classical and quantum derivations. Rindler, in his 1986 volume on spinors and space-time geometry, addressed these issues by clarifying the geometric role of the within the Lorentz group's structure, emphasizing consistent sign choices aligned with the . Earlier formulas from the 1930s and 1940s often contained sign errors due to varying conventions for parametrization, which were systematically resolved in the through group-theoretic analyses that standardized the 's expression as a pure SO(3) . In 2002, Herbert Goldstein and colleagues highlighted the paradoxical nature of the Wigner rotation in their textbook on , comparing it to the as a counterintuitive consequence of that challenges naive notions of velocity addition and frame synchronization. These discussions culminated in corrections to erroneous equations in prior pedagogical treatments, as noted in Krzysztof Rebilas's 2013 comment, which proposed a straightforward method to attribute formal calculations unambiguously to either the full Wigner rotation or its Thomas limit, thereby unifying the nomenclature and resolving lingering ambiguities.

Interpretations and Applications

Relation to Thomas Precession

The arises as the limit of the Wigner rotation when a particle undergoes continuous, non-collinear changes, such as in accelerated motion along a curved trajectory. In this context, successive result in a cumulative rotation of the particle's , manifesting as a of the relative to the lab frame. This effect is a direct consequence of the non-commutativity of non-collinear boosts in the , where the Wigner rotation angle for finite boosts becomes a rate under small increments. The angular velocity of the Thomas precession is given by \vec{\omega}_T = \frac{\gamma^2}{\gamma + 1} \frac{\vec{a} \times \vec{v}}{c^2}, where \gamma = (1 - v^2/c^2)^{-1/2} is the , \vec{v} is the particle's velocity, \vec{a} = d\vec{v}/dt is its , and c is the . This expression links the precession to the relativistic kinematics of spin transport, as described by the Bargmann-Michel-Telegdi equation, which governs the evolution of the spin four-vector under . In the non-relativistic (v \ll c), it approximates to \vec{\omega}_T \approx \frac{1}{2c^2} \vec{v} \times \vec{a}, highlighting its origin in the geometry of velocity space. Physically, the Thomas precession represents a rotation of the instantaneous rest frame of a spinning particle during circular or curved motion, arising from the need to parallel-transport the spin vector along the worldline while accounting for frame non-orthogonality due to relativity. For an electron in orbital motion, this precession rate matches half the rate expected from the naive spin-orbit magnetic interaction in the rest frame, effectively reducing the total spin-orbit coupling by a factor of 1/2. This correction explains why the electron's gyromagnetic ratio g = 2, as predicted by the , yields the observed fine structure splitting in hydrogen-like atoms when combined with the Thomas effect—without it, the relativistic Dirac prediction would overestimate the splitting by exactly twice the correct value.

Modern Uses in Physics

In particle physics, Wigner rotation plays a key role in describing the evolution of particle spin polarization during acceleration in high-energy colliders such as the Large Hadron Collider (LHC). It accounts for the additional rotations induced by successive non-collinear Lorentz boosts, which affect the transverse and longitudinal polarization states of beams, ensuring accurate predictions of spin-dependent scattering processes. For instance, in analyses of decays involving polarized particles, the Wigner D-matrix is employed to transform between helicity frames and total momentum systems, directly impacting measurements of asymmetries in events like B \to K^* \mu^+ \mu^-. This is particularly important for experiments probing new physics beyond the Standard Model, where spin correlations must be precisely modeled to interpret polarization data. Experimental confirmations of Wigner rotation have been obtained through observations in storage rings, where it manifests in conjunction with during . In the 1970s experiments, the anomalous spin precession frequency was measured with a precision of 7.3 parts per million, verifying the relativistic corrections including at the expected rate for muons at magic gamma (\gamma \approx 29.3). These results, derived from decay asymmetries in a uniform , aligned with theoretical predictions and provided early validation of the combined effects in accelerated charged particles. Similar verifications in electron storage rings, such as those studying and self-polarization, have further corroborated the rotation's influence on spin dynamics. In contemporary , post-2000 simulations of entangled particles incorporate Wigner rotations to capture the degradation of entanglement under Lorentz transformations. For massive particles in Bell states, these rotations induce momentum-dependent flips, quantified through measures like , revealing Lorentz invariance of entanglement despite frame changes. Such computations are essential for developing quantum communication in moving frames, linking back to the -focused interpretation via in one sentence: Wigner rotation extends the Thomas precession framework to entangled systems by introducing frame-dependent correlations.

Visual and Geometric Representations

Spacetime Diagrams

Spacetime diagrams provide an intuitive geometric visualization of the Wigner rotation arising from non-collinear Lorentz boosts in . These diagrams typically depict the worldlines of the origins of three inertial frames: the laboratory frame Σ, an intermediate frame Σ' obtained by boosting Σ along a u, and a final frame Σ'' reached by boosting Σ' along a non-parallel v. The worldlines trace the paths of these origins over time, with spatial axes represented as lines of constant time in each frame, revealing how successive boosts tilt these axes relative to one another. The rotation is illustrated by comparing the orientation of spatial axes in Σ'' to what would result from of the axes from Σ without rotation. In the , the axes in Σ'' appear rotated by the Wigner due to the non-commutativity of boosts, with the tilt manifesting as a discrepancy between the boosted frame's simultaneity and the original frame's. For instance, in a setup involving a Born-rigid object undergoing a closed of boosts, the final frame's axes show a clear rotational offset, such as 14.4° for a Lorentz factor γ = 2/√3, highlighting the geometric shear induced by the composition of boosts. This visualization underscores how the rotation emerges from the hyperbolic structure of , where trajectories curve in the laboratory frame but remain straight in instantaneous comoving frames. A specific example in the planar case confines the boosts to the xy-plane, with u along the x-direction and v at an angle in the xy-plane. The diagram plots the worldlines segmented into boost intervals, often including light cones at key events to delineate the and differences. For a boost parameter β = 0.7, the resulting Wigner rotation reaches approximately 33.7°, as seen in the misalignment of a grid's spatial positions across frames, demonstrating the rotation's dependence on the angle between velocities. Such diagrams emphasize the underlying , where the non-Euclidean metric of causes the apparent rotation through the during frame changes.

Parametrizations and Examples

The Euler angle parametrization provides a useful decomposition for elements of the proper Lorentz group SO+(2,1) in (2+1)-dimensional Minkowski space, analogous to the standard Euler angles for SO(3). A reflection-free Lorentz transformation can be expressed in the boost-rotation-boost form B(\psi, -\eta) R(\phi - \psi), where B(\psi, -\eta) is a boost of rapidity \eta at angle \psi to the x-axis, and R(\alpha) is a rotation by angle \alpha around the z-axis. The angles are determined explicitly from the matrix elements of the transformation: \cosh \eta = L_{33}, \cos \psi = L_{31} / \sqrt{L_{33}^2 - 1}, \sin \psi = L_{32} / \sqrt{L_{33}^2 - 1}, \cos \phi = -L_{31} / \sqrt{L_{33}^2 - 1}, and \sin \phi = -L_{32} / \sqrt{L_{33}^2 - 1}, with the Wigner rotation angle given by \theta = \phi - \psi. This parametrization simplifies the analysis of successive boosts, revealing the Wigner rotation as the difference in the pre- and post-boost orientation angles. A pedagogical example arises from two successive perpendicular boosts of equal speed, say \vec{u} = \beta c \hat{x} and \vec{v} = \beta c \hat{y} with \beta = |\vec{u}|/c = |\vec{v}|/c, corresponding to rapidities \eta_1 = \eta_2 = \artanh \beta and Lorentz factor \gamma = 1/\sqrt{1 - \beta^2}. The composition B(\vec{v}) B(\vec{u}) decomposes into a boost in the direction \hat{n} = (\hat{x} + \hat{y})/\sqrt{2} followed (or preceded) by a rotation around the z-axis by angle \theta, where \cos \theta = 2\gamma / (\gamma^2 + 1) or equivalently \tan \theta = \gamma \beta^2 / 2. The total boost has rapidity \eta_w = 2 \eta and magnitude |\vec{w}| = c \tanh(2 \eta) = c (2 \beta \gamma) / (\gamma^2 + 1). This case illustrates the non-commutativity of non-collinear boosts, with \theta increasing from 0 (non-relativistic limit) toward \pi/2 as \beta \to 1. For a numerical instance, consider unequal perpendicular boosts \vec{u} = 0.8c \hat{x} (\gamma_u = 5/3 \approx 1.667) and \vec{v} = 0.6c \hat{y} (\gamma_v = 5/4 = 1.25). The composition yields a total velocity \vec{w} with components w_x = 0.8c, w_y = 0.6c / \gamma_u = 0.36c, so |\vec{w}| \approx 0.877c (\gamma_w = \gamma_u \gamma_v \approx 2.083) and direction \hat{n} at angle \alpha = \arctan(w_y / w_x) \approx 24.2^\circ from the x-axis. The Wigner rotation angle is \theta = \arccos[ (\gamma_u + \gamma_v) / (1 + \gamma_u \gamma_v) ] \approx 19.2^\circ around the z-axis. In the ultra-relativistic limit where both boosts approach the (\gamma_u, \gamma_v \gg 1), the Wigner rotation angle \theta for perpendicular configurations approaches \pi/2, as \cos \theta \to 0.

References

  1. [1]
    Wigner rotation and Euler angle parametrization - AIP Publishing
    Jul 1, 2023 · Two successive boosts are equivalent to a boost followed or preceded by a rotation, which may be taken as the definition of Wigner rotation.
  2. [2]
    [PDF] an elementary derivation of the Wigner rotation
    Sep 22, 2010 · One of the most puzzling phenomena in special relativity is the composition of boosts. When one contemplates the form of an arbitrary boost ...
  3. [3]
    [PDF] ON UNITARY REPRESENTATIONS OF THE INHOMOGENEOUS ...
    This result will hold for all representations of the inhomogeneous Lorentz group with positive P, since we shall show that even the infinite dimensional ...
  4. [4]
    [PDF] Thomas-Wigner rotation and Thomas precession: actualized approach
    Abstract: We show that the explanation of Thomas-Wigner rotation (TWR) and Thomas preces- sion (TP) in the framework of special theory of relativity (STR) ...
  5. [5]
    [PDF] Elementary analysis of the special relativistic combination of ...
    This includes a description of what such velocities actually “mean”, what the Wigner rotation represents, and how this leads to the Thomas precession. In ...
  6. [6]
    [PDF] arXiv:1504.07432v1 [quant-ph] 28 Apr 2015
    Apr 28, 2015 · The Wigner rotation is known in special relativity as the product of two Lorentz boots in different directions which gives rise to a boost ...
  7. [7]
    [PDF] Wigner rotation and its SO(3) model: an active-frame approach - HAL
    Jun 24, 2021 · In special relativity, a boost transformation takes place between two inertial frames, hence each boost is defined by a constant velocity ...
  8. [8]
    None
    ### Summary of Physical Setup and Wigner Rotation (arXiv:2206.12406)
  9. [9]
    [PDF] Wigner Rotation: Theory and Application to Practical Relativistic ...
    Mar 19, 2019 · The re- sults for the Wigner rotation in the Lorentz lab frame obtained by many experts on special relativity such as Moeller and Jackson, are ...
  10. [10]
    [PDF] Visualization of Thomas-Wigner rotations - arXiv
    Jun 6, 2017 · The visualization highlights the close relationship between the Thomas-Wigner rotation and the relativity of simultaneity. ... Special Relativity ...
  11. [11]
    5.6 Relativistic Velocity Transformation - University Physics Volume 3
    Sep 29, 2016 · The kayak's velocity is the vector sum of its velocity relative to ... Do the calculation: u = v + u ′ 1 + v u ′ c 2 = 0.500 c + c 1 + ...
  12. [12]
    How do You Add Velocities in Special Relativity?
    How Do You Add Velocities in Special Relativity? Suppose an object A is moving with a velocity v relative to an object B, and B is moving with a velocity u (in ...Missing: c²) | Show results with:c²)
  13. [13]
    [PDF] 1.2 Einstein Velocity Addition
    γ u⊕v = γu γv. 1 − u·v c2. (1.8) for all u,v ∈ Rn c . Here, (1.8) is ... γu = 1. 1 − u2. 1+u2. 2+u2. 3 c2. (1.21). The three components of Einstein ...Missing: c²) | Show results with:c²)
  14. [14]
    an elementary derivation of the Wigner rotation - IOPscience
    Generic composition of boosts: an elementary derivation of the Wigner rotation. Rafael Ferraro and Marc Thibeault. Published under licence by IOP Publishing ...
  15. [15]
  16. [16]
  17. [17]
    The Wigner angle as an anholonomy in rapidity space - AIP Publishing
    Jul 1, 1997 · This paper demonstrates how the Wigner angle and final boost parameters can be calculated simply in terms of the initial boost parameters using ...
  18. [18]
    [PDF] Relativistic combination of non-collinear 3-velocities using quaternions
    Mar 13, 2020 · This rotation, known as the Wigner rotation, was first discovered by Llewellyn Thomas in 1926 whilst trying to describe the Zeeman effect from a ...
  19. [19]
  20. [20]
    [PDF] Thomas Precession for Dressed Particles - arXiv
    Jan 26, 2018 · It is the differential of a Wigner rotation ... composition of a sequence of infinitesimal Wigner rotations; it is a generalization of the.
  21. [21]
    [PDF] Useful Notes for the Lorentz Group
    We study vectorial quantities, such as velocity, momentum, force, etc. We need to be able to combine different velocities, and also other systems of vectors ...
  22. [22]
    [PDF] Exercises on Theoretical Particle Physics
    Oct 22, 2010 · Λ = exp. (. − i. 2. ωµν Mµν. ) . In exercise H 1.1 we have defined the Lorentz algebra through. [Mµν,Mρσ] = −i(ηµρMνσ − ηµσMνρ − ηνρMµσ + ...
  23. [23]
    Lorentz group and its representations - Book chapter - IOPscience
    When the three boost generators are added to the rotation group, these six generators also form a closed set of commutation relations, which is the Lie algebra ...Abstract · Generators of the Lorentz group · Two-by-two representation of...
  24. [24]
    [PDF] 1 Rotations - University of Oregon
    However, the transformation law under general Lorentz transformations is more complicated: the spin indices get transformed according to the Wigner rotation. 18 ...
  25. [25]
    History of special relativity - Wikipedia
    The history of special relativity consists of many theoretical results and empirical findings obtained by Albert A. Michelson, Hendrik Lorentz, Henri Poincaré ...<|control11|><|separator|>
  26. [26]
    A Method of É. Borel for calculation of the Thomas precession
    Mar 2, 2013 · In 1913, É. Borel (1871–1956) considered the relativistic kinematic rotation of a body that arises in the course of its orbital motion.Missing: composition | Show results with:composition
  27. [27]
    The theory of relativity : Silberstein, Ludwik, b. 1872 - Internet Archive
    Nov 30, 2006 · The theory of relativity. by: Silberstein, Ludwik, b. 1872. Publication date: 1914. Topics: Relativity (Physics). Publisher: London : Macmillan.Missing: rotation | Show results with:rotation
  28. [28]
  29. [29]
    [PDF] Thomas Precession and the BMT equation
    dSµ dτ + 1 c2 uµ duν dτ − uν duµ dτ Sν = 0 . The above equation is known as the equation for Fermi-Walker transport, and describes the trajectory of a ...
  30. [30]
    On the classical analysis of spin-orbit coupling in hydrogenlike atoms
    Apr 1, 2010 · The classical analysis of spin-orbit coupling in hydrogenlike atoms by Thomas1 played an important role in supporting the hypothesis of electron ...<|control11|><|separator|>
  31. [31]
  32. [32]
    [PDF] Muon (g − 2): experiment and theory - CLASSE (Cornell)
    Apr 23, 2007 · A review of the experimental and theoretical determinations of the anomalous magnetic moment of the muon is given. The anomaly is defined by ...
  33. [33]
    [PDF] An Introduction to Spin Polarisation in Accelerators and Storage Rings
    A first look at the stability of spin motion and spin gymnastics with spin rotators. 4. Electrons: the effect of synchrotron radiation I -- self polarisation. 5 ...
  34. [34]
    None
    Below is a merged summary of Wigner rotation in the context of GPS or satellite boosts, consolidating all the information from the provided segments into a single, dense response. To maximize detail retention, I’ll use a table in CSV format to organize key information systematically, followed by a narrative summary that ties it all together. The table will capture the essence of each segment, including context, relevance, key formulas, and URLs, while the narrative will provide a cohesive overview.
  35. [35]
    [2204.10395] Effect of Wigner rotation on estimating unitary-shift ...
    Apr 21, 2022 · Abstract page for arXiv paper 2204.10395: Effect of Wigner rotation on estimating unitary-shift parameter of relativistic spin-1/2 particle.Missing: 2020s | Show results with:2020s
  36. [36]
    [2411.09890] Wigner function under changes of reference frames
    Nov 15, 2024 · In this paper, we investigate the transformation laws of the Wigner function under changes of reference frames.
  37. [37]
    Wigner rotations and an apparent paradox in relativistic quantum ...
    Apr 3, 2013 · A Wigner rotation corresponds to a momentum-dependent change of the spin state of a relativistic particle with a change of reference frame [1, 2]Missing: Goldstein aspects
  38. [38]
    [2206.12406] Wigner rotation and Euler angle parametrization - arXiv
    Jun 22, 2022 · Title:Wigner rotation and Euler angle parametrization. Authors:Leehwa Yeh · Download PDF. Abstract: Analogous to the famous Euler angle ...
  39. [39]
    [PDF] Wigner rotation and Euler angle parametrization - HAL
    Dec 28, 2023 · As a demonstration of its effectiveness, we show how simple it is to derive important rules about the Wigner rotation problem (Sec. III) and how ...