Fact-checked by Grok 2 weeks ago

Rindler coordinates

Rindler coordinates are a system of coordinates in that describe the geometry of flat Minkowski from the perspective of an observer undergoing constant , resulting in hyperbolic motion. Discussed by physicist Wolfgang Rindler, notably in his 1960 book and 1966 paper, they provide a framework for analyzing the experiences of such an observer, revealing features like a causal horizon that limits the observer's access to information from certain regions of . The transformation from standard inertial (Minkowski) coordinates (ct, x, y, z) to Rindler coordinates (\bar{t}, \bar{x}, \bar{y}, \bar{z}) for an observer with constant proper acceleration g is given by
ct = \bar{x} \sinh\left(\frac{g \bar{t}}{c}\right), \quad x = \bar{x} \cosh\left(\frac{g \bar{t}}{c}\right), \quad y = \bar{y}, \quad z = \bar{z},
where the accelerating observer remains at fixed spatial coordinates \bar{x} = c^2/g, \bar{y} = 0, \bar{z} = 0, and \bar{t} corresponds to the observer's proper time. In these coordinates, the line element (spacetime metric) takes the form
ds^2 = -\frac{g^2 \bar{x}^2}{c^2} d\bar{t}^2 + d\bar{x}^2 + d\bar{y}^2 + d\bar{z}^2,
which highlights the non-uniform spatial geometry perceived by the observer and the presence of a Rindler horizon at \bar{x} = 0, analogous to event horizons in curved spacetimes of general relativity. This horizon arises because the worldlines of constant \bar{x} are hyperbolae in Minkowski space, confining the observable universe for the accelerated observer to the region \bar{x} > 0.
Rindler coordinates play a crucial role in bridging special and general relativity, particularly in discussions of equivalence principles and accelerated frames. They are essential for understanding quantum effects in accelerated frames, such as the , where an accelerating observer detects a thermal bath of particles in the Minkowski vacuum with temperature T = \frac{\hbar g}{2\pi k_B c}, linking in flat space to . Applications extend to studies of analogs, detector responses in quantum fields, and holographic dualities in .

Coordinate Definitions

Transformation to Minkowski coordinates

Rindler coordinates provide a system adapted to observers undergoing constant in flat Minkowski . The standard Minkowski coordinates (t, x, y, z) use the \eta_{\mu\nu} = \operatorname{diag}(-1, 1, 1, 1), where the is ds^2 = -dt^2 + dx^2 + dy^2 + dz^2 (with c = 1). Lorentz boosts in the x-direction preserve this and connect inertial frames related by velocity v, given by t' = \gamma (t - v x), x' = \gamma (x - v t), y' = y, z' = z, with \gamma = (1 - v^2)^{-1/2}. To derive the transformation for the right Rindler wedge, consider an observer with constant proper acceleration a, measured along their worldline. The proper acceleration is the magnitude of the four-acceleration vector, which is orthogonal to the four-velocity and satisfies u^\mu a_\mu = 0, with a^\mu a_\mu = a^2. For motion along the x-axis, the worldline satisfies the hyperbolic relation x^2 - t^2 = (1/a)^2, ensuring constant proper acceleration a at the point where \rho = 1/a. Parametrizing by proper time \tau, the coordinates are t = \frac{1}{a} \sinh(a \tau), x = \frac{1}{a} \cosh(a \tau), y = 0, z = 0. This follows from the conditions of constant proper acceleration along the worldline, yielding the hyperbolic trajectory. Extending to the full coordinate system, lines of constant spatial coordinate \rho > 0 are orthogonal to the worldlines, with \rho scaling the hyperbolae. The rapidity parameter \alpha = a \tau serves as the timelike coordinate, uniform across simultaneity surfaces. Thus, the transformation for the right wedge becomes t = \rho \sinh \alpha, x = \rho \cosh \alpha, y = y, z = z, where \alpha ranges from -\infty to \infty and \rho > 0. This covers the region x > |t| of Minkowski spacetime. Minkowski spacetime divides into four Rindler wedges: the right wedge (I: x > |t|), left wedge (II: x < -|t|), future wedge (III: t > |x|), and past wedge (IV: t < -|x|). These wedges exclude the t- and x-axes, which are horizons for the accelerated observers, and together they span the entire excluding the origin's light cone boundaries. The right and left wedges are timelike, suitable for observer families, while future and past are spacelike. Substituting the transformation into the Minkowski metric yields the Rindler form. Differentiating gives dt = d\rho \sinh \alpha + \rho \cosh \alpha \, d\alpha and dx = d\rho \cosh \alpha + \rho \sinh \alpha \, d\alpha. Then, -dt^2 + dx^2 = -(d\rho \sinh \alpha + \rho \cosh \alpha \, d\alpha)^2 + (d\rho \cosh \alpha + \rho \sinh \alpha \, d\alpha)^2. Expanding and simplifying using hyperbolic identities \cosh^2 \alpha - \sinh^2 \alpha = 1 and -\sinh^2 \alpha + \cosh^2 \alpha = 1 (with cross terms canceling) results in - \rho^2 d\alpha^2 + d\rho^2. Including transverse directions, the line element is ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + dz^2. The factor \rho^2 in the time term acts as a conformal factor, but the space remains flat, as the transformation is a coordinate change in Minkowski spacetime.

Variant transformation formulas

In addition to the standard transformation for the right Rindler wedge, several variant formulas provide equivalent mappings between Minkowski and Rindler coordinates, each tailored to emphasize different aspects of the geometry or facilitate specific calculations. A rapidity-based variant reparameterizes the transformation using a boost parameter \phi as the timelike coordinate, where the spatial coordinate \rho plays the role of a radial-like parameter. The relations are given by t = \rho \sinh \phi, \quad x = \rho \cosh \phi, with the identification \phi = \alpha, where \alpha denotes the Rindler timelike coordinate. This form highlights the hyperbolic nature of the boost, aligning directly with Lorentz transformations parameterized by rapidity. For the left Rindler wedge, which describes observers undergoing acceleration in the opposite direction (toward negative x), the transformation adjusts the signs to reflect the mirrored geometry: t = -\rho \sinh \alpha, \quad x = -\rho \cosh \alpha. This variant covers the region x < -|t| in , maintaining the same metric form but reversing the spatial orientation relative to the right wedge. An exponential variant, derived by expressing the transformation in terms of null (light-cone) coordinates u = t - x and v = t + x, takes the form x + t = \rho e^{\alpha}, \quad x - t = \rho e^{-\alpha}. Solving for t and x yields the hyperbolic relations but proves useful for analyzing or horizon structures directly in additive coordinates. These variants offer computational advantages in distinct contexts; for instance, the rapidity-based and standard hyperbolic forms simplify analytic solutions for observer worldlines and proper acceleration, while the exponential form aids numerical simulations of wave propagation near horizons by avoiding trigonometric functions. In geodesic integration, the exponential variant reduces complexity for lightlike paths, though it may require additional steps for timelike computations compared to the direct sinh/cosh expressions.
VariantKey Formulas for t, xProsCons
Rapidity-based (right)t = \rho \sinh \phi, x = \rho \cosh \phi (\phi = \alpha)Intuitive link to Lorentz boosts; easy for acceleration profilesRequires hyperbolic identities for null coordinates
Left wedget = -\rho \sinh \alpha, x = -\rho \cosh \alphaCovers opposite acceleration; symmetric to right wedgeSign flips can complicate unified simulations across wedges
Exponential (null-derived)x + t = \rho e^{\alpha}, x - t = \rho e^{-\alpha}Simplifies horizon and null geodesic calculations; avoids trig functionsLess direct for timelike distances; inversion to Cartesian may introduce logs

Observer Worldlines

Rindler observers and hyperbolic motion

The Rindler observers constitute a family of uniformly accelerating observers in flat Minkowski spacetime whose worldlines define the Rindler coordinate frame. These worldlines trace out branches of hyperbolas in the (t, x) plane of inertial Minkowski coordinates (with c = 1 and y = z = 0). For an observer at fixed spatial Rindler coordinate ρ > 0, the parametric equations are t(\tau) = \rho \sinh\left(\frac{\tau}{\rho}\right), \quad x(\tau) = \rho \cosh\left(\frac{\tau}{\rho}\right), where τ is the measured by the observer. These satisfy the invariant equation x^2 - t^2 = \rho^2, representing a centered at the with ρ. Each observer experiences constant α = 1/ρ, directed along the positive x-axis, where ρ measures the observer's "distance" from the , which acts as a horizon for the family. This acceleration ensures the observer never reaches the asymptotically, maintaining a . The family forms an orthogonal congruence of timelike curves, meaning the worldlines are everywhere orthogonal to a family of spacelike hypersurfaces of . In Minkowski coordinates, the 4-velocity of all observers at a given Rindler time α is u^\mu = (\cosh \alpha, \sinh \alpha, 0, 0), normalized such that u^\mu u_\mu = -1. Proper times along the worldlines are synchronized via the Rindler time coordinate α, defined such that α = τ / ρ for the observer at ρ, ensuring simultaneous events across the frame share the same α. For a reference observer at ρ₀ with acceleration a = 1/ρ₀, this corresponds to α = a τ, and the synchronization extends rigidly to other observers by scaling their proper times proportionally to ρ. Observers at larger ρ thus experience smaller proper accelerations but perceive fixed proper distances to neighbors, maintaining a rigid spatial configuration in the Rindler frame despite the varying accelerations. This structure arises because the proper distance between observers at ρ and ρ + dρ is simply dρ, independent of α. The line element experienced by these observers is the Rindler metric.

Characteristics of the Rindler frame

The Rindler frame is constructed from a of timelike worldlines traced by observers undergoing uniform in . This is geodesic-free, with each worldline characterized by a constant proper acceleration magnitude inversely proportional to the spatial coordinate \rho. The kinematic properties of the include zero expansion scalar \theta = 0, zero shear tensor \sigma_{\mu\nu} = 0, and zero rotation tensor \omega_{\mu\nu} = 0, rendering it Born-rigid such that infinitesimal volumes orthogonal to the worldlines remain constant under evolution. The acceleration vector field is a^\mu = (0, 1/\rho, 0, 0) in the coordinate basis (\alpha, \rho, y, z), pointing radially outward and determining the local proper acceleration a = 1/\rho. The in Rindler coordinates ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + [dz](/page/DZ)^2 is of the time coordinate \alpha, conferring time-translation invariance under shifts in \alpha. This static property positions the Rindler frame as an effective for systems of uniformly accelerating observers, where stationary configurations relative to the experience no temporal evolution in the metric structure. Rindler coordinates parametrize only the right of Minkowski , defined by x > |t| > 0, excluding regions causally disconnected from the . Covering the full Minkowski necessitates four distinct Rindler patches, each associated with one of the four wedges divided by the light cones at the . By the , the experiences of observers in the Rindler frame locally replicate those in a uniform , with the g = 1/\rho acting as an effective antiparallel to the \rho-direction. This analogy underscores the tidal-free uniformity of the field within the , absent the present in genuine gravitational contexts. The rigidity of spatial structures in the Rindler frame, such as "rods" spanning observers at fixed \rho, is preserved via Fermi-Walker transport of basis vectors along the worldlines, preventing spurious rotations from the and maintaining alignment with the instantaneous rest frames.

Paradoxical property of acceleration

In Rindler coordinates, observers following worldlines of constant spatial coordinate \rho undergo hyperbolic motion in the underlying Minkowski spacetime, maintaining a constant \alpha = 1/\rho (in units where c = 1) for each individual observer. This represents the physical felt locally by the observer, measured by an at rest relative to their worldline. However, when viewed from the inertial Minkowski frame, the coordinate \ddot{x} along these worldlines varies with position and time. Specifically, for a given worldline satisfying x^2 - t^2 = \rho^2, the coordinate is \ddot{x} = \rho^2 / x^3, which decreases as $1/x^3 for large x (or equivalently, as the observer approaches the asymptotically). Despite this apparent deceleration in coordinate terms, the remains invariant and constant due to the relativistic factor \alpha = \gamma^3 \ddot{x}, where \gamma = 1/\sqrt{1 - v^2} accounts for the increasing Lorentz contraction and . This discrepancy gives rise to a counterintuitive aspect of the Rindler frame: the coordinate effects, including the decreasing \ddot{x}, stem from the between the accelerating frame and the inertial frame. Events simultaneous in the Rindler frame (constant Rindler time \tau) are not simultaneous in Minkowski coordinates, leading to a desynchronization that manifests as varying coordinate accelerations even though each observer experiences unchanging locally. The is a boost-invariant quantity tied to the curvature of the worldline, ensuring that the "felt" force remains steady regardless of the observer's in the external frame. Consider two observers in the Rindler frame at different \rho, say \rho_1 < \rho_2. The inner observer (at smaller \rho_1) has higher proper acceleration \alpha_1 = 1/\rho_1 > \alpha_2 = 1/\rho_2, feeling a stronger g-force, while the outer observer requires less coordinate thrust in the Minkowski frame to maintain its motion but experiences a weaker local acceleration. This variation ensures the frame's Born rigidity: proper distances between observers remain constant along simultaneous spatial hypersurfaces in the instantaneous comoving inertial frame, preventing stretching or compression during acceleration. Without this differential acceleration, the structure would not remain rigid relativistically. This property of differential yet constant proper accelerations puzzled early relativists exploring in accelerated frames. Max introduced the concept of Born rigidity in 1909 while analyzing how extended bodies could accelerate without deforming in , concluding that only motion satisfies the condition of unchanging proper lengths in the body's at each instant. Wolfgang Rindler later formalized these coordinates in 1966, highlighting their role in resolving such kinematic puzzles through the geometry of uniformly accelerated observers.

Horizons and Causality

The Rindler horizon

In Rindler coordinates, the horizon is situated at the coordinate location \bar{x} = 0, which corresponds to the null emanating from the origin in the underlying Minkowski , defined by the surfaces x = |t|. This placement renders events in the region x < |t| causally disconnected from observers within the right Rindler wedge, as no timelike or lightlike paths can link them to points inside the wedge. For observers in the right wedge undergoing uniform proper acceleration, the Rindler horizon functions as a future horizon, preventing any signals or causal influences from the opposing left wedge from reaching them, regardless of the duration of their acceleration—even as their proper time approaches infinity. The worldlines of these Rindler observers asymptotically approach the horizon but remain confined to the wedge without crossing it. Geometrically, the horizon constitutes a null hypersurface, generated by a congruence of affinely parameterized that are tangent to the boundary of the Rindler wedge, thereby delineating the causal structure of the spacetime region accessible to the observers. The presence of this horizon in the classical setup is closely associated with the Unruh temperature T = a / (2\pi) (in natural units where \hbar = c = k_B = 1), at which accelerated observers perceive a thermal bath arising from the horizon's structure. In Penrose diagrams of , the Rindler horizon manifests as the null boundaries separating the right wedge from the left and the future and past wedges, visually emphasizing the causal isolation of the regions and the diamond-shaped conformal structure of the full spacetime.

Minkowski observers and horizon effects

From the perspective of inertial observers using Minkowski coordinates, the Rindler horizon corresponds precisely to the past and future light cones originating at the coordinate origin (t=0, x=0). These light cones define the boundary beyond which signals cannot reach the accelerating Rindler observers, who inhabit the right wedge (x > |t|) of Minkowski . The worldlines of Rindler observers, characterized by constant , trace out branches of hyperbolae that asymptotically approach this horizon as their increases, but they never cross it. In this inertial frame, the horizon is thus a purely geometric feature of flat spacetime, without any intrinsic or curvature. A key observational contrast arises in the propagation of signals across the horizon. For Rindler observers, signals originating from distant regions (large \bar{x}) within the accessible and received near the horizon (small \bar{x}) undergo an infinite gravitational-like in the limit as the receiver approaches the boundary, rendering such signals unobservable due to the extreme frequency shift. This effect stems from the position-dependent in the Rindler . However, an inertial Minkowski observer receiving the same signals experiences only a standard Doppler shift due to relative motion, without the extreme associated with the accelerated frame. This disparity highlights the observer-dependent nature of the effect, stemming from the motion rather than any global property of the . Mutual unobservability further underscores the of the horizon. Inertial observers located outside the Rindler wedge can detect events occurring within the wedge, including those involving the accelerating observers, since their light cones encompass the entire region. Conversely, Rindler observers are causally isolated from events beyond their horizon, unable to receive signals from the left wedge or other disconnected regions, even though these events are part of the same flat . This asymmetry arises because the Rindler coordinate patch excludes the left wedge, partitioning into causally disjoint sectors for the accelerated frame. Consider, for instance, a Minkowski observer at rest relative to the origin watching a Rindler observer undergoing hyperbolic motion. As the accelerating observer's worldline nears the horizon, relativistic causes the observed to slow dramatically, making the motion appear to "freeze" asymptotically close to the . Signals from the accelerating observer become increasingly infrequent and redshifted in the inertial frame, mimicking the behavior of objects approaching a , though here it is an artifact of alone. These horizon effects pose no causality paradoxes in flat spacetime, as all observer perspectives remain consistent within the global Minkowski structure. The apparent barriers are relative, dependent on the observer's motion: inertial frames see the full causal connectivity, while accelerated ones experience partitioned regions due to their trajectories approaching light-like asymptotes. This observer dependence exemplifies how reframes intuitive notions of and visibility without violating the light-speed limit or introducing true singularities.

Spacetime Geometry

Geodesics

In Rindler coordinates, the geodesic equations are derived from the of the ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + dz^2, where the non-vanishing symbols relevant to radial motion in the (\alpha, \rho) plane include \Gamma^\rho_{\alpha\alpha} = \rho and \Gamma^\alpha_{\alpha\rho} = \Gamma^\alpha_{\rho\alpha} = 1/\rho. These yield the coupled equations \ddot{\alpha} + \frac{2}{\rho} \dot{\alpha} \dot{\rho} = 0 and \ddot{\rho} + \rho \dot{\alpha}^2 = 0, where dots denote derivatives with respect to an affine parameter \lambda. Null geodesics, corresponding to light rays with ds^2 = 0, satisfy \dot{\rho}^2 = \rho^2 \dot{\alpha}^2 for radial paths. Combined with the geodesic equations, this implies \rho \dot{\alpha} = \pm k for some constant k, leading to solutions \rho(\alpha) = \rho_0 e^{\pm (\alpha - \alpha_0)} in the (\alpha, \rho) , though the exact form depends on the affine parameterization. These paths generate the Rindler horizon at \rho = 0, as the outgoing and ingoing geodesics asymptote to the of the Rindler , becoming incomplete as \alpha \to \pm \infty. Timelike geodesics, representing the worldlines of free-falling observers, obey the normalization -\rho^2 \dot{\alpha}^2 + \dot{\rho}^2 = -1. The conservation law from the Killing vector \partial/\partial\alpha gives \rho^2 \dot{\alpha} = k (constant), substituting to yield \dot{\rho}^2 = k^2/\rho^2 - 1. Integrating provides \rho(\tau) = \sqrt{ k^2 - (\tau - \tau_0)^2 }, where \tau is proper time and the turning point is at maximum \rho = |k|; these paths deviate from the constant-\rho hyperbolic worldlines of accelerated Rindler observers due to initial velocity components. Such geodesics are incomplete, terminating at the horizon after finite proper time \Delta\tau = 2|k| from past to future horizon. Spacelike geodesics, with normalization -\rho^2 \dot{\alpha}^2 + \dot{\rho}^2 = +1, follow analogous forms but describe maximal spatial separations, such as the shortest paths between simultaneous events for rigid rods in the Rindler frame. The same Killing symmetry conserves the \alpha- p_\alpha = -\rho^2 \dot{\alpha} = constant along all types, reflecting the translational invariance in \alpha.

Notions of distance

In Rindler coordinates, the spacetime metric in flat Minkowski space takes the form ds^2 = -\rho^2 \, d\alpha^2 + d\rho^2 + dy^2 + dz^2, where \rho > 0 is the radial coordinate, \alpha is the time-like coordinate, and y, z are Cartesian transverse coordinates. For measurements of spatial distance at constant \alpha, which defines hypersurfaces orthogonal to the worldlines of Rindler observers, the line element simplifies to dl^2 = d\rho^2 + dy^2 + dz^2. This yields a Euclidean geometry in the spatial sections, with the proper distance between two points at fixed y, z and \alpha given by integrating along the path, such as \Delta l = \int_{\rho_1}^{\rho_2} d\rho = \rho_2 - \rho_1 for radial separation. In cylindrical coordinates for the transverse plane, the spatial metric becomes dl^2 = d\rho^2 + \rho^2 d\phi^2 + dy^2 + dz^2, but the radial proper distance remains \rho_2 - \rho_1. The proper distance between Rindler observers at different \rho but same transverse position and constant \alpha is thus \Delta l = |\rho_2 - \rho_1|, measured along these orthogonal hypersurfaces of . This distance relates logarithmically to the effective , interpreted via the as the proper a(\rho) = 1/\rho (in units where c = 1), since \rho = 1/a. Consequently, \Delta l = |1/a_1 - 1/a_2|, or approximately \Delta a / a^2 for nearby observers with small acceleration difference \Delta a, highlighting how scales inversely with acceleration in the accelerated frame. An alternative operational measure is the , determined from the round-trip for a signal between observers. For an observer at \rho_1 with a = 1/\rho_1 sending a radial signal to \rho_2 > \rho_1 and back, the geodesics satisfy d\rho = \pm \rho \, d\alpha, leading to \Delta \alpha = 2 \ln(\rho_2 / \rho_1). The corresponding interval is \Delta \tau = \rho_1 \Delta \alpha = 2 \ln(\rho_2 / \rho_1) / a, so the is \Delta \tau / 2 = \ln(\rho_2 / \rho_1) / a. This differs markedly from the coordinate or \rho_2 - \rho_1, as the logarithmic form arises from the scaling of propagation in the . These notions reveal key challenges in the Rindler frame: there is no simultaneity across the entire , as the constant-\alpha slices cover only the Rindler and exclude regions behind the horizon. Distances appear to "stretch" near the horizon (\rho \to 0) due to the \rho factor in the metric, which causes extreme and , making nearby events at small \rho effectively remote in measurements. In comparison to Minkowski coordinates, where spatial distances are invariant and is global via Lorentz frames, Rindler distances are frame-dependent, reflecting the non-inertial nature of the accelerating observers; for instance, the same pair of events separated by fixed Minkowski interval yields \Delta l = \rho_2 - \rho_1 in Rindler but a Lorentz-contracted or dilated when viewed from an inertial frame. This dependence underscores how acceleration warps perceptions of length, analogous to gravitational effects in .

Fermat metric

In Rindler coordinates, the propagation of follows null geodesics of the underlying Minkowski spacetime, but as perceived by uniformly accelerating observers, these paths can be analyzed using an effective spatial known as the Fermat metric. This arises from projecting the null condition ds^2 = 0 onto constant-time spatial hypersurfaces in the Rindler frame. The standard Rindler is ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + dz^2, where \rho > 0 is the spatial coordinate to the acceleration direction, \alpha is the timelike coordinate (proportional to at fixed \rho), and y, z are transverse coordinates. Setting ds^2 = 0 yields \rho^2 d\alpha^2 = d\rho^2 + dy^2 + dz^2, so the infinitesimal interval for signals is d\alpha = \sqrt{(d\rho^2 + dy^2 + dz^2)} / \rho. Thus, the Fermat , which governs the paths of least \Delta \alpha, is dl_F^2 = (d\rho^2 + dy^2 + dz^2) / \rho^2. This Fermat is conformally flat, equivalent to the of the Poincaré half-plane (with negative K = -1), where light rays correspond to geodesics of this Riemannian structure. Fermat's principle states that light takes the path of stationary \int d\alpha = \int dl_F, analogous to the in a medium with position-dependent n = 1/\rho embedded in , where the physical spatial distance is dl^2 = d\rho^2 + dy^2 + dz^2. For transverse propagation (fixed \rho, varying y, z), the simplifies to dl_F^2 = (dy^2 + dz^2)/\rho^2, independent of \rho in form but scaled by the local value, leading to straight-line paths in the (y, z) at constant \rho. In polar coordinates (\rho, \phi) for motion involving the radial and azimuthal directions in the transverse , the becomes dl_F^2 = (d\rho^2 / \rho^2) + d\phi^2, so dl_F = \rho d\phi along angular directions at fixed \rho, and full paths are straight lines in the transformed coordinates (\ln \rho, \phi), appearing curved (as semicircles) in the physical (\rho, \phi) . This framework reveals optical effects for Rindler observers, such as aberration and gravitational lensing analogs when viewing distant . Due to the varying effective n = 1/\rho, from asymptotic directions bends toward higher \rho (lower regions), causing apparent distortion of stellar positions: appear shifted forward in the direction of , with increased blueshift and crowding near the forward horizon. This mimics in a stratified medium with a uniform gradient, where the decreases with "height" \rho, leading to upward deflection of rays relative to inertial observers. For example, a at in the transverse direction follows a hyperbolic geodesic in the Fermat , resulting in an observed that scales with the observer's a = 1/\rho. These effects are classical manifestations of the coordinate and have implications for signal propagation in accelerated frames, though they do not alter the underlying flat causality.

Symmetries and Structure

Isometries of the Rindler wedge

The isometries of the Rindler wedge are generated by Killing vector fields that preserve the Rindler metric ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + dz^2 (with \rho > 0) and map the wedge region x > |t| in Minkowski coordinates to itself. These include the time-translation Killing vector \partial_\alpha, which generates translations along the hyperbolic time coordinate and corresponds to a boost in the embedding Minkowski space; the spatial translation Killing vectors \partial_y and \partial_z, which preserve the transverse directions; and the boost-like Killing vector \alpha \partial_\rho + \rho \partial_\alpha, which generates hyperbolic rotations in the (\alpha, \rho) plane. In the full four-dimensional setting, an additional Killing vector arises from rotations in the transverse y-z plane, given explicitly by the infinitesimal generator \xi = y \partial_z - z \partial_y, which preserves the metric due to its independence from angular coordinates in cylindrical formulations. The Lie algebra of these Killing vectors closes under the Lie bracket, forming a with commutators such as [\partial_\alpha, \alpha \partial_\rho + \rho \partial_\alpha] = \partial_\rho, reflecting the boost-translation relations in the longitudinal sector, while the transverse vectors commute among themselves and with the longitudinal ones. The isometry group generated by these Killing vectors is a subgroup of the full Poincaré group of Minkowski space, specifically isomorphic to \mathrm{ISO}(2,1), the group of isometries of (2+1)-dimensional Minkowski spacetime, comprising three translations (one longitudinal and two transverse) and three Lorentz transformations (two boosts in the \alpha-\rho plane and one rotation in the y-z plane). This group structure ensures that all isometries preserve the horizon at \rho = 0, mapping the causal wedge to itself without mixing it with the complementary regions, in contrast to the unrestricted Poincaré transformations that would violate the wedge's boundaries. In representations of this isometry group, particularly in quantum field theory contexts, the Casimir invariants of the Lie algebra \mathfrak{iso}(2,1) play a key role: the quadratic Casimir C_1 = P_\alpha^2 + P_y^2 + P_z^2 corresponds to the squared mass in particle representations, while a linear Casimir derived from the Pauli-Lubanski pseudovector components governs helicity or spin labels, providing invariants under the restricted symmetries of the wedge. These invariants facilitate the classification of unitary irreducible representations relevant to fields propagating in the Rindler frame.

Relation to Lorentz group

The isometries of the Rindler wedge form a subgroup of the Poincaré group, which is the full isometry group of Minkowski spacetime. This subgroup preserves the wedge region W_R = \{ (t, x, y, z) \mid x > |t| \} and is generated by Lorentz boosts in the t-x plane, rotations in the y-z plane, along with translations in the perpendicular y and z directions. These generators ensure that the Minkowski metric, when expressed in Rindler coordinates, admits Killing vectors corresponding to these transformations. The structure of this subgroup can be understood as the stabilizer of the origin's light cone in the positive x-direction, restricting the full Lorentz group actions to those that leave the wedge invariant. The full group acting on the wedge aligns with the little group of the Lorentz group for null vectors defining the wedge boundaries; for a null vector in 3+1 dimensions, this little group is \mathrm{ISO}(2), comprising transverse rotations and translations augmented by the longitudinal boost. This embedding highlights how Rindler symmetries arise as a specialized Poincaré subgroup tailored to accelerated observers. Lorentz boosts along the x-direction realize the \alpha-translations—time translations in Rindler coordinates—through the nonlinear coordinate transformation relating Rindler and Minkowski frames. In this setup, a boost with rapidity parameter \phi shifts the Rindler time \alpha by \Delta \alpha = \phi / \kappa, where \kappa is the surface gravity parameter, linking global Lorentz transformations to local stationary symmetries. From the perspective of , Rindler coordinates diagonalize the boost generators of the , providing a natural basis for irreducible representations associated with accelerating trajectories. The boost operator, in the Rindler frame, acts multiplicatively on the spatial coordinate \xi, manifesting the hyperbolic motion and enabling explicit construction of unitary representations for accelerated systems. These symmetries underpin the invariance of physical laws within the , ensuring of quantities like defined relative to Rindler time \alpha. For instance, the Rindler , derived from the boost Killing , commutes with the field equations, yielding a conserved for observers at constant . This framework extends the inertial symmetries of the to noninertial settings while preserving .

Extensions and Applications

Generalization to curved spacetimes

In , Rindler-like coordinates generalize the flat description to curved geometries, particularly for regions probed by uniformly accelerating observers or near event horizons, where curvature introduces effects like tidal forces via that are absent in the Minkowski case. These coordinates facilitate the analysis of horizon structures and observer-dependent slicing in metrics such as Schwarzschild and de Sitter, maintaining the character of worldlines while incorporating gravitational influences. The flat Rindler wedge serves as the local approximation in the weak-field limit. A key analogue in the Schwarzschild geometry is provided by Painlevé-Gullstrand coordinates, which extend the Rindler framework by offering a regular across the event horizon, unlike the singular . In these coordinates, the takes the form ds^2 = -\left(1 - \frac{2M}{r}\right) dt^2 + 2\sqrt{\frac{2M}{r}}\, dt\, dr + dr^2 + r^2 d\Omega^2, where ingoing geodesics appear as straight radial lines dr/dt = -\sqrt{2M/r}, mirroring the causal structure of rays in Rindler coordinates. This formulation describes the from the perspective of freely falling observers, with the horizon at r = 2M behaving analogously to the Rindler horizon, and physical dynamics for accelerating observers near the horizon aligning locally with those in flat Rindler space due to the . In de Sitter , which models an expanding universe with positive , Rindler-like coordinates arise through hyperbolic slicing adapted for uniformly accelerating observers. These observers follow worldlines of constant , generalizing the hyperbolic trajectories of flat space Rindler motion, with the metric foliated by hyperbolic spatial sections that account for the global expansion. The geometry covers a wedge-like region bounded by a , where the accelerating observer's accessible exhibits a Rindler horizon structure, influenced by the de Sitter radius l = \sqrt{3/\Lambda}. This slicing highlights how acceleration interacts with cosmic expansion, differing from static de Sitter coordinates by emphasizing the observer's hyperbolic path. Near horizons, the Rindler approximation emerges in the near-horizon limit using Gaussian null coordinates (v, r, x^A), where the expands as ds^2 = - \kappa r\, dv^2 - 2\, dv\, dr + \Omega_{AB}(r, x) dx^A dx^B + \mathcal{O}(r), with \kappa the surface gravity playing the role of , and r = 0 locating the horizon. This form directly approximates the Rindler , capturing the universal near-horizon geometry for holes, where curvature effects are subleading close to the bifurcation surface. The approximation holds for the Schwarzschild case and extends to rotating or charged holes, providing a local flat-space-like description for horizon physics. For uniformly accelerating observers in general curved spacetimes, the definition of "uniform acceleration" generalizes Rindler's flat-space condition to trajectories satisfying \mathbf{a} \cdot \mathbf{u} = \alpha (constant proper acceleration \alpha), but geodesic deviation introduces tidal perturbations that stretch or compress nearby test particles, unlike the rigid congruence in . These tidal effects, quantified by the Riemann tensor components along the worldline, lead to non-zero relative accelerations between comoving observers, potentially causing trajectory incompleteness in strong fields. In the , for instance, radial acceleration amplifies tidal stretching along the direction of motion. An illustrative example is the Kruskal-Szekeres coordinates for the maximal extension of the Schwarzschild spacetime, given by ds^2 = \frac{32 M^3}{r} e^{-r/2M} (-dT^2 + dX^2) + r^2 d\Omega^2, with T and X related to null Eddington-Finkelstein coordinates via T - X = -e^{-(v/4M)} and T + X = e^{(u/4M)}, where r is implicitly defined by T^2 - X^2 = (r/2M - 1) e^{r/2M}. In the flat limit M \to 0, this reduces to Minkowski spacetime, and the region near the horizon (r \approx 2M) approximates the Rindler wedge, demonstrating how the curved horizon geometry emerges as a deformation of the flat Rindler structure.

Unruh effect and quantum applications

In quantum field theory formulated in Rindler coordinates, the Unruh effect arises as a consequence of the observer-dependent nature of the vacuum state. An observer with uniform proper acceleration a through the Minkowski vacuum perceives a thermal spectrum of particles, characterized by the Unruh temperature T = \frac{\hbar a}{2\pi k_B c}, where \hbar is the reduced Planck constant, k_B is Boltzmann's constant, and c is the speed of light. This effect was first derived by considering the response of idealized detectors to the quantum field, revealing that the Minkowski vacuum appears as a heat bath to the accelerating observer. The underlying mechanism involves Bogoliubov transformations that relate the Minkowski and Rindler mode decompositions of the quantum field. In the Rindler frame, the field operator \hat{\phi} is expanded in modes that are positive frequency with respect to Rindler time \eta, leading to creation and annihilation operators \hat{a}^R_\omega and \hat{a}^{\dagger R}_\omega for the right wedge. These Rindler operators mix Minkowski creation and annihilation operators via coefficients \alpha_{\omega\omega'} and \beta_{\omega\omega'}, such that \hat{a}^R_\omega = \int d\omega' \left( \alpha_{\omega\omega'} \hat{a}^M_{\omega'} + \beta_{\omega\omega'} {\hat{a}^M_{\omega'}}^\dagger \right), with |\beta_{\omega\omega'}|^2 determining the particle density. Equivalently, in canonical quantization, the Rindler annihilation operator takes the form \hat{a}_R = \int (u_R \hat{\phi} + i v_R \hat{\pi}), where \hat{\pi} is the conjugate momentum, u_R and v_R are mode functions satisfying the Klein-Gordon equation in Rindler coordinates, and the mixing of \hat{\phi} and \hat{\pi} reflects the non-trivial vacuum structure. The Rindler vacuum |0_R\rangle, defined by \hat{a}^R_\omega |0_R\rangle = 0 for all frequencies \omega in the right wedge, differs fundamentally from the Minkowski vacuum |0_M\rangle. It is a highly entangled state in the Minkowski basis, with correlations between modes in the right and left Rindler wedges separated by the horizon. This entanglement manifests as for the Rindler observer, where the expectation value of the energy-momentum tensor yields a . The Rindler horizon acts as the boundary delineating this vacuum entanglement. The serves as a flat-spacetime analog to near horizons, where the a corresponds to the surface gravity \kappa, yielding an identical formula T = \frac{\hbar \kappa}{2\pi k_B c}. This analogy underscores the role of horizons in inducing particle creation from the in both scenarios. In applications, the Unruh-DeWitt detector—a model of a two-level quantum system linearly coupled to —quantifies the excitation rate, which for constant matches the thermal response at the Unruh , enabling operational tests of the effect. Quantum inequalities impose bounds on the integrated densities experienced by accelerated observers, preventing violations of energy conditions while accommodating the Unruh flux. Additionally, entanglement harvesting protocols use pairs of accelerated Unruh-DeWitt detectors to extract bipartite entanglement from the Minkowski , with negativity measures peaking when detectors straddle the Rindler horizon.

Historical Development

Overview

The concept of hyperbolic motion in , central to what would become Rindler coordinates, emerged shortly after the formulation of the theory. In 1909, introduced the idea of rigid body motion under constant , coining the term "hyperbolic motion" to describe the trajectory, which traces a in diagrams due to the relativistic invariance of proper acceleration. This built on earlier insights from Hermann Minkowski's 1908 formalism, where Lorentz transformations were interpreted as hyperbolic rotations preserving the interval. Born's work addressed the challenges of defining rigidity in accelerated frames, motivated by the need to extend Newtonian concepts to relativistic dynamics without violating causality. Rindler's 1960 textbook offered an early detailed treatment of hyperbolic motion for accelerating observers, paving the way for his formal coordinate system. Preceding Rindler's formalization, significant contributions appeared in the context of accelerated reference frames and gravitational analogies. In 1943, Christian Møller developed coordinates based on for observers in uniformly accelerated motion, analyzing homogeneous gravitational fields as a local simulation of acceleration via the . Møller's approach, detailed in his study of field equations in such frames, provided a precursor to the coordinate transformation now associated with Rindler, emphasizing the synchronization of clocks along worldlines of constant . These efforts were driven by practical motivations, such as modeling propulsion and resolving apparent paradoxes in the , where acceleration mimics gravity but introduces horizons beyond which signals cannot reach the observer. Wolfgang Rindler formalized the framework in his 1966 paper "Kruskal Space and the Uniformly Accelerated Frame," introducing coordinates tailored to uniformly accelerating observers to highlight formal similarities with Kruskal diagrams in and issues related to horizons in accelerated frames. Motivated by analogies between acceleration and , Rindler's system described the wedge accessible to such observers, highlighting event horizons analogous to those in cosmology. This work unified scattered results on horizons and accelerated motion, providing a natural coordinate chart for the Rindler wedge in Minkowski . Following Rindler's contribution, the coordinates saw widespread adoption in literature during the and , appearing in seminal texts that integrated relativistic acceleration with gravitational theory. For instance, they featured prominently in Misner, Thorne, and Wheeler's 1973 "Gravitation," where they illustrated local inertial frames and applications. Interest surged in the with quantum connections, notably William Unruh's 1976 demonstration that accelerating observers perceive a thermal bath in the vacuum, linking Rindler coordinates to in curved . These developments underscored the coordinates' role in simulating uniform fields and probing foundational issues.

Evolution of key formulas

The mathematical formulation of coordinates describing uniformly accelerated observers in began with early attempts to model hyperbolic motion in the context of the and gravitational analogies. These initial efforts focused on two-dimensional transformations, often overlooking transverse dimensions, and were motivated by resolving clock discrepancies in accelerated frames. Over time, the coordinates evolved from ad hoc expressions for specific trajectories to a standardized system covering the Rindler wedge, incorporating variable and full four-dimensional structure, with quantum field applications emerging in the . Key milestones in this development are summarized in the following table, highlighting the progression of transformation forms from inertial (Minkowski) coordinates (t, x) to accelerated coordinates (e.g., (\tau, x') or (\alpha, \rho)):
YearAuthorTransformation Form
1943Møllert = \left( \frac{c}{g} + \frac{X}{c} \right) \sinh\left( \frac{g T}{c} \right), x = \left( \frac{c^2}{g} + X \right) \cosh\left( \frac{g T}{c} \right) - \frac{c^2}{g} (2D form for constant acceleration g, relating inertial t, x to accelerated T, X; ignores y, z).
1966RindlerT = \rho \sinh \alpha, X = \rho \cosh \alpha (standard form introducing spatial coordinate \rho > 0 for variable proper acceleration a = 1/\rho, with angular-like \alpha; metric ds^2 = -\rho^2 d\alpha^2 + d\rho^2 + dy^2 + dz^2; full 4D, focused on right wedge, with extension to left wedge symmetry: analogous transformations for X < 0, T = -\rho \sinh \alpha, X = -\rho \cosh \alpha, formalizing dual wedges and full causal structure).
1973FullingMode expansions for quantum fields: scalar field \phi = \int_0^\infty d\omega \left[ e^{-i \omega \ln(\rho)} (a_\omega^R u_\omega^R + a_\omega^L u_\omega^L) + \text{h.c.} \right] (Rindler modes u_\omega^{R/L} in right/left wedges, highlighting vacuum non-uniqueness and thermal spectrum).
This evolution marked a shift from trajectory-specific, two-dimensional mappings in Møller's work—designed to mimic homogeneous gravitational fields and address the clock paradox—to Rindler's 1966 generalization, which introduced the \rho scaling to accommodate observers at varying distances from the origin, thus allowing to depend on position (a = 1/\rho). Early formulations, such as Møller's, were limited to the plane of motion and did not explicitly include transverse coordinates y and z, treating them implicitly as unchanged; modern treatments, starting with Rindler's extensions, incorporate the full for completeness. By the , Rindler incorporated the left to describe observers accelerating in the opposite direction, completing the causal partitioning of into four wedges and emphasizing event horizons. The 1970s brought recognition of conformal invariance in the Rindler metric, facilitating applications, as seen in Fulling's mode expansions that revealed observer-dependent vacua. Standardization in texts, such as Hawking and (1973), further solidified the notation, adopting Rindler's form for wedge metrics and influencing subsequent curved- generalizations.

References

  1. [1]
    [PDF] on the uniformly accelerated observer - MIT
    Figure 3: An illustration of Rindler coordinates. The red curve is the worldline of an accelerating observer who starts at x = c2/g and experiences constant ...
  2. [2]
    Kruskal Space and the Uniformly Accelerated Frame - AIP Publishing
    Search Site. Citation. W. Rindler; Kruskal Space and the Uniformly Accelerated Frame. ... This content is only available via PDF. Open the PDF for in another ...
  3. [3]
    The Unruh effect and its applications | Rev. Mod. Phys.
    Jul 1, 2008 · The Unruh effect is defined here as the fact that the usual vacuum state for QFT in Minkowski spacetime restricted to the right Rindler wedge is ...Missing: primary | Show results with:primary
  4. [4]
    [PDF] Chapter 24: From Special to General Relativity [version 1224.1.K]
    Specifically, we introduce so-called Rindler coordinates 2 t′ = x. 0. , x′ = x. 1. + 1/κ , y′ = x. 2. , z′ = x. 3 . (24.71). 2Named for Wolfgang Rindler. Page ...
  5. [5]
  6. [6]
    [PDF] On the spectral shift and the time delay of light in a Rindler ... - arXiv
    Jul 2, 2006 · A set of coordinates with the desired properties mentioned above is provided by the so-called Rindler coordinates. We now proceed to define ...
  7. [7]
    [PDF] Accelerated Observers - UCSB Physics
    Figure 1: An accelerated observer out in empty space could come to the correct physical conclusions by assuming that he was not accelerating, but rather in a.Missing: Wolfgang | Show results with:Wolfgang
  8. [8]
    [PDF] Static Observers in Curved Spaces and Non-inertial Frames ... - arXiv
    Apr 22, 2010 · Therefore the Rindler congruence and the static Schwarzschild observers have the same acceleration field a(ρ). This equivalence reinforce the ...
  9. [9]
    Space and observers in cosmology
    4 The Rindler observer. The Rindler observer in Minkowski spacetime has constant acceleration a. Its velocity is defined by. $\displaystyle {u}^{0}=\cosh (a ...
  10. [10]
  11. [11]
    [PDF] Kinematic quantities for a spherical distribution of uniformly ... - arXiv
    Oct 17, 2014 · The generalized Rindler reference frame (or the spherical Rindler coordi- ... we obtain aα = (0, 1/ρ, 0, 0) . The invariant acceleration ...
  12. [12]
    [1411.7019] Inertial non-vacuum states viewed from the Rindler frame
    Nov 25, 2014 · Abstract page for arXiv paper 1411.7019: Inertial non-vacuum states viewed from the Rindler frame. ... time translation invariance and discuss ...
  13. [13]
    [PDF] arXiv:gr-qc/0202085v2 9 Jan 2003
    transport and Fermi-Walker transport frequencies for stationary circular orbits. ... Rindler metric built on an observer family each member of which ...
  14. [14]
  15. [15]
    [1612.03158] Remarks on Rindler Quantization - arXiv
    Dec 9, 2016 · We review the quantization of scalar and gauge fields using Rindler coordinates with an emphasis on the physics of the Rindler horizon.
  16. [16]
  17. [17]
    [PDF] Rindler space, Unruh effect and Hawking temperature - M. Socolovsky
    Note : A static observer at r must be accelerated; on the contrary its motion would be geodesic i.e. in free fall. Let us compute κ. The 4-acceleration of a ...
  18. [18]
    Unruh effect - Scholarpedia
    Oct 10, 2014 · The mathematical relationship between Rindler and Minkowski coordinates in flat space is practically identical to that between Schwarzschild ...Missing: primary | Show results with:primary
  19. [19]
    [PDF] What a Rindler Observer Sees in a Minkowski Vacuum
    We label the left- and right-hand wedges by L and R respectively. The null rays act as event horizons for Rindler observers: an observer in R cannot see events ...
  20. [20]
  21. [21]
    [PDF] arXiv:1911.03950v1 [gr-qc] 10 Nov 2019
    Nov 10, 2019 · Such a localized observer moving in a constant velocity in Minkowski space can operationally define a reference frame, called “radar coordinates ...
  22. [22]
    [PDF] Entropy of Rindler space
    Oct 29, 2020 · First, the solutions for a rotating black hole was found by Kerr [5] and an extension to a charged rotating black hole was found by Kerr and.
  23. [23]
    [PDF] Revisiting quantum field theory in Rindler spacetime with ... - arXiv
    Rindler coordinates map Minkowski spacetime onto regions with horizons, effectively dividing accelerated observers into causally disconnected sec- tors.
  24. [24]
    [PDF] arXiv:hep-th/0310099v2 19 Apr 2005
    ... ISO (2, 1). Let us start by introducing Minkowski coordinates X0, X1, X2 and ... Rindler wedge. This. 20. Page 21. time, though, we do not have a natural ...
  25. [25]
    [PDF] Some aspects of QFT in non-inertial frames. - arXiv
    Apr 24, 2025 · After a brief recap of Killing vectors and their horizons, we study in some detail a few explicit examples, namely the Rindler frame, the ...
  26. [26]
  27. [27]
    [PDF] The Rich Structure of Minkowski Space - arXiv
    Feb 29, 2008 · The Poincaré group in n ≥ 2 dimensions, which is the same as the inhomogeneous Lorentz group in n ≥ 2 dimensions and therefore will be denoted ...
  28. [28]
    Quasilocal first law of black hole dynamics from local Lorentz ...
    Jul 5, 2018 · As is well known, these must belong to , the little group of the Lorentz group [11] that keeps a null vector invariant. Further, using the ...
  29. [29]
    [PDF] Coherent States of Accelerated Relativistic Quantum Particles ...
    Aug 10, 2011 · Appendix A reminds. Rindler coordinates and Bogoliubov transformations in the standard derivation of Unruh ... and D (i.e., a Poincaré subgroup P ...
  30. [30]
  31. [31]
  32. [32]
  33. [33]
    Scalar production in Schwarzschild and Rindler metrics - IOPscience
    The procedure used recently by Hawking (1975) to demonstrate the creation of massless particles by black holes is applied to the Rindler coordinate system in ...
  34. [34]
    [quant-ph/0212044] Entanglement from the vacuum - arXiv
    Dec 8, 2002 · We explore the entanglement of the vacuum of a relativistic field by letting a pair of causally disconnected probes interact with the field.
  35. [35]
    [PDF] arXiv:2307.00023v1 [gr-qc] 28 Jun 2023
    Jun 28, 2023 · The free fall rocket (proper) time is simply γ0 times the ground (proper) time, as might have been guessed, and a little larger than the ...
  36. [36]