Rindler space
   HOME

TheInfoList



OR:

In
relativistic physics In physics, relativistic mechanics refers to mechanics compatible with special relativity (SR) and general relativity (GR). It provides a non-quantum mechanical description of a system of particles, or of a fluid, in cases where the velocities of ...
, the coordinates of a ''hyperbolically accelerated reference frame'' constitute an important and useful
coordinate chart In topology, a branch of mathematics, a topological manifold is a topological space that locally resembles real ''n''-dimensional Euclidean space. Topological manifolds are an important class of topological spaces, with applications throughout mathe ...
representing part of flat
Minkowski spacetime In mathematical physics, Minkowski space (or Minkowski spacetime) () is a combination of Three-dimensional space, three-dimensional Euclidean space and time into a four-dimensional manifold where the spacetime interval between any two Event (rel ...
. In
special relativity In physics, the special theory of relativity, or special relativity for short, is a scientific theory regarding the relationship between space and time. In Albert Einstein's original treatment, the theory is based on two postulates: # The laws ...
, a uniformly accelerating particle undergoes
hyperbolic motion In geometry, hyperbolic motions are isometric automorphisms of a hyperbolic space. Under composition of mappings, the hyperbolic motions form a continuous group. This group is said to characterize the hyperbolic space. Such an approach to geom ...
, for which a uniformly
accelerating In mechanics, acceleration is the rate of change of the velocity of an object with respect to time. Accelerations are vector quantities (in that they have magnitude and direction). The orientation of an object's acceleration is given by th ...
frame of reference in which it is at rest can be chosen as its proper reference frame. The phenomena in this hyperbolically accelerated frame can be compared to effects arising in a homogeneous gravitational field. For general overview of accelerations in flat spacetime, see
Acceleration (special relativity) Accelerations in special relativity (SR) follow, as in Newtonian Mechanics, by differentiation of velocity with respect to time. Because of the Lorentz transformation and time dilation, the concepts of time and distance become more complex, which ...
and
Proper reference frame (flat spacetime) A proper reference frame in the theory of relativity is a particular form of accelerated reference frame, that is, a reference frame in which an accelerated observer can be considered as being at rest. It can describe phenomena in curved spacetime, ...
. In this article, the
speed of light The speed of light in vacuum, commonly denoted , is a universal physical constant that is important in many areas of physics. The speed of light is exactly equal to ). According to the special theory of relativity, is the upper limit ...
is defined by , the inertial coordinates are , and the hyperbolic coordinates are . These hyperbolic coordinates can be separated into two main variants depending on the accelerated observer's position: If the observer is located at time at position (with as the constant
proper acceleration In relativity theory, proper acceleration is the physical acceleration (i.e., measurable acceleration as by an accelerometer) experienced by an object. It is thus acceleration relative to a free-fall, or inertial, observer who is momentarily at ...
measured by a comoving
accelerometer An accelerometer is a tool that measures proper acceleration. Proper acceleration is the acceleration (the rate of change of velocity) of a body in its own instantaneous rest frame; this is different from coordinate acceleration, which is acc ...
), then the hyperbolic coordinates are often called Rindler coordinates with the corresponding ''Rindler metric''. If the observer is located at time at position , then the hyperbolic coordinates are sometimes called ''Møller coordinates'' or ''Kottler–Møller coordinates'' with the corresponding ''Kottler–Møller metric''. An alternative chart often related to observers in hyperbolic motion is obtained using
Radar Radar is a detection system that uses radio waves to determine the distance ('' ranging''), angle, and radial velocity of objects relative to the site. It can be used to detect aircraft, ships, spacecraft, guided missiles, motor vehicles, we ...
coordinates which are sometimes called ''Lass coordinates''. Both the Kottler–Møller coordinates as well as Lass coordinates are denoted as Rindler coordinates as well. Regarding the history, such coordinates were introduced soon after the advent of special relativity, when they were studied (fully or partially) alongside the concept of hyperbolic motion: In relation to flat Minkowski spacetime by
Albert Einstein Albert Einstein ( ; ; 14 March 1879 – 18 April 1955) was a German-born theoretical physicist, widely acknowledged to be one of the greatest and most influential physicists of all time. Einstein is best known for developing the theory ...
(1907, 1912), Max Born (1909), Arnold Sommerfeld (1910), Max von Laue (1911), Hendrik Lorentz (1913),
Friedrich Kottler Friedrich Kottler (December 10, 1886 – May 11, 1965) was an Austrian theoretical physicist. He was a Privatdozent before he got a professorship in 1923 at the University of Vienna. Life In 1938, after the Anschluss, he lost his profess ...
(1914),
Wolfgang Pauli Wolfgang Ernst Pauli (; ; 25 April 1900 – 15 December 1958) was an Austrian theoretical physicist and one of the pioneers of quantum physics. In 1945, after having been nominated by Albert Einstein, Pauli received the Nobel Prize in Physics ...
(1921), Karl Bollert (1922),
Stjepan Mohorovičić Stjepan Mohorovičić (August 20, 1890 – February 13, 1980) was a Croatian physicist, geophysicist and meteorologist. Biography Mohorovičić was born in the town of Bakar. His father is the world-famous geophysicist Andrija Mohorovičić. He ...
(1922),
Georges Lemaître Georges Henri Joseph Édouard Lemaître ( ; ; 17 July 1894 – 20 June 1966) was a Belgian Catholic priest, theoretical physicist, mathematician, astronomer, and professor of physics at the Catholic University of Louvain. He was the first to t ...
(1924), Einstein &
Nathan Rosen Nathan Rosen (Hebrew: נתן רוזן; March 22, 1909 – December 18, 1995) was an American-Israeli physicist noted for his study on the structure of the hydrogen atom and his work with Albert Einstein and Boris Podolsky on entangled wave functio ...
(1935), Christian Møller (1943, 1952),
Fritz Rohrlich Fritz Rohrlich (May 12, 1921 – November 14, 2018) was an American theoretical physicist and educator who published in the fields of quantum electrodynamics, classical electrodynamics of charged particles, and the philosophy of science. Life and ...
(1963), Harry Lass (1963), and in relation to both flat and
curved spacetime Curved space often refers to a spatial geometry which is not "flat", where a flat space is described by Euclidean geometry. Curved spaces can generally be described by Riemannian geometry though some simple cases can be described in other ways. Cu ...
of
general relativity General relativity, also known as the general theory of relativity and Einstein's theory of gravity, is the geometric theory of gravitation published by Albert Einstein in 1915 and is the current description of gravitation in modern physics ...
by
Wolfgang Rindler Wolfgang Rindler (18 May 1924 – 8 February 2019) was a physicist working in the field of general relativity where he is known for introducing the term "event horizon", Rindler coordinates, and (in collaboration with Roger Penrose) for the use of ...
(1960, 1966). For details and sources, see '.


Characteristics of the Rindler frame

The
worldline The world line (or worldline) of an object is the path that an object traces in 4-dimensional spacetime. It is an important concept in modern physics, and particularly theoretical physics. The concept of a "world line" is distinguished from con ...
of a body in
hyperbolic motion In geometry, hyperbolic motions are isometric automorphisms of a hyperbolic space. Under composition of mappings, the hyperbolic motions form a continuous group. This group is said to characterize the hyperbolic space. Such an approach to geom ...
having constant proper acceleration \alpha in the X-direction as a function of
proper time In relativity, proper time (from Latin, meaning ''own time'') along a timelike world line is defined as the time as measured by a clock following that line. It is thus independent of coordinates, and is a Lorentz scalar. The proper time interval ...
\tau and rapidity \alpha\tau can be given by :T=x\sinh(\alpha\tau),\quad X=x\cosh(\alpha\tau) where x=1/\alpha is constant and \alpha\tau is variable, with the worldline resembling the hyperbola X^-T^=x^2. Sommerfeld showed that the equations can be reinterpreted by defining x as variable and \alpha\tau as constant, so that it represents the simultaneous "rest shape" of a body in hyperbolic motion measured by a comoving observer. By using the proper time of the observer as the time of the entire hyperbolically accelerated frame by setting \tau=t, the transformation formulas between the inertial coordinates and the hyperbolic coordinates are consequently: with the inverse :t=\frac\operatorname\left(\frac\right),\quad x=\sqrt,\quad y=Y,\quad z=Z Differentiated and inserted into the Minkowski metric ds^=-dT^+dX^+dY^+dZ^, the
metric Metric or metrical may refer to: * Metric system, an internationally adopted decimal system of measurement * An adjective indicating relation to measurement in general, or a noun describing a specific type of measurement Mathematics In mathem ...
in the hyperbolically accelerated frame follows These transformations define the ''Rindler observer'' as an observer that is "at rest" in Rindler coordinates, i.e., maintaining constant ''x'', ''y'', ''z'' and only varying ''t'' as time passes. The coordinates are valid in the region 0 < X < \infty,\; -X < T < X, which is often called the ''Rindler wedge'', if \alpha represents the proper acceleration (along the hyperbola x=1 / \alpha) of the Rindler observer whose proper time is defined to be equal to Rindler coordinate time. To maintain this world line, the observer must accelerate with a constant proper acceleration, with Rindler observers closer to x=0 (the Rindler horizon) having greater proper acceleration. All the Rindler observers are instantaneously at rest at time T=0 in the inertial frame, and at this time a Rindler observer with proper acceleration \alpha_ will be at position X=1/\alpha_ (really X=c^/\alpha_, but we assume units where c=1), which is also that observer's constant distance from the Rindler horizon in Rindler coordinates. If all Rindler observers set their clocks to zero at T=0, then when defining a Rindler coordinate system we have a choice of which Rindler observer's
proper time In relativity, proper time (from Latin, meaning ''own time'') along a timelike world line is defined as the time as measured by a clock following that line. It is thus independent of coordinates, and is a Lorentz scalar. The proper time interval ...
will be equal to the coordinate time t in Rindler coordinates, and this observer's proper acceleration defines the value of \alpha above (for other Rindler observers at different distances from the Rindler horizon, the coordinate time will equal some constant multiple of their own proper time). It is a common convention to define the Rindler coordinate system so that the Rindler observer whose proper time matches coordinate time is the one who has proper acceleration \alpha=1, so that \alpha can be eliminated from the equations. The above equation has been simplified for c=1. The unsimplified equation is more convenient for finding the Rindler Horizon distance, given an acceleration \alpha. :\begint & =\frac\operatorname\left(\frac\right)\;\overset\;\frac\\ X & \approx\frac\;\overset\;\frac \end The remainder of the article will follow the convention of setting both \alpha=1 and c=1, so units for X and x will be 1 unit =c^/\alpha=1. Be mindful that setting \alpha=1 light-second/second2 is very different from setting \alpha=1 light-year/year2. Even if we pick units where c=1, the magnitude of the proper acceleration \alpha will depend on our choice of units: for example, if we use units of light-years for distance, (X or x) and years for time, (T or t), this would mean \alpha=1 light year/year2, equal to about 9.5 meters/second2, while if we use units of light-seconds for distance, (X or x), and seconds for time, (T or t), this would mean \alpha=1 light-second/second2, or 299 792 458 meters/second2).


Variants of transformation formulas

A more general derivation of the transformation formulas is given, when the corresponding Fermi–Walker tetrad is formulated from which the Fermi coordinates or Proper coordinates can be derived. Depending on the choice of origin of these coordinates, one can derive the metric, the time dilation between the time at the origin dt_ and dt at point x, and the coordinate light speed , dx, /, dt, (this
variable speed of light A variable speed of light (VSL) is a feature of a family of hypotheses stating that the speed of light may in some way not be constant, for example, that it varies in space or time, or depending on frequency. Accepted classical theories of physi ...
does not contradict special relativity, because it is only an artifact of the accelerated coordinates employed, while in inertial coordinates it remains constant). Instead of Fermi coordinates, also Radar coordinates can be used, which are obtained by determining the distance using light signals (see section Notions of distance), by which metric, time dilation and speed of light do not depend on the coordinates anymore – in particular, the coordinate speed of light remains identical with the speed of light (c=1) in inertial frames:


The Rindler observers

In the new chart () with c=1 and \alpha=1, it is natural to take the coframe field : d\sigma^0 = x \, dt,\;\; d\sigma^1 = dx,\;\; d\sigma^2 = dy,\;\; d\sigma^3 = dz which has the dual frame field : \vec_0 = \frac\partial_t,\;\; \vec_1 = \partial_x,\;\; \vec_2 = \partial_y,\;\; \vec_3 = \partial_z This defines a ''local Lorentz frame'' in the
tangent space In mathematics, the tangent space of a manifold generalizes to higher dimensions the notion of '' tangent planes'' to surfaces in three dimensions and ''tangent lines'' to curves in two dimensions. In the context of physics the tangent space to a ...
at each
event Event may refer to: Gatherings of people * Ceremony, an event of ritual significance, performed on a special occasion * Convention (meeting), a gathering of individuals engaged in some common interest * Event management, the organization of e ...
(in the region covered by our Rindler chart, namely the Rindler wedge). The
integral curve In mathematics, an integral curve is a parametric curve that represents a specific solution to an ordinary differential equation or system of equations. Name Integral curves are known by various other names, depending on the nature and interpret ...
s of the
timelike In physics, spacetime is a mathematical model that combines the three dimensions of space and one dimension of time into a single four-dimensional manifold. Spacetime diagrams can be used to visualize relativistic effects, such as why differen ...
unit vector field \vec_0 give a timelike congruence, consisting of the world lines of a family of observers called the ''Rindler observers''. In the Rindler chart, these world lines appear as the vertical coordinate lines x = x_0,\; y = y_0,\; z = z_0. Using the coordinate transformation above, we find that these correspond to hyperbolic arcs in the original Cartesian chart. As with any timelike congruence in any Lorentzian manifold, this congruence has a ''kinematic decomposition'' (see
Raychaudhuri equation In general relativity, the Raychaudhuri equation, or Landau–Raychaudhuri equation, is a fundamental result describing the motion of nearby bits of matter. The equation is important as a fundamental lemma for the Penrose–Hawking singularity the ...
). In this case, the ''expansion'' and ''vorticity'' of the congruence of Rindler observers ''vanish''. The vanishing of the expansion tensor implies that ''each of our observers maintains constant distance to his neighbors''. The vanishing of the vorticity tensor implies that the world lines of our observers are not twisting about each other; this is a kind of local absence of "swirling". The acceleration vector of each observer is given by the covariant derivative :\nabla_ \vec_0 = \frac\vec_1 That is, each Rindler observer is accelerating in the \partial_x direction. Individually speaking, each observer is in fact accelerating with ''constant magnitude'' in this direction, so their world lines are the Lorentzian analogs of circles, which are the curves of constant path curvature in the Euclidean geometry. Because the Rindler observers are ''vorticity-free'', they are also ''hypersurface orthogonal''. The orthogonal spatial hyperslices are t = t_0; these appear as horizontal half-planes in the Rindler chart and as half-planes through T = X = 0 in the Cartesian chart (see the figure above). Setting dt = 0 in the line element, we see that these have the ordinary Euclidean geometry, d\sigma^2 = dx^2 + dy^2 + dz^2,\; \forall x > 0, \forall y, z. Thus, the spatial coordinates in the Rindler chart have a very simple interpretation consistent with the claim that the Rindler observers are mutually stationary. We will return to this rigidity property of the Rindler observers a bit later in this article.


A "paradoxical" property

Note that Rindler observers with smaller constant x coordinate are accelerating ''harder'' to keep up. This may seem surprising because in Newtonian physics, observers who maintain constant relative distance must share the ''same'' acceleration. But in relativistic physics, we see that the trailing endpoint of a rod which is accelerated by some external force (parallel to its symmetry axis) must accelerate a bit harder than the leading endpoint, or else it must ultimately break. This is a manifestation of Lorentz contraction. As the rod accelerates, its velocity increases and its length decreases. Since it is getting shorter, the back end must accelerate harder than the front. Another way to look at it is: the back end must achieve the same change in velocity in a shorter period of time. This leads to a differential equation showing that, at some distance, the acceleration of the trailing end diverges, resulting in the Rindler horizon. This phenomenon is the basis of a well known "paradox",
Bell's spaceship paradox Bell's spaceship paradox is a thought experiment in special relativity. It was designed by E. Dewan and M. Beran in 1959 and became more widely known when J. S. Bell included a modified version.J. S. Bell: ''How to teach special relativity'', Prog ...
. However, it is a simple consequence of relativistic kinematics. One way to see this is to observe that the magnitude of the acceleration vector is just the path curvature of the corresponding world line. But ''the world lines of our Rindler observers are the analogs of a family of concentric circles'' in the Euclidean plane, so we are simply dealing with the Lorentzian analog of a fact familiar to speed skaters: in a family of concentric circles, ''inner circles must bend faster (per unit arc length) than the outer ones''.


Minkowski observers

It is worthwhile to also introduce an alternative frame, given in the Minkowski chart by the natural choice :\vec_0 = \partial_T, \; \vec_1 = \partial_X, \; \vec_2 = \partial_Y, \; \vec_3 = \partial_Z Transforming these vector fields using the coordinate transformation given above, we find that in the Rindler chart (in the Rinder wedge) this frame becomes :\begin \vec_0 &= \frac\cosh(t) \, \partial_t - \sinh(t) \, \partial_x\\ \vec_1 &= -\frac\sinh(t) \, \partial_t + \cosh(t) \, \partial_x\\ \vec_2 &= \partial_y, \; \vec_3 = \partial_z \end Computing the kinematic decomposition of the timelike congruence defined by the timelike unit vector field \vec_0 , we find that the expansion and vorticity again vanishes, and in addition the acceleration vector vanishes, \nabla_ \vec_0 = 0. In other words, this is a ''geodesic congruence''; the corresponding observers are in a state of ''inertial motion''. In the original Cartesian chart, these observers, whom we will call ''Minkowski observers'', are at rest. In the Rindler chart, the world lines of the Minkowski observers appear as hyperbolic secant curves asymptotic to the coordinate plane x = 0. Specifically, in Rindler coordinates, the world line of the Minkowski observer passing through the event t = t_0,\; x = x_0,\; y = y_0,\; z = z_0 is :\begin t &= \operatorname\left(\frac\right),\; -x_0 < s < x_0\\ x &= \sqrt,\; -x_0 < s < x_0\\ y &= y_0\\ z &= z_0 \end where s is the proper time of this Minkowski observer. Note that only a small portion of his history is covered by the Rindler chart. This shows explicitly why the Rindler chart is ''not'' geodesically complete; timelike geodesics run outside the region covered by the chart in finite proper time. Of course, we already knew that the Rindler chart cannot be geodesically complete, because it covers only a portion of the original Cartesian chart, which ''is'' a geodesically complete chart. In the case depicted in the figure, x_0 = 1 and we have drawn (correctly scaled and boosted) the light cones at s \in \left\.


The Rindler horizon

The Rindler coordinate chart has a ''coordinate singularity'' at ''x'' = 0, where the metric tensor (expressed in the Rindler coordinates) has vanishing
determinant In mathematics, the determinant is a scalar value that is a function of the entries of a square matrix. It characterizes some properties of the matrix and the linear map represented by the matrix. In particular, the determinant is nonzero if a ...
. This happens because as ''x'' → 0 the acceleration of the Rindler observers diverges. As we can see from the figure illustrating the Rindler wedge, the locus ''x'' = 0 in the Rindler chart corresponds to the locus ''T''2 = ''X''2, ''X'' > 0 in the Cartesian chart, which consists of two null half-planes, each ruled by a null geodesic congruence. For the moment, we simply consider the Rindler horizon as the boundary of the Rindler coordinates. If we consider the set of accelerating observers who have a constant position in Rindler coordinates, none of them can ever receive light signals from events with ''T'' ≥ ''X'' (on the diagram, these would be events on or to the left of the line ''T'' = ''X'' which the upper red horizon lies along; these observers could however receive signals from events with ''T'' ≥ ''X'' if they stopped their acceleration and crossed this line themselves) nor could they have ever sent signals to events with ''T'' ≤ −''X'' (events on or to the left of the line ''T'' = −''X'' which the lower red horizon lies along; those events lie outside all future light cones of their past world line). Also, if we consider members of this set of accelerating observers closer and closer to the horizon, in the limit as the distance to the horizon approaches zero, the constant proper acceleration experienced by an observer at this distance (which would also be the G-force experienced by such an observer) would approach infinity. Both of these facts would also be true if we were considering a set of observers hovering outside the
event horizon In astrophysics, an event horizon is a boundary beyond which events cannot affect an observer. Wolfgang Rindler coined the term in the 1950s. In 1784, John Michell proposed that gravity can be strong enough in the vicinity of massive compact ob ...
of a black hole, each observer hovering at a constant radius in
Schwarzschild coordinates In the theory of Lorentzian manifolds, spherically symmetric spacetimes admit a family of ''nested round spheres''. In such a spacetime, a particularly important kind of coordinate chart is the Schwarzschild chart, a kind of polar spherical coord ...
. In fact, in the close neighborhood of a black hole, the geometry close to the event horizon can be described in Rindler coordinates. Hawking radiation in the case of an accelerating frame is referred to as Unruh radiation. The connection is the equivalence of acceleration with gravitation.


Geodesics

The geodesic equations in the Rindler chart are easily obtained from the geodesic
Lagrangian Lagrangian may refer to: Mathematics * Lagrangian function, used to solve constrained minimization problems in optimization theory; see Lagrange multiplier ** Lagrangian relaxation, the method of approximating a difficult constrained problem with ...
; they are : \ddot + \frac \, \dot \, \dot = 0, \; \ddot + x \, \dot^2 = 0, \; \ddot = 0, \; \ddot = 0 Of course, in the original Cartesian chart, the geodesics appear as straight lines, so we could easily obtain them in the Rindler chart using our coordinate transformation. However, it is instructive to obtain and study them independently of the original chart, and we shall do so in this section. From the first, third, and fourth we immediately obtain the ''first integrals'' : \dot = \frac, \; \; \dot = P, \; \; \dot = Q But from the line element we have \varepsilon = -x^2 \, \dot^2 + \dot^2 + \dot^2 + \dot^2 where \varepsilon \in \left\ for timelike, null, and spacelike geodesics, respectively. This gives the fourth first integral, namely : \dot^2 = \left(\varepsilon + \frac \right) - P^2 - Q^2. This suffices to give the complete solution of the geodesic equations. In the case of
null geodesic In general relativity, a geodesic generalizes the notion of a "straight line" to curved spacetime. Importantly, the world line of a particle free from all external, non-gravitational forces is a particular type of geodesic. In other words, a fre ...
s, from \frac \,-\, P^2 \,-\, Q^2 with nonzero E, we see that the x coordinate ranges over the interval : 0 \,<\, x \,<\, \frac. The complete seven parameter family giving any null geodesic through any event in the Rindler wedge, is :\begin t - t_0 &= \operatorname \left( \frac\left \left(P^2 + Q^2\right) - \sqrt\right \right) +\\ & \qquad \operatorname \left( \frac\sqrt \right)\\ x &= \sqrt\\ y - y_0 &= Ps;\;\; z - z_0 = Qs \end Plotting the ''tracks'' of some representative null geodesics through a given event (that is, projecting to the hyperslice t = 0), we obtain a picture which looks suspiciously like the family of all semicircles through a point and orthogonal to the Rindler horizon (See the figure).


The Fermat metric

The fact that in the Rindler chart, the projections of null geodesics into any spatial hyperslice for the Rindler observers are simply semicircular arcs can be verified directly from the general solution just given, but there is a very simple way to see this. A static spacetime is one in which a vorticity-free timelike
Killing vector In mathematics, a Killing vector field (often called a Killing field), named after Wilhelm Killing, is a vector field on a Riemannian manifold (or pseudo-Riemannian manifold) that preserves the metric. Killing fields are the infinitesimal gene ...
field can be found. In this case, we have a uniquely defined family of (identical) spatial hyperslices orthogonal to the corresponding static observers (who need not be inertial observers). This allows us to define a new metric on any of these hyperslices which is conformally related to the original metric inherited from the spacetime, but with the property that geodesics in the new metric (note this is a
Riemannian metric In differential geometry, a Riemannian manifold or Riemannian space , so called after the German mathematician Bernhard Riemann, is a real, smooth manifold ''M'' equipped with a positive-definite inner product ''g'p'' on the tangent space '' ...
on a Riemannian three-manifold) are precisely the projections of the null geodesics of spacetime. This new metric is called the ''Fermat metric'', and in a static spacetime endowed with a coordinate chart in which the line element has the form : ds^2 = g_ \, dt^2 + g_ \, dx^j \, dx^k,\;\; j,\; k \in \ the Fermat metric on t = 0 is simply : d\rho^2 = \frac\left(g_ \, dx^j \, dx^k\right) (where the metric coeffients are understood to be evaluated at t = 0). In the Rindler chart, the timelike translation \partial_t is such a Killing vector field, so this is a static spacetime (not surprisingly, since Minkowski spacetime is of course trivially a static vacuum solution of the
Einstein field equation In the general theory of relativity, the Einstein field equations (EFE; also known as Einstein's equations) relate the geometry of spacetime to the distribution of matter within it. The equations were published by Einstein in 1915 in the form ...
). Therefore, we may immediately write down the Fermat metric for the Rindler observers: : d\rho^2 = \frac\left(dx^2 + dy^2 + dz^2\right),\;\; \forall x > 0,\;\; \forall y, z But this is the well-known line element of ''hyperbolic three-space'' H3 in the ''upper half space chart''. This is closely analogous to the well known ''upper half plane chart'' for the hyperbolic plane H2, which is familiar to generations of complex analysis students in connection with ''conformal mapping problems'' (and much more), and many mathematically minded readers already know that the geodesics of H2 in the upper half plane model are simply semicircles (orthogonal to the circle at infinity represented by the real axis).


Symmetries

Since the Rindler chart is a coordinate chart for Minkowski spacetime, we expect to find ten linearly independent Killing vector fields. Indeed, in the Cartesian chart we can readily find ten linearly independent Killing vector fields, generating respectively one parameter subgroups of
time translation Time translation symmetry or temporal translation symmetry (TTS) is a mathematical transformation in physics that moves the times of events through a common interval. Time translation symmetry is the law that the laws of physics are unchanged ( ...
, three spatials, three rotations and three boosts. Together these generate the (proper isochronous) Poincaré group, the symmetry group of Minkowski spacetime. However, it is instructive to write down and solve the Killing vector equations directly. We obtain four familiar looking Killing vector fields : \partial_t, \; \; \partial_y, \; \; \partial_z, \; \; -z \, \partial_y + y \, \partial_z (time translation, spatial translations orthogonal to the direction of acceleration, and spatial rotation orthogonal to the direction of acceleration) plus six more: :\begin &\exp(\pm t) \, \left( \frac \, \partial_t \pm \left y \, \partial_x - x \, \partial_y \right\right)\\ &\exp(\pm t) \, \left( \frac \, \partial_t \pm \left z \, \partial_x - x \, \partial_z \right\right)\\ &\exp(\pm t) \, \left( \frac \, \partial_t \pm \partial_x \right) \end (where the signs are chosen consistently + or −). We leave it as an exercise to figure out how these are related to the standard generators; here we wish to point out that we must be able to obtain generators equivalent to \partial_T in the Cartesian chart, yet the Rindler wedge is obviously not invariant under this translation. How can this be? The answer is that like anything defined by a system of partial differential equations on a smooth manifold, the Killing equation will in general have locally defined solutions, but these might not exist globally. That is, with suitable restrictions on the group parameter, a Killing flow can always be defined in a suitable ''local neighborhood'', but the flow might not be well-defined globally. This has nothing to do with Lorentzian manifolds per se, since the same issue arises in the study of general
smooth manifold In mathematics, a differentiable manifold (also differential manifold) is a type of manifold that is locally similar enough to a vector space to allow one to apply calculus. Any manifold can be described by a collection of charts (atlas). One ma ...
s.


Notions of distance

One of the many valuable lessons to be learned from a study of the Rindler chart is that there are in fact several ''distinct'' (but reasonable) notions of
distance Distance is a numerical or occasionally qualitative measurement of how far apart objects or points are. In physics or everyday usage, distance may refer to a physical length or an estimation based on other criteria (e.g. "two counties over"). ...
which can be used by the Rindler observers. The first is the one we have tacitly employed above: the induced Riemannian metric on the spatial hyperslices t = t_0. We will call this the ''ruler distance'' since it corresponds to this induced Riemannian metric, but its operational meaning might not be immediately apparent. From the standpoint of physical measurement, a more natural notion of distance between two world lines is the ''radar distance''. This is computed by sending a null geodesic from the world line of our observer (event A) to the world line of some small object, whereupon it is reflected (event B) and returns to the observer (event C). The radar distance is then obtained by dividing the round trip travel time, as measured by an ideal clock carried by our observer. (In Minkowski spacetime, fortunately, we can ignore the possibility of multiple null geodesic paths between two world lines, but in cosmological models and other applications things are not so simple. We should also caution against assuming that this notion of distance between two observers gives a notion which is symmetric under interchanging the observers.) In particular, consider a pair of Rindler observers with coordinates x = x_0, \; y = 0,\; z = 0 and x = x_0 + h, \; y = 0,\; z = 0 respectively. (Note that the first of these, the trailing observer, is accelerating a bit harder, in order to keep up with the leading observer). Setting dy = dz = 0 in the Rindler line element, we readily obtain the equation of null geodesics moving in the direction of acceleration: : t - t_0 = \log\left(\frac\right) Therefore, the radar distance between these two observers is given by : x_0 \, \log \left(1 + \frac \right) = h - \frac + O \left( h^3 \right) This is a bit smaller than the ruler distance, but for nearby observers the discrepancy is negligible. A third possible notion of distance is this: our observer measures the ''angle'' subtended by a unit disk placed on some object (not a point object), as it appears from his location. We call this the ''optical diameter distance''. Because of the simple character of null geodesics in Minkowski spacetime, we can readily determine the optical distance between our pair of Rindler observers (aligned with the direction of acceleration). From a sketch it should be plausible that the optical diameter distance scales like h + \frac + O \left( h^3 \right) . Therefore, in the case of a trailing observer estimating distance to a leading observer (the case h > 0), the optical distance is a bit larger than the ruler distance, which is a bit larger than the radar distance. The reader should now take a moment to consider the case of a leading observer estimating distance to a trailing observer. There are other notions of distance, but the main point is clear: while the values of these various notions will in general disagree for a given pair of Rindler observers, they all agree that ''every pair of Rindler observers maintains constant distance''. The fact that ''very nearby'' Rindler observers are mutually stationary follows from the fact, noted above, that the expansion tensor of the Rindler congruence vanishes identically. However, we have shown here that in various senses, this rigidity property holds at larger scales. This is truly a remarkable rigidity property, given the well-known fact that in relativistic physics, ''no rod can be accelerated rigidly'' (and ''no disk can be spun up rigidly'') — at least, not without sustaining inhomogeneous stresses. The easiest way to see this is to observe that in Newtonian physics, if we "kick" a rigid body, all elements of matter in the body will immediately change their state of motion. This is of course incompatible with the relativistic principle that no information having any physical effect can be transmitted faster than the speed of light. It follows that if a rod is accelerated by some external force applied anywhere along its length, the elements of matter in various different places in the rod cannot all feel the same magnitude of acceleration if the rod is not to extend without bound and ultimately break. In other words, an accelerated rod which does not break must sustain stresses which vary along its length. Furthermore, in any thought experiment with time varying forces, whether we "kick" an object or try to accelerate it gradually, we cannot avoid the problem of avoiding mechanical models which are inconsistent with relativistic kinematics (because distant parts of the body respond too quickly to an applied force). Returning to the question of the operational significance of the ruler distance, we see that this should be the distance which our observers will obtain should they very slowly pass from hand to hand a small ruler which is repeatedly set end to end. But justifying this interpretation in detail would require some kind of material model.


Generalization to curved spacetimes

Rindler coordinates as described above can be generalized to curved spacetime, as Fermi normal coordinates. The generalization essentially involves constructing an appropriate orthonormal tetrad and then transporting it along the given trajectory using the
Fermi–Walker transport Fermi–Walker transport is a process in general relativity used to define a coordinate system or reference frame such that all curvature in the frame is due to the presence of mass/energy density and not to arbitrary spin or rotation of the fram ...
rule. For details, see the paper by Ni and Zimmermann in the references below. Such a generalization actually enables one to study inertial and gravitational effects in an Earth-based laboratory, as well as the more interesting coupled inertial-gravitational effects.


History


Overview

;Kottler–Møller and Rindler coordinates
Albert Einstein Albert Einstein ( ; ; 14 March 1879 – 18 April 1955) was a German-born theoretical physicist, widely acknowledged to be one of the greatest and most influential physicists of all time. Einstein is best known for developing the theory ...
(1907) studied the effects within a uniformly accelerated frame, obtaining equations for coordinate dependent
time dilation In physics and relativity, time dilation is the difference in the elapsed time as measured by two clocks. It is either due to a relative velocity between them ( special relativistic "kinetic" time dilation) or to a difference in gravitational ...
and
speed of light The speed of light in vacuum, commonly denoted , is a universal physical constant that is important in many areas of physics. The speed of light is exactly equal to ). According to the special theory of relativity, is the upper limit ...
equivalent to (), and in order to make the formulas independent of the observer's origin, he obtained time dilation () in formal agreement with Radar coordinates. While introducing the concept of
Born rigidity Born rigidity is a concept in special relativity. It is one answer to the question of what, in special relativity, corresponds to the rigid body of non-relativistic classical mechanics. The concept was introduced by Max Born (1909),Born (1909b) wh ...
, Max Born (1909) noted that the formulas for hyperbolic motion can be used as transformations into a "hyperbolically accelerated reference system" (german: hyperbolisch beschleunigtes Bezugsystem) equivalent to (). Born's work was further elaborated by Arnold Sommerfeld (1910) and Max von Laue (1911) who both obtained () using imaginary numbers, which was summarized by
Wolfgang Pauli Wolfgang Ernst Pauli (; ; 25 April 1900 – 15 December 1958) was an Austrian theoretical physicist and one of the pioneers of quantum physics. In 1945, after having been nominated by Albert Einstein, Pauli received the Nobel Prize in Physics ...
(1921) who besides coordinates () also obtained metric () using imaginary numbers. Einstein (1912) studied a static gravitational field and obtained the Kottler–Møller metric () as well as approximations to formulas () using a coordinate dependent speed of light. Hendrik Lorentz (1913) obtained coordinates similar to (, , ) while studying Einstein's equivalence principle and the uniform gravitational field. A detailed description was given by
Friedrich Kottler Friedrich Kottler (December 10, 1886 – May 11, 1965) was an Austrian theoretical physicist. He was a Privatdozent before he got a professorship in 1923 at the University of Vienna. Life In 1938, after the Anschluss, he lost his profess ...
(1914),Kottler (1912), pp. 1715; Kottler (1914a), Table I; pp. 747–748; Kottler (1914b), pp. 488–489, 503; Kottler (1916), pp. 958–959; (1918), pp. 453–454; who formulated the corresponding orthonormal
tetrad Tetrad ('group of 4') or tetrade may refer to: * Tetrad (area), an area 2 km x 2 km square * Tetrad (astronomy), four total lunar eclipses within two years * Tetrad (chromosomal formation) * Tetrad (general relativity), or frame field ** Tetra ...
, transformation formulas and metric (, ). Also Karl Bollert (1922) obtained the metric () in his study of uniform acceleration and uniform gravitational fields. In a paper concerned with Born rigidity,
Georges Lemaître Georges Henri Joseph Édouard Lemaître ( ; ; 17 July 1894 – 20 June 1966) was a Belgian Catholic priest, theoretical physicist, mathematician, astronomer, and professor of physics at the Catholic University of Louvain. He was the first to t ...
(1924) obtained coordinates and metric (, ).
Albert Einstein Albert Einstein ( ; ; 14 March 1879 – 18 April 1955) was a German-born theoretical physicist, widely acknowledged to be one of the greatest and most influential physicists of all time. Einstein is best known for developing the theory ...
and
Nathan Rosen Nathan Rosen (Hebrew: נתן רוזן; March 22, 1909 – December 18, 1995) was an American-Israeli physicist noted for his study on the structure of the hydrogen atom and his work with Albert Einstein and Boris Podolsky on entangled wave functio ...
(1935) described (, ) as the "well known" expressions for a homogeneous gravitational field. After Christian Møller (1943) obtained (, ) in as study related to homogeneous gravitational fields, he (1952) as well as Misner & Thorne & Wheeler (1973) used
Fermi–Walker transport Fermi–Walker transport is a process in general relativity used to define a coordinate system or reference frame such that all curvature in the frame is due to the presence of mass/energy density and not to arbitrary spin or rotation of the fram ...
to obtain the same equations. While these investigations were concerned with flat spacetime,
Wolfgang Rindler Wolfgang Rindler (18 May 1924 – 8 February 2019) was a physicist working in the field of general relativity where he is known for introducing the term "event horizon", Rindler coordinates, and (in collaboration with Roger Penrose) for the use of ...
(1960) analyzed hyperbolic motion in curved spacetime, and showed (1966) the analogy between the hyperbolic coordinates (, ) in flat spacetime with Kruskal coordinates in Schwarzschild space. This influenced subsequent writers in their formulation of Unruh radiation measured by an observer in hyperbolic motion, which is similar to the description of
Hawking radiation Hawking radiation is theoretical black body radiation that is theorized to be released outside a black hole's event horizon because of relativistic quantum effects. It is named after the physicist Stephen Hawking, who developed a theoretical a ...
of black holes. ;Horizon Born (1909) showed that the inner points of a Born rigid body in hyperbolic motion can only be in the region X/\left(X^-T^\right)>0.Born (1909), p. 35 Sommerfeld (1910) defined that the coordinates allowed for the transformation between inertial and hyperbolic coordinates must satisfy T.Sommerfeld (1910), p. 672 Kottler (1914)Kottler (1914), pp. 489-490 defined this region as X^-T^>0, and pointed out the existence of a "border plane" (german: Grenzebene) c^2/\alpha+x, beyond which no signal can reach the observer in hyperbolic motion. This was called the "horizon of the observer" (german: Horizont des Beobachters) by Bollert (1922).Bollert (192), pp. 194-196 Rindler (1966) demonstrated the relation between such a horizon and the horizon in Kruskal coordinates. ;Radar coordinates Using Bollert's formalism,
Stjepan Mohorovičić Stjepan Mohorovičić (August 20, 1890 – February 13, 1980) was a Croatian physicist, geophysicist and meteorologist. Biography Mohorovičić was born in the town of Bakar. His father is the world-famous geophysicist Andrija Mohorovičić. He ...
(1922) made a different choice for some parameter and obtained metric () with a printing error, which was corrected by Bollert (1922b) with another printing error, until a version without printing error was given by Mohorovičić (1923). In addition, Mohorovičić erroneously argued that metric (, now called Kottler–Møller metric) is incorrect, which was rebutted by Bollert (1922).Bollert (1922b), p. 189 Metric () was rediscovered by Harry Lass (1963), who also gave the corresponding coordinates () which are sometimes called "Lass coordinates". Metric (), as well as (, ), was also derived by
Fritz Rohrlich Fritz Rohrlich (May 12, 1921 – November 14, 2018) was an American theoretical physicist and educator who published in the fields of quantum electrodynamics, classical electrodynamics of charged particles, and the philosophy of science. Life and ...
(1963). Eventually, the Lass coordinates (, ) were identified with Radar coordinates by Desloge & Philpott (1987).


Table with historical formulas

{, style="text-align: center;" , valign="top", {, class="wikitable" !Einstein (1907)Einstein (1907), §§ 18-21 , - , {\scriptstyle \begin{matrix}\sigma=\tau\left(1+\frac{\gamma\xi}{c^{2\right)\\ \sigma=\tau e^{\gamma\xi/c^{2\\ c\left(1+\frac{\gamma\xi}{c^{2\right) \end{matrix , - !Born (1909)Born (1909), p. 25 , - , {\scriptstyle \begin{matrix}x=-q\xi,\ y=\eta,\ z=\zeta,\ t=\frac{p}{c^{2\xi\\ \left(p=x_{\tau},\ q=-t_{\tau}=\sqrt{1+p^{2}/c^{2\right)\\ \boldsymbol{\downarrow}\\ x^{2}-c^{2}t^{2}=\xi^{2} \end{matrix , - !Herglotz (1909)Herglotz (1909), pp. 408, 414 , - , {\scriptstyle \begin{matrix}\begin{align}x & =x'\\ y & =y'\\ t-z & =(t'-z')e^{\vartheta}\\ t+z & =(t'+z')e^{-\vartheta} \end{align} \\ \boldsymbol{\downarrow}\\ x=x_{0},\quad y=y_{0},\quad z=\sqrt{z_{0}^{2}+t^{2 \end{matrix , - !Sommerfeld (1910)Sommerfeld (1910), pp. 670-671 , - , \scriptstyle\begin{align} x & =r\cos\varphi\\ y & =y'\\ z & =z'\\ l & =r\sin\varphi\\ \varphi & =i\psi,\ l =ict \end{align} , - !von Laue (1911)von Laue (1911), p. 109 , - , \scriptstyle\begin{align} X & =R\cos\varphi\\ L & =R\sin\varphi\\ R^{2} & =X^{2}+L^{2}\\ \tan\varphi & =\frac{L}{X} \end{align} , - !Einstein (1912)Einstein (1912), pp. 358-359 , - , {\scriptstyle \begin{matrix}d\xi^{2}-d\tau^{2}=dx^{2}-c^{2}dt^{2}\\ \boldsymbol{\downarrow}\\ c=c_{0}+ax\\ \boldsymbol{\downarrow}\\ \begin{align}\xi & =x+\frac{ac}{2}t^{2}\\ \eta & =y\\ \zeta & =z\\ \tau & =ct \end{align} \end{matrix , - !Kottler (1912)Kottler (1912), pp. 1715 , - , {\scriptstyle \begin{align}x^{(1)} & =x_{0}^{(1)}\\ x^{(2)} & =x_{0}^{(2)}\\ x^{(3)} & =b\cos i\varphi\\ x^{(4)} & =b\sin i\varphi \end{align , - !Lorentz (1913)Lorentz (1913), pp. 34-38; 50-52 , - , {\scriptstyle \begin{matrix}dc=\frac{g}{c}dz\\ \hline \begin{align}z & =a\left(z'-z_{0}^{\prime}\right)\\ ct & =b\left(z'-z_{0}^{\prime}\right)\\ a & =\frac{1}{2}\left(e^{kt'}+e^{-kt}\right)\\ b & =\frac{1}{2}\left(e^{kt'}-e^{-kt}\right) \end{align} \\ \boldsymbol{\downarrow}\\ c'=k\left(z'-z_{0}^{\prime}\right),\ z'-z_{0}^{\prime}=\frac{c^{2{g}\\ \boldsymbol{\downarrow}\\ \begin{align} & dx^{2}+dy^{2}+dz^{2}-c^{2}dt\\ & =dx^{\prime2}+dy^{\prime2}+dz^{\prime2}-c^{\prime2}dt^{\prime2} \end{align} \end{matrix , valign="top", {, class="wikitable" !Kottler (1914a)Kottler (1914a), Table I; pp. 747-748 , - , {\scriptstyle \begin{matrix}\begin{align}x^{(1)} & =x_{0}^{(1)}\\ x^{(2)} & =x_{0}^{(2)}\\ x^{(3)} & =b\cos iu\\ x^{(4)} & =b\sin iu \end{align} \\ \boldsymbol{\downarrow}\\ ds^{2}=-c^{2}d\tau^{2}=b^{2}(du)^{2}\\ \boldsymbol{\downarrow}\\ \begin{matrix}c_{1}^{(1)}=0, & & c_{1}^{(2)}=0, & & c_{1}^{(3)}=-\sin iu, & & c_{1}^{(4)}=\cos iu,\\ c_{2}^{(1)}=0, & & c_{2}^{(2)}=0, & & c_{2}^{(3)}=-\cos iu, & & c_{2}^{(4)}=-\sin iu, \end{matrix}\\ \boldsymbol{\downarrow}\\ dS^{2}=(dX')^{2}+(dY')^{2}+(dZ')^{2}-\left(c+\frac{Z'c}{b}\right)^{2}dT'\\ \boldsymbol{\downarrow}\\ c'=c+\frac{Z'c^{2{b}\cdot\frac{1}{c} \end{matrix , - !Kottler (1914b)Kottler (1914b), pp. 488-489, 503 , - , \scriptstyle\begin{matrix}\begin{matrix} c_{1}^{(1)}=0, & & c_{1}^{(2)}=0, & & c_{1}^{(3)}=\frac{1}{i}\sinh u, & & c_{1}^{(4)}=\cosh u,\\ c_{2}^{(1)}=0, & & c_{2}^{(2)}=0, & & c_{2}^{(3)}=\frac{1}{i}\cosh u, & & c_{2}^{(4)}=-\sinh u,\\ c_{3}^{(1)}=1, & & c_{3}^{(2)}=0, & & c_{3}^{(3)}=0, & & c_{3}^{(4)}=0,\\ c_{4}^{(1)}=0, & & c_{4}^{(2)}=1, & & c_{4}^{(3)}=0, & & c_{4}^{(4)}=0, \end{matrix}\\ \boldsymbol{\downarrow}\\ X=x+\Delta^{(2)}c_{2}+\Delta^{(3)}c_{3}+\Delta^{(4)}c_{4}\\ \boldsymbol{\downarrow}\\ \begin{align} X & =x_{0}+\mathfrak{X}'\\ Y & =y_{0}+\mathfrak{Y}'\\ Z & =\left(b+\mathfrak{Z}'\right)\cosh\mathfrak{u}\\ cT & =\left(b+\mathfrak{Z}'\right)\sinh\mathfrak{u} \end{align}\\ \left(\Delta^{(2)}=\mathfrak{X}',\ \Delta^{(3)}=\mathfrak{Y}',\ \Delta^{(4)}=\mathfrak{Z}'\right)\\ \boldsymbol{\downarrow}\\ \begin{align} \mathfrak{X}' & =X_{0}-x_{0}+q_{x}T\\ \mathfrak{Y}' & =Y_{0}-y_{0}+q_{y}T\\ b+\mathfrak{Z}' & =\sqrt{\left(Z_{0}+q_{x}T\right)^{2}-c^{2}T^{2\\ c\mathfrak{T}' & =b\operatorname{artanh}\frac{cT}{Z_{0}+q_{x}T} \end{align}\\ \left(X=X_{0}+q_{x}T,\ Y=Y_{0}+q_{y}T,\ Z=Z_{0}+q_{x}T\right)\\ \boldsymbol{\downarrow}\\ dS^{2}=(d\mathfrak{X}')^{2}+(d\mathfrak{Y}')^{2}+(d\mathfrak{Z}')^{2}-c^{2}\left(\frac{b+\mathfrak{Z}'}{b^{2\right)^{2}(d\mathfrak{T}')^{2} \end{matrix} , - !Kottler (1916, 1918)Kottler (1916), pp. 958-959; (1918), pp. 453-454 , - , \scriptstyle\begin{matrix}\begin{align} x & =x'\\ y & =y'\\ \frac{c^{2{\gamma}+z & =\left(\frac{c^{2{\gamma}+z'\right)\cosh\frac{\gamma t'}{c}\\ ct & =\left(\frac{c^{2{\gamma}+z'\right)\sinh\frac{\gamma t'}{c} \end{align}\\ \boldsymbol{\downarrow}\\ ds^{2}=dx^{\prime2}+dy^{\prime2}+dz^{\prime2}-\left(c+\frac{\gamma}{c}z'\right){}^{2}dt^{\prime2} \end{matrix} , valign="top", {, class="wikitable" !Pauli (1921)Pauli (1921), pp. 647-648 , - , \scriptstyle\begin{matrix}\begin{align} x^{1} & =\varrho\cos\varphi\\ x^{4} & =\varrho\sin\varphi \end{align}\\ \boldsymbol{\downarrow}\\ ds^{2}=\left(d\xi^{1}\right)^{2}+\left(d\xi^{2}\right)^{2}+\left(d\xi^{3}\right)^{2}+\left(\xi^{1}\right)^{2}\left(d\xi^{4}\right)^{2}\\ \left(\xi^{(1)}=\varrho,\ \xi^{(2)}=x^{(2)},\ \xi^{(3)}=x^{(3)},\ \xi^{(4)}=\varphi\right) \end{matrix} , - !Bollert (1922)Bollert (1922a), p. 261, 266 , - , {\scriptstyle \begin{matrix}ds^{2}=c^{2}\left(1+\frac{\gamma_{0}x}{c^{2\right)d\tau^{2}-dx^{2}-dy^{2}-dz^{2}\\ \hline ds^{2}=g_{44}dx_{4}^{2}+g_{11}dx_{1}^{2}+g_{22}\left(dx_{2}^{2}+dx_{3}^{2}\right)\\ \boldsymbol{\downarrow}\\ V''-\frac{g_{11{2g_{11V'=0\\ \left(g_{22}=-1,\ g_{11}=-1,\ V''=0,\ V=ax+b\right)\\ \boldsymbol{\downarrow}\\ ds^{2}=dx_{4}^{2}(ax+b)^{2}-dx^{2}-dy^{2}-dz^{2} \end{matrix , - !Mohorovičić (1922, 1923); Bollert (1922b)Mohorovičić (1922), p. 92, without x_1 in the exponent due to a printing error, which was corrected by Bollert (1922b), p.189, as well as Mohorovičić (1923), p. 54 , - , {\scriptstyle \begin{matrix}\text{Mohorovičić (1922):}\\ g_{11}=g_{44}=V^{2},\ VV''-V'^{2}=0,\ V\left(x_{1}\right)=e^{ax_{1\\ \boldsymbol{\downarrow}\\ ds^{2}=e^{2a}\left(-dx_{4}^{2}+dx_{1}^{2}\right)+dx_{2}^{2}+dx_{3}^{2}\\ \\ \text{corrected by Bollert (1922b):}\\ ds^{2}=e^{2ax}\left(-dx_{4}^{2}+dx_{1}^{2}\right)+dx_{2}^{2}+dx_{3}^{2}\\ \\ \text{final correction by Mohorovičić (1923):}\\ ds^{2}=e^{2ax_{1\left(-dx_{4}^{2}+dx_{1}^{2}\right)+dx_{2}^{2}+dx_{3}^{2} \end{matrix , - !Lemaître (1924)Lemaitre (1921), pp. 166, 168 , - , {\scriptstyle \begin{matrix}\begin{align}1+g\xi= & (1+gx)\cosh gt\\ g\tau= & (1+gx)\sinh gt \end{align} \\ \boldsymbol{\downarrow}\\ ds^{2}=-dx^{2}-dy^{2}-dz^{2}+(1+gx)^{2}dt^{2} \end{matrix , - !Einstein & Rosen (1935)Einstein & Rosen (1935, p. 74 , - , \scriptstyle\begin{matrix}\begin{align} \xi_{1} & =x_{1}\cosh\alpha x_{4}\\ \xi_{2} & =x_{2}\\ \xi_{3} & =x_{3}\\ \xi_{4} & =x_{1}\sinh\alpha x_{4} \end{align}\\ \boldsymbol{\downarrow}\\ ds^{2}=-dx_{1}^{2}-dx_{2}^{2}-dx_{3}^{2}+\alpha{}^{2}x_{1}^{2}dx_{4}^{2} \end{matrix} , - !Møller (1952)Møller (1952), pp. 121-123; 255-258 , - , \scriptstyle\begin{matrix}\alpha_{ik}=\left(\begin{matrix}U_{4}/ic & 0 & 0 & iU_{1}/c\\ 0 & 1 & 0 & 0\\ 0 & 0 & 1 & 0\\ U_{1}/ic & 0 & 0 & U_{4}/ic \end{matrix}\right)\\ U_{i}=\left(c\sinh\frac{g\tau}{c},\ 0,0,\ ig\cosh\frac{g\tau}{c}\right)\\ \boldsymbol{\downarrow}\\ X_{i}=\mathbf{f}_{i}(t)+x^{\prime\kappa}\alpha_{\kappa i}(\tau)\\ \boldsymbol{\downarrow}\\ \begin{align} X & =\frac{c^{2{g}\left(\cosh\frac{gt}{c}-1\right)+x\cosh\frac{gt}{c}\\ Y & =y\\ Z & =z\\ T & =\frac{c}{g}\sinh\frac{gt}{c}+x\frac{\sinh\frac{gt}{c{c} \end{align}\\ \boldsymbol{\downarrow}\\ ds^{2}=dx^{2}+dy^{2}+dz^{2}-c^{2}dt^{2}\left(1+gx/c^{2}\right)^{2}\\ \\ \end{matrix}


See also

*
Bell's spaceship paradox Bell's spaceship paradox is a thought experiment in special relativity. It was designed by E. Dewan and M. Beran in 1959 and became more widely known when J. S. Bell included a modified version.J. S. Bell: ''How to teach special relativity'', Prog ...
, for a sometimes controversial subject often studied using Rindler coordinates. * Born coordinates, for another important coordinate system adapted to the motion of certain accelerated observers in
Minkowski spacetime In mathematical physics, Minkowski space (or Minkowski spacetime) () is a combination of Three-dimensional space, three-dimensional Euclidean space and time into a four-dimensional manifold where the spacetime interval between any two Event (rel ...
. *
Congruence (general relativity) In general relativity, a congruence (more properly, a congruence of curves) is the set of integral curves of a (nowhere vanishing) vector field in a four-dimensional Lorentzian manifold which is interpreted physically as a model of spacetime. Ofte ...
*
Ehrenfest paradox The Ehrenfest paradox concerns the rotation of a "rigid" disc in the theory of relativity. In its original 1909 formulation as presented by Paul Ehrenfest in relation to the concept of Born rigidity within special relativity, it discusses an idea ...
, for a sometimes controversial subject often studied using Born coordinates. *
Frame fields in general relativity A frame field in general relativity (also called a tetrad or vierbein) is a set of four pointwise-orthonormal vector fields, one timelike and three spacelike, defined on a Lorentzian manifold that is physically interpreted as a model of spacetime ...
*
General relativity resources General relativity, also known as the general theory of relativity and Einstein's theory of gravity, is the geometric theory of gravitation published by Albert Einstein in 1915 and is the current description of gravitation in modern physics. G ...
* Milne model *
Raychaudhuri equation In general relativity, the Raychaudhuri equation, or Landau–Raychaudhuri equation, is a fundamental result describing the motion of nearby bits of matter. The equation is important as a fundamental lemma for the Penrose–Hawking singularity the ...
*
Unruh effect The Unruh effect (also known as the Fulling–Davies–Unruh effect) is a kinematic prediction of quantum field theory that an accelerating observer will observe a thermal bath, like blackbody radiation, whereas an inertial observer would observe ...


References

; First edition 1911, second expanded edition 1913, third expanded edition 1919.
In English:


Historical sources

; English translatio
On the relativity principle and the conclusions drawn from it
at Einstein paper project. , English translatio
The Speed of Light and the Statics of the Gravitational Field
at Einstein paper project.
; First edition 1911, second expanded edition 1913, third expanded edition 1919. New edition 2013: Editor: Domenico Giulini, Springer, 2013 .


Further reading

Useful background: * See ''Chapter 4'' for background concerning vector fields on smooth manifolds. * See ''Chapter 8'' for a derivation of the Fermat metric. Rindler coordinates: * * See ''Section 6.6''. * * Rindler horizon:
eprint version
*{{cite journal , author1=Barceló, Carlos , author2=Liberati, Stefano , author3=Visser, Matt , name-list-style=amp , title=Analogue Gravity , journal=Living Reviews in Relativity , arxiv=gr-qc/0505065, doi=10.12942/lrr-2005-12, bibcode=2005LRR.....8...12B , volume=8 , issue=1 , pages=12 , year=2005 , pmid=28179871 , pmc=5255570 Theory of relativity Coordinate charts in general relativity Acceleration