fbpx
Wikipedia

Rindler coordinates

Rindler coordinates are a coordinate system used in the context of special relativity to describe the hyperbolic acceleration of a uniformly accelerating reference frame in flat spacetime. In relativistic physics the coordinates of a hyperbolically accelerated reference frame[H 1][1] constitute an important and useful coordinate chart representing part of flat Minkowski spacetime.[2][3][4][5] In special relativity, a uniformly accelerating particle undergoes hyperbolic motion, for which a uniformly accelerating 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) and Proper reference frame (flat spacetime).

In this article, the speed of light is defined by c = 1, the inertial coordinates are (X, Y, Z, T), and the hyperbolic coordinates are (x, y, z, t). These hyperbolic coordinates can be separated into two main variants depending on the accelerated observer's position: If the observer is located at time T = 0 at position X = 1/α (with α as the constant proper acceleration measured by a comoving accelerometer), then the hyperbolic coordinates are often called Rindler coordinates with the corresponding Rindler metric.[6] If the observer is located at time T = 0 at position X = 0, then the hyperbolic coordinates are sometimes called Møller coordinates[1] or Kottler–Møller coordinates with the corresponding Kottler–Møller metric.[7] An alternative chart often related to observers in hyperbolic motion is obtained using Radar coordinates[8] which are sometimes called Lass coordinates.[9][10] Both the Kottler–Møller coordinates as well as Lass coordinates are denoted as Rindler coordinates as well.[11]

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 (1907, 1912),[H 2] Max Born (1909),[H 1] Arnold Sommerfeld (1910),[H 3] Max von Laue (1911),[H 4] Hendrik Lorentz (1913),[H 5] Friedrich Kottler (1914),[H 6] Wolfgang Pauli (1921),[H 7] Karl Bollert (1922),[H 8] Stjepan Mohorovičić (1922),[H 9] Georges Lemaître (1924),[H 10] Einstein & Nathan Rosen (1935),[H 2] Christian Møller (1943, 1952),[H 11] Fritz Rohrlich (1963),[12] Harry Lass (1963),[13] and in relation to both flat and curved spacetime of general relativity by Wolfgang Rindler (1960, 1966).[14][15] For details and sources, see § History.

Characteristics of the Rindler frame edit

 
Rindler chart, for   in equation (1a), plotted on a Minkowski diagram. The dashed lines are the Rindler horizons

The worldline of a body in hyperbolic motion having constant proper acceleration   in the  -direction as a function of proper time   and rapidity   can be given by[16]

 

where   is constant and   is variable, with the worldline resembling the hyperbola  . Sommerfeld[H 3][17] showed that the equations can be reinterpreted by defining   as variable and   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  , the transformation formulas between the inertial coordinates and the hyperbolic coordinates are consequently:[6][9]

 

(1a)

with the inverse

 

Differentiated and inserted into the Minkowski metric

 

the metric in the hyperbolically accelerated frame follows as

 

(1b)

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  , which is often called the Rindler wedge, if   represents the proper acceleration (along the hyperbola  ) 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   (the Rindler horizon) having greater proper acceleration. All the Rindler observers are instantaneously at rest at time   in the inertial frame, and at this time a Rindler observer with proper acceleration   will be at position   (really  , but we assume units where  ), 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  , then when defining a Rindler coordinate system we have a choice of which Rindler observer's proper time will be equal to the coordinate time   in Rindler coordinates, and this observer's proper acceleration defines the value of   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).[18] 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  , so that   can be eliminated from the equations.

The above equation has been simplified for  . The unsimplified equation is more convenient for finding the Rindler Horizon distance, given an acceleration  .

 

The remainder of the article will follow the convention of setting both   and  , so units for   and   will be 1 unit  . Be mindful that setting   light-second/second2 is very different from setting   light-year/year2. Even if we pick units where  , the magnitude of the proper acceleration   will depend on our choice of units: for example, if we use units of light-years for distance, (  or  ) and years for time, (  or  ), this would mean   light year/year2, equal to about 9.5 meters/second2, while if we use units of light-seconds for distance, (  or  ), and seconds for time, (  or  ), this would mean   light-second/second2, or 299 792 458 meters/second2).

Variants of transformation formulas edit

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.[19] Depending on the choice of origin of these coordinates, one can derive the metric, the time dilation between the time at the origin   and   at point  , and the coordinate light speed   (this variable speed of light 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   in inertial frames:

  at   Transformation, Metric, Time dilation and Coordinate speed of light
  Kottler–Møller coordinates[H 12][20][21][22]
 

(2a)

 

(2b)

 

(2c)
Rindler coordinates[23][24][18]
 
 

(2d)

 

(2e)

 

(2f)
Radar coordinates (Lass coordinates)[25][26][8][9]
 
 

(2g)
 

(2h)

 

(2i)

The Rindler observers edit

In the new chart (1a) with   and  , it is natural to take the coframe field

 

which has the dual frame field

 

This defines a local Lorentz frame in the tangent space at each event (in the region covered by our Rindler chart, namely the Rindler wedge). The integral curves of the timelike unit vector field   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  . Using the coordinate transformation above, we find that these correspond to hyperbolic arcs in the original Cartesian chart.

 
Some representative Rindler observers (navy blue hyperbolic arcs) depicted using the Cartesian chart. The red lines at 45 degrees from the vertical represent the Rindler horizon; the Rindler coordinate system is only defined to the right of this boundary.

As with any timelike congruence in any Lorentzian manifold, this congruence has a kinematic decomposition (see Raychaudhuri equation). 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

 

That is, each Rindler observer is accelerating in the   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  ; these appear as horizontal half-planes in the Rindler chart and as half-planes through   in the Cartesian chart (see the figure above). Setting   in the line element, we see that these have the ordinary Euclidean geometry,  . 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 edit

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. 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 edit

 
A representative Minkowski observer (navy blue hyperbolic secant curve) depicted using the Rindler chart. The Rindler horizon is shown in red.

It is worthwhile to also introduce an alternative frame, given in the Minkowski chart by the natural choice

 

Transforming these vector fields using the coordinate transformation given above, we find that in the Rindler chart (in the Rinder wedge) this frame becomes

 

Computing the kinematic decomposition of the timelike congruence defined by the timelike unit vector field  , we find that the expansion and vorticity again vanishes, and in addition the acceleration vector vanishes,  . 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  . Specifically, in Rindler coordinates, the world line of the Minkowski observer passing through the event   is

 

where   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,   and we have drawn (correctly scaled and boosted) the light cones at  .

The Rindler horizon edit

The Rindler coordinate chart has a coordinate singularity at x = 0, where the metric tensor (expressed in the Rindler coordinates) has vanishing determinant. 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 T2 = X2X > 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 of a black hole, each observer hovering at a constant radius in Schwarzschild coordinates. 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 edit

The geodesic equations in the Rindler chart are easily obtained from the geodesic Lagrangian; they are

 

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.

 
Some representative null geodesics (black hyperbolic semicircular arcs) projected into the spatial hyperslice t = 0 of the Rindler observers. The Rindler horizon is shown as a magenta plane.

From the first, third, and fourth we immediately obtain the first integrals

 

But from the line element we have   where   for timelike, null, and spacelike geodesics, respectively. This gives the fourth first integral, namely

 .

This suffices to give the complete solution of the geodesic equations.

In the case of null geodesics, from   with nonzero  , we see that the x coordinate ranges over the interval

 .

The complete seven parameter family giving any null geodesic through any event in the Rindler wedge, is

 

Plotting the tracks of some representative null geodesics through a given event (that is, projecting to the hyperslice  ), 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).[27]

The Fermat metric edit

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 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 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

 

the Fermat metric on   is simply

 

(where the metric coeffients are understood to be evaluated at  ).

In the Rindler chart, the timelike translation   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). Therefore, we may immediately write down the Fermat metric for the Rindler observers:

 

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 edit

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, 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

 

(time translation, spatial translations orthogonal to the direction of acceleration, and spatial rotation orthogonal to the direction of acceleration) plus six more:

 

(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   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 manifolds.

Notions of distance edit

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 which can be used by the Rindler observers.

 
Operational meaning of the radar distance between two Rindler observers (navy blue vertical lines). The Rindler horizon is shown at left (red vertical line). The world line of the radar pulse is also depicted, together with the (properly scaled) light cones at events A, B, C.

The first is the one we have tacitly employed above: the induced Riemannian metric on the spatial hyperslices  . 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[which?] 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   and   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   in the Rindler line element, we readily obtain the equation of null geodesics moving in the direction of acceleration:

 

Therefore, the radar distance between these two observers is given by

 

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  . Therefore, in the case of a trailing observer estimating distance to a leading observer (the case  ), 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 edit

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 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 edit

Overview edit

Kottler–Møller and Rindler coordinates

Albert Einstein (1907)[H 13] studied the effects within a uniformly accelerated frame, obtaining equations for coordinate dependent time dilation and speed of light equivalent to (2c), and in order to make the formulas independent of the observer's origin, he obtained time dilation (2i) in formal agreement with Radar coordinates. While introducing the concept of Born rigidity, Max Born (1909)[H 14] noted that the formulas for hyperbolic motion can be used as transformations into a "hyperbolically accelerated reference system" (German: hyperbolisch beschleunigtes Bezugsystem) equivalent to (2d). Born's work was further elaborated by Arnold Sommerfeld (1910)[H 15] and Max von Laue (1911)[H 16] who both obtained (2d) using imaginary numbers, which was summarized by Wolfgang Pauli (1921)[16] who besides coordinates (2d) also obtained metric (2e) using imaginary numbers. Einstein (1912)[H 17] studied a static gravitational field and obtained the Kottler–Møller metric (2b) as well as approximations to formulas (2a) using a coordinate dependent speed of light.[28] Hendrik Lorentz (1913)[H 18] obtained coordinates similar to (2d, 2e, 2f) while studying Einstein's equivalence principle and the uniform gravitational field.

A detailed description was given by Friedrich Kottler (1914),[H 19] who formulated the corresponding orthonormal tetrad, transformation formulas and metric (2a, 2b). Also Karl Bollert (1922)[H 20] obtained the metric (2b) in his study of uniform acceleration and uniform gravitational fields. In a paper concerned with Born rigidity, Georges Lemaître (1924)[H 21] obtained coordinates and metric (2a, 2b). Albert Einstein and Nathan Rosen (1935) described (2d, 2e) as the "well known" expressions for a homogeneous gravitational field.[H 22] After Christian Møller (1943)[H 11] obtained (2a, 2b) in as study related to homogeneous gravitational fields, he (1952)[H 23] as well as Misner & Thorne & Wheeler (1973)[2] used Fermi–Walker transport to obtain the same equations.

While these investigations were concerned with flat spacetime, Wolfgang Rindler (1960)[14] analyzed hyperbolic motion in curved spacetime, and showed (1966)[15] the analogy between the hyperbolic coordinates (2d, 2e) 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 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  .[H 24] Sommerfeld (1910) defined that the coordinates allowed for the transformation between inertial and hyperbolic coordinates must satisfy  .[H 25] Kottler (1914)[H 26] defined this region as  , and pointed out the existence of a "border plane" (German: Grenzebene)  , 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).[H 27] Rindler (1966)[15] demonstrated the relation between such a horizon and the horizon in Kruskal coordinates.

Radar coordinates

Using Bollert's formalism, Stjepan Mohorovičić (1922)[H 28] made a different choice for some parameter and obtained metric (2h) 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 (2b, now called Kottler–Møller metric) is incorrect, which was rebutted by Bollert (1922).[H 29] Metric (2h) was rediscovered by Harry Lass (1963),[13] who also gave the corresponding coordinates (2g) which are sometimes called "Lass coordinates".[9] Metric (2h), as well as (2a, 2b), was also derived by Fritz Rohrlich (1963).[12] Eventually, the Lass coordinates (2g, 2h) were identified with Radar coordinates by Desloge & Philpott (1987).[29][8]

Table with historical formulas edit

Einstein (1907)[H 30]
 
Born (1909)[H 14]
 
Herglotz (1909)[H 31][30]
 
Sommerfeld (1910)[H 15]
 
von Laue (1911)[H 32]
 
Einstein (1912)[H 17]
 
Kottler (1912)[H 33]
 
Lorentz (1913)[H 18]
 
Kottler (1914a)[H 34]
 
Kottler (1914b)[H 35]
 
Kottler (1916, 1918)[H 36]
 
Pauli (1921)[H 37]
 
Bollert (1922)[H 20]
 
Mohorovičić (1922, 1923); Bollert (1922b)[H 28]
 
Lemaître (1924)[H 21]
 
Einstein & Rosen (1935)[H 22]
 
Møller (1952)[H 23]
 

See also edit

References edit

  1. ^ a b Øyvind Grøn (2010). Lecture Notes on the General Theory of Relativity. Lecture Notes in Physics. Vol. 772. Springer. pp. 86–91. ISBN 978-0387881348.
  2. ^ a b Misner, C. W.; Thorne, K. S.; Wheeler, J. A. (1973). Gravitation. Freeman. ISBN 0716703440.
  3. ^ Kopeikin, S., Efroimsky, M., Kaplan, G. (2011). Relativistic Celestial Mechanics of the Solar System. John Wiley & Sons. ISBN 978-3527408566.{{cite book}}: CS1 maint: multiple names: authors list (link)
  4. ^ Padmanabhan, T. (2010). Gravitation: Foundations and Frontiers. Cambridge University Press. ISBN 978-1139485395.
  5. ^ N. D. Birrell, P. C. W. Davies (1982). Quantum Fields in Curved Space. Cambridge University Press. ISBN 1107392810.
  6. ^ a b Leonard Susskind, James Lindesay (2005). An Introduction to Black Holes, Information and the String Theory Revolution: The Holographic Universe. World Scientific. pp. 8–10. ISBN 9812561315.
  7. ^ Muñoz, Gerardo; Jones, Preston (2010). "The equivalence principle, uniformly accelerated reference frames, and the uniform gravitational field". American Journal of Physics. 78 (4): 377–383. arXiv:1003.3022. Bibcode:2010AmJPh..78..377M. doi:10.1119/1.3272719. S2CID 118616525.
  8. ^ a b c Minguzzi, E. (2005). "The Minkowski metric in non-inertial observer radar coordinates". American Journal of Physics. 73 (12): 1117–1121. arXiv:physics/0412024. Bibcode:2005AmJPh..73.1117M. doi:10.1119/1.2060716. S2CID 119359878.
  9. ^ a b c d David Tilbrook (1997). "General Coordinatisations of the Flat Space-Time of Constant Proper-acceleration". Australian Journal of Physics. 50 (5): 851–868. doi:10.1071/P96111.
  10. ^ Jones, Preston; Wanex, Lucas F. (2006). "The Clock Paradox in a Static Homogeneous Gravitational Field". Foundations of Physics Letters. 19 (1): 75–85. arXiv:physics/0604025. Bibcode:2006FoPhL..19...75J. doi:10.1007/s10702-006-1850-3. S2CID 14583590.
  11. ^ For instance, Birrell & Davies (1982), pp. 110–111 or Padmanabhan (2010), p. 126 denote equations (2g, 2h) as Rindler coordinates or Rindler frame; Tilbrook (1997) pp. 864–864 or Jones & Wanex (2006) denote equations (2a, 2b) as Rindler coordinates
  12. ^ a b Rohrlich, Fritz (1963). "The principle of equivalence". Annals of Physics. 22 (2): 169–191. Bibcode:1963AnPhy..22..169R. doi:10.1016/0003-4916(63)90051-4.
  13. ^ a b Harry Lass (1963). "Accelerating Frames of Reference and the Clock Paradox". American Journal of Physics. 31 (4): 274–276. Bibcode:1963AmJPh..31..274L. doi:10.1119/1.1969430. S2CID 121608504.
  14. ^ a b Rindler, W. (1960). "Hyperbolic Motion in Curved Space Time". Physical Review. 119 (6): 2082–2089. Bibcode:1960PhRv..119.2082R. doi:10.1103/PhysRev.119.2082.
  15. ^ a b c Rindler, W. (1966). "Kruskal Space and the Uniformly Accelerated Frame". American Journal of Physics. 34 (12): 1174–1178. Bibcode:1966AmJPh..34.1174R. doi:10.1119/1.1972547.
  16. ^ a b Pauli, Wolfgang (1921), "Die Relativitätstheorie", Encyclopädie der Mathematischen Wissenschaften, 5 (2): 539–776
    In English: Pauli, W. (1981) [1921]. Theory of Relativity. Vol. 165. Dover Publications. ISBN 0-486-64152-X. {{cite book}}: |journal= ignored (help)
  17. ^ von Laue, M. (1921). Die Relativitätstheorie, Band 1 (fourth edition of "Das Relativitätsprinzip" ed.). Vieweg.; First edition 1911, second expanded edition 1913, third expanded edition 1919.
  18. ^ a b Don Koks (2006). Explorations in Mathematical Physics. Springer. pp. 235–269. ISBN 0387309438.
  19. ^ Møller, C. (1955) [1952]. The theory of relativity. Oxford Clarendon Press.
  20. ^ Møller (1952), eq. 154
  21. ^ Misner & Thorne & Wheeler (1973), section 6.6
  22. ^ Muñoz & Jones (2010), eq. 37, 38
  23. ^ Pauli (1921), section 32-y
  24. ^ Rindler (1966), p. 1177
  25. ^
rindler, coordinates, coordinate, system, used, context, special, relativity, describe, hyperbolic, acceleration, uniformly, accelerating, reference, frame, flat, spacetime, relativistic, physics, coordinates, hyperbolically, accelerated, reference, frame, con. Rindler coordinates are a coordinate system used in the context of special relativity to describe the hyperbolic acceleration of a uniformly accelerating reference frame in flat spacetime In relativistic physics the coordinates of a hyperbolically accelerated reference frame H 1 1 constitute an important and useful coordinate chart representing part of flat Minkowski spacetime 2 3 4 5 In special relativity a uniformly accelerating particle undergoes hyperbolic motion for which a uniformly accelerating 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 and Proper reference frame flat spacetime In this article the speed of light is defined by c 1 the inertial coordinates are X Y Z T and the hyperbolic coordinates are x y z t These hyperbolic coordinates can be separated into two main variants depending on the accelerated observer s position If the observer is located at time T 0 at position X 1 a with a as the constant proper acceleration measured by a comoving accelerometer then the hyperbolic coordinates are often called Rindler coordinates with the corresponding Rindler metric 6 If the observer is located at time T 0 at position X 0 then the hyperbolic coordinates are sometimes called Moller coordinates 1 or Kottler Moller coordinates with the corresponding Kottler Moller metric 7 An alternative chart often related to observers in hyperbolic motion is obtained using Radar coordinates 8 which are sometimes called Lass coordinates 9 10 Both the Kottler Moller coordinates as well as Lass coordinates are denoted as Rindler coordinates as well 11 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 1907 1912 H 2 Max Born 1909 H 1 Arnold Sommerfeld 1910 H 3 Max von Laue 1911 H 4 Hendrik Lorentz 1913 H 5 Friedrich Kottler 1914 H 6 Wolfgang Pauli 1921 H 7 Karl Bollert 1922 H 8 Stjepan Mohorovicic 1922 H 9 Georges Lemaitre 1924 H 10 Einstein amp Nathan Rosen 1935 H 2 Christian Moller 1943 1952 H 11 Fritz Rohrlich 1963 12 Harry Lass 1963 13 and in relation to both flat and curved spacetime of general relativity by Wolfgang Rindler 1960 1966 14 15 For details and sources see History Contents 1 Characteristics of the Rindler frame 2 Variants of transformation formulas 3 The Rindler observers 4 A paradoxical property 5 Minkowski observers 6 The Rindler horizon 7 Geodesics 8 The Fermat metric 9 Symmetries 10 Notions of distance 11 Generalization to curved spacetimes 12 History 12 1 Overview 12 2 Table with historical formulas 13 See also 14 References 15 Historical sources 16 Further readingCharacteristics of the Rindler frame edit nbsp Rindler chart for a 0 5 displaystyle alpha 0 5 nbsp in equation 1a plotted on a Minkowski diagram The dashed lines are the Rindler horizons The worldline of a body in hyperbolic motion having constant proper acceleration a displaystyle alpha nbsp in the X displaystyle X nbsp direction as a function of proper time t displaystyle tau nbsp and rapidity a t displaystyle alpha tau nbsp can be given by 16 T x sinh a t X x cosh a t displaystyle T x sinh alpha tau quad X x cosh alpha tau nbsp where x 1 a displaystyle x 1 alpha nbsp is constant and a t displaystyle alpha tau nbsp is variable with the worldline resembling the hyperbola X 2 T 2 x 2 displaystyle X 2 T 2 x 2 nbsp Sommerfeld H 3 17 showed that the equations can be reinterpreted by defining x displaystyle x nbsp as variable and a t displaystyle alpha tau nbsp 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 t t displaystyle tau t nbsp the transformation formulas between the inertial coordinates and the hyperbolic coordinates are consequently 6 9 T x sinh a t X x cosh a t Y y Z z displaystyle T x sinh alpha t quad X x cosh alpha t quad Y y quad Z z nbsp 1a with the inverse t 1 a artanh T X x X 2 T 2 y Y z Z displaystyle t frac 1 alpha operatorname artanh left frac T X right quad x sqrt X 2 T 2 quad y Y quad z Z nbsp Differentiated and inserted into the Minkowski metric d s 2 d T 2 d X 2 d Y 2 d Z 2 displaystyle mathrm d s 2 mathrm d T 2 mathrm d X 2 mathrm d Y 2 mathrm d Z 2 nbsp the metric in the hyperbolically accelerated frame follows as d s 2 a x 2 d t 2 d x 2 d y 2 d z 2 displaystyle mathrm d s 2 alpha x 2 mathrm d t 2 mathrm d x 2 mathrm d y 2 mathrm d z 2 nbsp 1b 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 lt X lt X lt T lt X displaystyle 0 lt X lt infty X lt T lt X nbsp which is often called the Rindler wedge if a displaystyle alpha nbsp represents the proper acceleration along the hyperbola x 1 a displaystyle x 1 alpha nbsp 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 displaystyle x 0 nbsp the Rindler horizon having greater proper acceleration All the Rindler observers are instantaneously at rest at time T 0 displaystyle T 0 nbsp in the inertial frame and at this time a Rindler observer with proper acceleration a i displaystyle alpha i nbsp will be at position X 1 a i displaystyle X 1 alpha i nbsp really X c 2 a i displaystyle X c 2 alpha i nbsp but we assume units where c 1 displaystyle c 1 nbsp 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 displaystyle T 0 nbsp then when defining a Rindler coordinate system we have a choice of which Rindler observer s proper time will be equal to the coordinate time t displaystyle t nbsp in Rindler coordinates and this observer s proper acceleration defines the value of a displaystyle alpha nbsp 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 18 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 a 1 displaystyle alpha 1 nbsp so that a displaystyle alpha nbsp can be eliminated from the equations The above equation has been simplified for c 1 displaystyle c 1 nbsp The unsimplified equation is more convenient for finding the Rindler Horizon distance given an acceleration a displaystyle alpha nbsp t c a artanh c T X X c T c 2 T a X X c 2 T a t T t c 2 a displaystyle begin aligned amp t frac c alpha operatorname artanh left frac cT X right overset X gg cT approx frac c 2 T alpha X amp Rightarrow X approx frac c 2 T alpha t overset T approx t approx frac c 2 alpha end aligned nbsp The remainder of the article will follow the convention of setting both a 1 displaystyle alpha 1 nbsp and c 1 displaystyle c 1 nbsp so units for X displaystyle X nbsp and x displaystyle x nbsp will be 1 unit c 2 a 1 displaystyle c 2 alpha 1 nbsp Be mindful that setting a 1 displaystyle alpha 1 nbsp light second second2 is very different from setting a 1 displaystyle alpha 1 nbsp light year year2 Even if we pick units where c 1 displaystyle c 1 nbsp the magnitude of the proper acceleration a displaystyle alpha nbsp will depend on our choice of units for example if we use units of light years for distance X displaystyle X nbsp or x displaystyle x nbsp and years for time T displaystyle T nbsp or t displaystyle t nbsp this would mean a 1 displaystyle alpha 1 nbsp light year year2 equal to about 9 5 meters second2 while if we use units of light seconds for distance X displaystyle X nbsp or x displaystyle x nbsp and seconds for time T displaystyle T nbsp or t displaystyle t nbsp this would mean a 1 displaystyle alpha 1 nbsp light second second2 or 299 792 458 meters second2 Variants of transformation formulas editA 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 19 Depending on the choice of origin of these coordinates one can derive the metric the time dilation between the time at the origin d t 0 displaystyle dt 0 nbsp and d t displaystyle dt nbsp at point x displaystyle x nbsp and the coordinate light speed d x d t displaystyle dx dt nbsp this variable speed of light 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 displaystyle c 1 nbsp in inertial frames X displaystyle X nbsp at T 0 displaystyle T 0 nbsp Transformation Metric Time dilation and Coordinate speed of light X 0 displaystyle X 0 nbsp Kottler Moller coordinates H 12 20 21 22 T x 1 a sinh a t X x 1 a cosh a t 1 a Y y Z z t 1 a artanh T X 1 a x X 1 a 2 T 2 1 a y Y z Z displaystyle begin array c c begin aligned T amp left x frac 1 alpha right sinh alpha t X amp left x frac 1 alpha right cosh alpha t frac 1 alpha Y amp y Z amp z end aligned amp begin aligned t amp frac 1 alpha operatorname artanh left frac T X frac 1 alpha right x amp sqrt left X frac 1 alpha right 2 T 2 frac 1 alpha y amp Y z amp Z end aligned end array nbsp 2a d s 2 1 a x 2 d t 2 d x 2 d y 2 d z 2 displaystyle ds 2 1 alpha x 2 dt 2 dx 2 dy 2 dz 2 nbsp 2b d t 1 a x d t 0 d x d t 1 a x displaystyle dt 1 alpha x dt 0 qquad frac dx dt 1 alpha x nbsp 2c Rindler coordinates 23 24 18 X 1 a displaystyle X frac 1 alpha nbsp T x sinh a t X x cosh a t Y y Z z t 1 a artanh T X x X 2 T 2 y Y z Z displaystyle begin array c c begin aligned T amp x sinh alpha t X amp x cosh alpha t Y amp y Z amp z end aligned amp begin aligned t amp frac 1 alpha operatorname artanh frac T X x amp sqrt X 2 T 2 y amp Y z amp Z end aligned end array nbsp 2d d s 2 a x 2 d t 2 d x 2 d y 2 d z 2 displaystyle ds 2 alpha x 2 dt 2 dx 2 dy 2 dz 2 nbsp 2e d t a x d t 0 d x d t a x displaystyle dt alpha x dt 0 qquad frac dx dt alpha x nbsp 2f Radar coordinates Lass coordinates 25 26 8 9 X 1 a displaystyle X frac 1 alpha nbsp T 1 a e a x sinh a t X 1 a e a x cosh a t Y y Z z t 1 a artanh T X x 1 2 a ln a 2 X 2 T 2 y Y z Z displaystyle begin array c c begin aligned T amp frac 1 alpha e alpha x sinh alpha t X amp frac 1 alpha e alpha x cosh alpha t Y amp y Z amp z end aligned amp begin aligned t amp frac 1 alpha operatorname artanh frac T X x amp frac 1 2 alpha ln left alpha 2 left X 2 T 2 right right y amp Y z amp Z end aligned end array nbsp 2g d s 2 e 2 a x d t 2 d x 2 d y 2 d z 2 displaystyle ds 2 e 2 alpha x left dt 2 dx 2 right dy 2 dz 2 nbsp 2h d t e a x d t 0 d x d t 1 displaystyle dt e alpha x dt 0 qquad frac dx dt 1 nbsp 2i The Rindler observers editIn the new chart 1a with c 1 displaystyle c 1 nbsp and a 1 displaystyle alpha 1 nbsp it is natural to take the coframe field d s 0 x d t d s 1 d x d s 2 d y d s 3 d z displaystyle d sigma 0 x dt d sigma 1 dx d sigma 2 dy d sigma 3 dz nbsp which has the dual frame field e 0 1 x t e 1 x e 2 y e 3 z displaystyle vec e 0 frac 1 x partial t vec e 1 partial x vec e 2 partial y vec e 3 partial z nbsp This defines a local Lorentz frame in the tangent space at each event in the region covered by our Rindler chart namely the Rindler wedge The integral curves of the timelike unit vector field e 0 displaystyle vec e 0 nbsp 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 displaystyle x x 0 y y 0 z z 0 nbsp Using the coordinate transformation above we find that these correspond to hyperbolic arcs in the original Cartesian chart nbsp Some representative Rindler observers navy blue hyperbolic arcs depicted using the Cartesian chart The red lines at 45 degrees from the vertical represent the Rindler horizon the Rindler coordinate system is only defined to the right of this boundary As with any timelike congruence in any Lorentzian manifold this congruence has a kinematic decomposition see Raychaudhuri equation 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 e 0 e 0 1 x e 1 displaystyle nabla vec e 0 vec e 0 frac 1 x vec e 1 nbsp That is each Rindler observer is accelerating in the x displaystyle partial x nbsp 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 displaystyle t t 0 nbsp these appear as horizontal half planes in the Rindler chart and as half planes through T X 0 displaystyle T X 0 nbsp in the Cartesian chart see the figure above Setting d t 0 displaystyle dt 0 nbsp in the line element we see that these have the ordinary Euclidean geometry d s 2 d x 2 d y 2 d z 2 x gt 0 y z displaystyle d sigma 2 dx 2 dy 2 dz 2 forall x gt 0 forall y z nbsp 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 editNote 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 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 edit nbsp A representative Minkowski observer navy blue hyperbolic secant curve depicted using the Rindler chart The Rindler horizon is shown in red It is worthwhile to also introduce an alternative frame given in the Minkowski chart by the natural choice f 0 T f 1 X f 2 Y f 3 Z displaystyle vec f 0 partial T vec f 1 partial X vec f 2 partial Y vec f 3 partial Z nbsp Transforming these vector fields using the coordinate transformation given above we find that in the Rindler chart in the Rinder wedge this frame becomes f 0 1 x cosh t t sinh t x f 1 1 x sinh t t cosh t x f 2 y f 3 z displaystyle begin aligned vec f 0 amp frac 1 x cosh t partial t sinh t partial x vec f 1 amp frac 1 x sinh t partial t cosh t partial x vec f 2 amp partial y vec f 3 partial z end aligned nbsp Computing the kinematic decomposition of the timelike congruence defined by the timelike unit vector field f 0 displaystyle vec f 0 nbsp we find that the expansion and vorticity again vanishes and in addition the acceleration vector vanishes f 0 f 0 0 displaystyle nabla vec f 0 vec f 0 0 nbsp 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 displaystyle x 0 nbsp 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 displaystyle t t 0 x x 0 y y 0 z z 0 nbsp is t artanh s x 0 x 0 lt s lt x 0 x x 0 2 s 2 x 0 lt s lt x 0 y y 0 z z 0 displaystyle begin aligned t amp operatorname artanh left frac s x 0 right x 0 lt s lt x 0 x amp sqrt x 0 2 s 2 x 0 lt s lt x 0 y amp y 0 z amp z 0 end aligned nbsp where s displaystyle s nbsp 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 displaystyle x 0 1 nbsp and we have drawn correctly scaled and boosted the light cones at s 1 2 0 1 2 displaystyle s in left frac 1 2 0 frac 1 2 right nbsp The Rindler horizon editThe Rindler coordinate chart has a coordinate singularity at x 0 where the metric tensor expressed in the Rindler coordinates has vanishing determinant 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 T2 X2 X gt 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 of a black hole each observer hovering at a constant radius in Schwarzschild coordinates 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 editThe geodesic equations in the Rindler chart are easily obtained from the geodesic Lagrangian they are t 2 x x t 0 x x t 2 0 y 0 z 0 displaystyle ddot t frac 2 x dot x dot t 0 ddot x x dot t 2 0 ddot y 0 ddot z 0 nbsp 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 nbsp Some representative null geodesics black hyperbolic semicircular arcs projected into the spatial hyperslice t 0 of the Rindler observers The Rindler horizon is shown as a magenta plane From the first third and fourth we immediately obtain the first integrals t E x 2 y P z Q displaystyle dot t frac E x 2 dot y P dot z Q nbsp But from the line element we have e x 2 t 2 x 2 y 2 z 2 displaystyle varepsilon x 2 dot t 2 dot x 2 dot y 2 dot z 2 nbsp where e 1 0 1 displaystyle varepsilon in left 1 0 1 right nbsp for timelike null and spacelike geodesics respectively This gives the fourth first integral namely x 2 e E 2 x 2 P 2 Q 2 displaystyle dot x 2 left varepsilon frac E 2 x 2 right P 2 Q 2 nbsp This suffices to give the complete solution of the geodesic equations In the case of null geodesics from E 2 x 2 P 2 Q 2 displaystyle frac E 2 x 2 P 2 Q 2 nbsp with nonzero E displaystyle E nbsp we see that the x coordinate ranges over the interval 0 lt x lt E P 2 Q 2 displaystyle 0 lt x lt frac E sqrt P 2 Q 2 nbsp The complete seven parameter family giving any null geodesic through any event in the Rindler wedge is t t 0 artanh 1 E s P 2 Q 2 E 2 P 2 Q 2 x 0 2 artanh 1 E E 2 P 2 Q 2 x 0 2 x x 0 2 2 s E 2 P 2 Q 2 x 0 2 s 2 P 2 Q 2 y y 0 P s z z 0 Q s displaystyle begin aligned t t 0 amp operatorname artanh left frac 1 E left s left P 2 Q 2 right sqrt E 2 left P 2 Q 2 right x 0 2 right right amp qquad operatorname artanh left frac 1 E sqrt E 2 P 2 Q 2 x 0 2 right x amp sqrt x 0 2 2s sqrt E 2 P 2 Q 2 x 0 2 s 2 P 2 Q 2 y y 0 amp Ps z z 0 Qs end aligned nbsp Plotting the tracks of some representative null geodesics through a given event that is projecting to the hyperslice t 0 displaystyle t 0 nbsp 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 27 The Fermat metric editThe 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 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 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 d s 2 g 00 d t 2 g j k d x j d x k j k 1 2 3 displaystyle ds 2 g 00 dt 2 g jk dx j dx k j k in 1 2 3 nbsp the Fermat metric on t 0 displaystyle t 0 nbsp is simply d r 2 1 g 00 g j k d x j d x k displaystyle d rho 2 frac 1 g 00 left g jk dx j dx k right nbsp where the metric coeffients are understood to be evaluated at t 0 displaystyle t 0 nbsp In the Rindler chart the timelike translation t displaystyle partial t nbsp 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 Therefore we may immediately write down the Fermat metric for the Rindler observers d r 2 1 x 2 d x 2 d y 2 d z 2 x gt 0 y z displaystyle d rho 2 frac 1 x 2 left dx 2 dy 2 dz 2 right forall x gt 0 forall y z nbsp 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 editSince 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 three spatials three rotations and three boosts Together these generate the proper isochronous Poincare 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 t y z z y y z displaystyle partial t partial y partial z z partial y y partial z nbsp time translation spatial translations orthogonal to the direction of acceleration and spatial rotation orthogonal to the direction of acceleration plus six more exp t y x t y x x y exp t z x t z x x z exp t 1 x t x displaystyle begin aligned amp exp pm t left frac y x partial t pm left y partial x x partial y right right amp exp pm t left frac z x partial t pm left z partial x x partial z right right amp exp pm t left frac 1 x partial t pm partial x right end aligned nbsp 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 T displaystyle partial T nbsp 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 manifolds Notions of distance editOne 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 which can be used by the Rindler observers nbsp Operational meaning of the radar distance between two Rindler observers navy blue vertical lines The Rindler horizon is shown at left red vertical line The world line of the radar pulse is also depicted together with the properly scaled light cones at events A B C The first is the one we have tacitly employed above the induced Riemannian metric on the spatial hyperslices t t 0 displaystyle t t 0 nbsp 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 which 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 displaystyle x x 0 y 0 z 0 nbsp and x x 0 h y 0 z 0 displaystyle x x 0 h y 0 z 0 nbsp 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 d y d z 0 displaystyle dy dz 0 nbsp in the Rindler line element we readily obtain the equation of null geodesics moving in the direction of acceleration t t 0 log x x 0 displaystyle t t 0 log left frac x x 0 right nbsp Therefore the radar distance between these two observers is given by x 0 log 1 h x 0 h h 2 2 x 0 O h 3 displaystyle x 0 log left 1 frac h x 0 right h frac h 2 2 x 0 O left h 3 right nbsp 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 1 x 0 O h 3 textstyle h frac 1 x 0 O left h 3 right nbsp Therefore in the case of a trailing observer estimating distance to a leading observer the case h gt 0 displaystyle h gt 0 nbsp 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 editRindler 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 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 editOverview edit Kottler Moller and Rindler coordinates Albert Einstein 1907 H 13 studied the effects within a uniformly accelerated frame obtaining equations for coordinate dependent time dilation and speed of light equivalent to 2c and in order to make the formulas independent of the observer s origin he obtained time dilation 2i in formal agreement with Radar coordinates While introducing the concept of Born rigidity Max Born 1909 H 14 noted that the formulas for hyperbolic motion can be used as transformations into a hyperbolically accelerated reference system German hyperbolisch beschleunigtes Bezugsystem equivalent to 2d Born s work was further elaborated by Arnold Sommerfeld 1910 H 15 and Max von Laue 1911 H 16 who both obtained 2d using imaginary numbers which was summarized by Wolfgang Pauli 1921 16 who besides coordinates 2d also obtained metric 2e using imaginary numbers Einstein 1912 H 17 studied a static gravitational field and obtained the Kottler Moller metric 2b as well as approximations to formulas 2a using a coordinate dependent speed of light 28 Hendrik Lorentz 1913 H 18 obtained coordinates similar to 2d 2e 2f while studying Einstein s equivalence principle and the uniform gravitational field A detailed description was given by Friedrich Kottler 1914 H 19 who formulated the corresponding orthonormal tetrad transformation formulas and metric 2a 2b Also Karl Bollert 1922 H 20 obtained the metric 2b in his study of uniform acceleration and uniform gravitational fields In a paper concerned with Born rigidity Georges Lemaitre 1924 H 21 obtained coordinates and metric 2a 2b Albert Einstein and Nathan Rosen 1935 described 2d 2e as the well known expressions for a homogeneous gravitational field H 22 After Christian Moller 1943 H 11 obtained 2a 2b in as study related to homogeneous gravitational fields he 1952 H 23 as well as Misner amp Thorne amp Wheeler 1973 2 used Fermi Walker transport to obtain the same equations While these investigations were concerned with flat spacetime Wolfgang Rindler 1960 14 analyzed hyperbolic motion in curved spacetime and showed 1966 15 the analogy between the hyperbolic coordinates 2d 2e 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 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 X 2 T 2 gt 0 displaystyle X left X 2 T 2 right gt 0 nbsp H 24 Sommerfeld 1910 defined that the coordinates allowed for the transformation between inertial and hyperbolic coordinates must satisfy T lt X displaystyle T lt X nbsp H 25 Kottler 1914 H 26 defined this region as X 2 T 2 gt 0 displaystyle X 2 T 2 gt 0 nbsp and pointed out the existence of a border plane German Grenzebene c 2 a x displaystyle c 2 alpha x nbsp 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 H 27 Rindler 1966 15 demonstrated the relation between such a horizon and the horizon in Kruskal coordinates Radar coordinates Using Bollert s formalism Stjepan Mohorovicic 1922 H 28 made a different choice for some parameter and obtained metric 2h with a printing error which was corrected by Bollert 1922b with another printing error until a version without printing error was given by Mohorovicic 1923 In addition Mohorovicic erroneously argued that metric 2b now called Kottler Moller metric is incorrect which was rebutted by Bollert 1922 H 29 Metric 2h was rediscovered by Harry Lass 1963 13 who also gave the corresponding coordinates 2g which are sometimes called Lass coordinates 9 Metric 2h as well as 2a 2b was also derived by Fritz Rohrlich 1963 12 Eventually the Lass coordinates 2g 2h were identified with Radar coordinates by Desloge amp Philpott 1987 29 8 Table with historical formulas edit Einstein 1907 H 30 s t 1 g 3 c 2 s t e g 3 c 2 c 1 g 3 c 2 displaystyle 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 nbsp Born 1909 H 14 x q 3 y h z z t p c 2 3 p x t q t t 1 p 2 c 2 x 2 c 2 t 2 3 2 displaystyle 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 nbsp Herglotz 1909 H 31 30 x x y y t z t z e ϑ t z t z e ϑ x x 0 y y 0 z z 0 2 t 2 displaystyle scriptstyle begin matrix begin aligned x amp x y amp y t z amp t z e vartheta t z amp t z e vartheta end aligned boldsymbol downarrow x x 0 quad y y 0 quad z sqrt z 0 2 t 2 end matrix nbsp Sommerfeld 1910 H 15 x r cos f y y z z l r sin f f i ps l i c t displaystyle scriptstyle begin aligned x amp r cos varphi y amp y z amp z l amp r sin varphi varphi amp i psi l ict end aligned nbsp von Laue 1911 H 32 X R cos f L R sin f R 2 X 2 L 2 tan f L X displaystyle scriptstyle begin aligned X amp R cos varphi L amp R sin varphi R 2 amp X 2 L 2 tan varphi amp frac L X end aligned nbsp Einstein 1912 H 17 d 3 2 d t 2 d x 2 c 2 d t 2 c c 0 a x 3 x a c 2 t 2 h y z z t c t displaystyle scriptstyle begin matrix d xi 2 d tau 2 dx 2 c 2 dt 2 boldsymbol downarrow c c 0 ax boldsymbol downarrow begin aligned xi amp x frac ac 2 t 2 eta amp y zeta amp z tau amp ct end aligned end matrix nbsp Kottler 1912 H 33 x 1 x 0 1 x 2 x 0 2 x 3 b cos i f x 4 b sin i f displaystyle scriptstyle begin aligned x 1 amp x 0 1 x 2 amp x 0 2 x 3 amp b cos i varphi x 4 amp b sin i varphi end aligned nbsp Lorentz 1913 H 18 d c g c d z z a z z 0 c t b z z 0 a 1 2 e k t e k t b 1 2 e k t e k t c k z z 0 z z 0 c 2 g d x 2 d y 2 d z 2 c 2 d t d x 2 d y 2 d z 2 c 2 d t 2 displaystyle scriptstyle begin matrix dc frac g c dz hline begin aligned z amp a left z z 0 prime right ct amp b left z z 0 prime right a amp frac 1 2 left e kt e kt right b amp frac 1 2 left e kt e kt right end aligned boldsymbol downarrow c k left z z 0 prime right z z 0 prime frac c 2 g boldsymbol downarrow begin aligned amp dx 2 dy 2 dz 2 c 2 dt amp dx prime 2 dy prime 2 dz prime 2 c prime 2 dt prime 2 end aligned end matrix nbsp Kottler 1914a H 34 x 1 x 0 1 x 2 x 0 2 x 3 b cos i u x 4 b sin i u d s 2 c 2 d t 2 b 2 d u 2 c 1 1 0 c 1 2 0 c 1 3 sin i u c 1 4 cos i u c 2 1 0 c 2 2 0 c 2 3 cos i u c 2 4 sin i u d S 2 d X 2 d Y 2 d Z 2 c Z c b 2 d T c c Z c 2 b 1 c displaystyle scriptstyle begin matrix begin aligned x 1 amp x 0 1 x 2 amp x 0 2 x 3 amp b cos iu x 4 amp b sin iu end aligned boldsymbol downarrow ds 2 c 2 d tau 2 b 2 du 2 boldsymbol downarrow begin matrix c 1 1 0 amp amp c 1 2 0 amp amp c 1 3 sin iu amp amp c 1 4 cos iu c 2 1 0 amp amp c 2 2 0 amp amp c 2 3 cos iu amp amp 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 nbsp Kottler 1914b H 35 c 1 1 0 c 1 2 0 c 1 3 1 i sinh u c 1 4 cosh u c 2 1 0 c 2 2 0 c 2 3 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 X x D 2 c 2 D 3 c 3 D 4 c 4 X x 0 X Y y 0 Y Z b Z cosh u c T b Z sinh u D 2 X D 3 Y D 4 Z X X 0 x 0 q x T Y Y 0 y 0 q y T b Z Z 0 q x T 2 c 2 T 2 c T b artanh c T Z 0 q x T X X 0 q x T Y Y 0 q y T Z Z 0 q x T d S 2 d X 2 d Y 2 d Z 2 c 2 b Z b 2 2 d T 2 displaystyle scriptstyle begin matrix begin matrix c 1 1 0 amp amp c 1 2 0 amp amp c 1 3 frac 1 i sinh u amp amp c 1 4 cosh u c 2 1 0 amp amp c 2 2 0 amp amp c 2 3 frac 1 i cosh u amp amp c 2 4 sinh u c 3 1 1 amp amp c 3 2 0 amp amp c 3 3 0 amp amp c 3 4 0 c 4 1 0 amp amp c 4 2 1 amp amp c 4 3 0 amp amp 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 aligned X amp x 0 mathfrak X Y amp y 0 mathfrak Y Z amp left b mathfrak Z right cosh mathfrak u cT amp left b mathfrak Z right sinh mathfrak u end aligned left Delta 2 mathfrak X Delta 3 mathfrak Y Delta 4 mathfrak Z right boldsymbol downarrow begin aligned mathfrak X amp X 0 x 0 q x T mathfrak Y amp Y 0 y 0 q y T b mathfrak Z amp sqrt left Z 0 q x T right 2 c 2 T 2 c mathfrak T amp b operatorname artanh frac cT Z 0 q x T end aligned 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 nbsp Kottler 1916 1918 H 36 x x y y c 2 g z c 2 g z cosh g t c c t c 2 g z sinh g t c d s 2 d x 2 d y 2 d z 2 c g c z 2 d t 2 displaystyle scriptstyle begin matrix begin aligned x amp x y amp y frac c 2 gamma z amp left frac c 2 gamma z right cosh frac gamma t c ct amp left frac c 2 gamma z right sinh frac gamma t c end aligned boldsymbol downarrow ds 2 dx prime 2 dy prime 2 dz prime 2 left c frac gamma c z right 2 dt prime 2 end matrix nbsp Pauli 1921 H 37 x 1 ϱ cos f x 4 ϱ sin f d s 2 d 3 1 2 d 3 2 2 d 3 3 2 3 1 2 d 3 4 2 3 1 ϱ 3 2 x 2 3 3 x 3 3 4 f displaystyle scriptstyle begin matrix begin aligned x 1 amp varrho cos varphi x 4 amp varrho sin varphi end aligned 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 nbsp Bollert 1922 H 20 d s 2 c 2 1 g 0 x c 2 d t 2 d x 2 d y 2 d z 2 d s 2 g 44 d x 4 2 g 11 d x 1 2 g 22 d x 2 2 d x 3 2 V g 11 2 g 11 V 0 g 22 1 g 11 1 V 0 V a x b d s 2 d x 4 2 a x b 2 d x 2 d y 2 d z 2 displaystyle 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 11 V 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 nbsp Mohorovicic 1922 1923 Bollert 1922b H 28 Mohorovicic 1922 g 11 g 44 V 2 V V V 2 0 V x 1 e a x 1 d s 2 e 2 a d x 4 2 d x 1 2 d x 2 2 d x 3 2 corrected by Bollert 1922b d s 2 e 2 a x d x 4 2 d x 1 2 d x 2 2 d x 3 2 final correction by Mohorovicic 1923 d s 2 e 2 a x 1 d x 4 2 d x 1 2 d x 2 2 d x 3 2 displaystyle scriptstyle begin matrix text Mohorovicic 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 Mohorovicic 1923 ds 2 e 2ax 1 left dx 4 2 dx 1 2 right dx 2 2 dx 3 2 end matrix nbsp Lemaitre 1924 H 21 1 g 3 1 g x cosh g t g t 1 g x sinh g t d s 2 d x 2 d y 2 d z 2 1 g x 2 d t 2 displaystyle scriptstyle begin matrix begin aligned 1 g xi amp 1 gx cosh gt g tau amp 1 gx sinh gt end aligned boldsymbol downarrow ds 2 dx 2 dy 2 dz 2 1 gx 2 dt 2 end matrix nbsp Einstein amp Rosen 1935 H 22 3 1 x 1 cosh a x 4 3 2 x 2 3 3 x 3 3 4 x 1 sinh a x 4 d s 2 d x 1 2 d x 2 2 d x 3 2 a 2 x 1 2 d x 4 2 displaystyle scriptstyle begin matrix begin aligned xi 1 amp x 1 cosh alpha x 4 xi 2 amp x 2 xi 3 amp x 3 xi 4 amp x 1 sinh alpha x 4 end aligned boldsymbol downarrow ds 2 dx 1 2 dx 2 2 dx 3 2 alpha 2 x 1 2 dx 4 2 end matrix nbsp Moller 1952 H 23 a i k U 4 i c 0 0 i U 1 c 0 1 0 0 0 0 1 0 U 1 i c 0 0 U 4 i c U i c sinh g t c 0 0 i g cosh g t c X i f i t x k a k i t X c 2 g cosh g t c 1 x cosh g t c Y y Z z T c g sinh g t c x sinh g t c c d s 2 d x 2 d y 2 d z 2 c 2 d t 2 1 g x c 2 2 displaystyle scriptstyle begin matrix alpha ik left begin matrix U 4 ic amp 0 amp 0 amp iU 1 c 0 amp 1 amp 0 amp 0 0 amp 0 amp 1 amp 0 U 1 ic amp 0 amp 0 amp 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 aligned X amp frac c 2 g left cosh frac gt c 1 right x cosh frac gt c Y amp y Z amp z T amp frac c g sinh frac gt c x frac sinh frac gt c c end aligned boldsymbol downarrow ds 2 dx 2 dy 2 dz 2 c 2 dt 2 left 1 gx c 2 right 2 end matrix nbsp See also editBell s spaceship paradox 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 Congruence general relativity Ehrenfest paradox for a sometimes controversial subject often studied using Born coordinates Frame fields in general relativity General relativity resources Milne model Raychaudhuri equation Unruh effectReferences edit a b Oyvind Gron 2010 Lecture Notes on the General Theory of Relativity Lecture Notes in Physics Vol 772 Springer pp 86 91 ISBN 978 0387881348 a b Misner C W Thorne K S Wheeler J A 1973 Gravitation Freeman ISBN 0716703440 Kopeikin S Efroimsky M Kaplan G 2011 Relativistic Celestial Mechanics of the Solar System John Wiley amp Sons ISBN 978 3527408566 a href Template Cite book html title Template Cite book cite book a CS1 maint multiple names authors list link Padmanabhan T 2010 Gravitation Foundations and Frontiers Cambridge University Press ISBN 978 1139485395 N D Birrell P C W Davies 1982 Quantum Fields in Curved Space Cambridge University Press ISBN 1107392810 a b Leonard Susskind James Lindesay 2005 An Introduction to Black Holes Information and the String Theory Revolution The Holographic Universe World Scientific pp 8 10 ISBN 9812561315 Munoz Gerardo Jones Preston 2010 The equivalence principle uniformly accelerated reference frames and the uniform gravitational field American Journal of Physics 78 4 377 383 arXiv 1003 3022 Bibcode 2010AmJPh 78 377M doi 10 1119 1 3272719 S2CID 118616525 a b c Minguzzi E 2005 The Minkowski metric in non inertial observer radar coordinates American Journal of Physics 73 12 1117 1121 arXiv physics 0412024 Bibcode 2005AmJPh 73 1117M doi 10 1119 1 2060716 S2CID 119359878 a b c d David Tilbrook 1997 General Coordinatisations of the Flat Space Time of Constant Proper acceleration Australian Journal of Physics 50 5 851 868 doi 10 1071 P96111 Jones Preston Wanex Lucas F 2006 The Clock Paradox in a Static Homogeneous Gravitational Field Foundations of Physics Letters 19 1 75 85 arXiv physics 0604025 Bibcode 2006FoPhL 19 75J doi 10 1007 s10702 006 1850 3 S2CID 14583590 For instance Birrell amp Davies 1982 pp 110 111 or Padmanabhan 2010 p 126 denote equations 2g 2h as Rindler coordinates or Rindler frame Tilbrook 1997 pp 864 864 or Jones amp Wanex 2006 denote equations 2a 2b as Rindler coordinates a b Rohrlich Fritz 1963 The principle of equivalence Annals of Physics 22 2 169 191 Bibcode 1963AnPhy 22 169R doi 10 1016 0003 4916 63 90051 4 a b Harry Lass 1963 Accelerating Frames of Reference and the Clock Paradox American Journal of Physics 31 4 274 276 Bibcode 1963AmJPh 31 274L doi 10 1119 1 1969430 S2CID 121608504 a b Rindler W 1960 Hyperbolic Motion in Curved Space Time Physical Review 119 6 2082 2089 Bibcode 1960PhRv 119 2082R doi 10 1103 PhysRev 119 2082 a b c Rindler W 1966 Kruskal Space and the Uniformly Accelerated Frame American Journal of Physics 34 12 1174 1178 Bibcode 1966AmJPh 34 1174R doi 10 1119 1 1972547 a b Pauli Wolfgang 1921 Die Relativitatstheorie Encyclopadie der Mathematischen Wissenschaften 5 2 539 776 In English Pauli W 1981 1921 Theory of Relativity Vol 165 Dover Publications ISBN 0 486 64152 X a href Template Cite book html title Template Cite book cite book a journal ignored help von Laue M 1921 Die Relativitatstheorie Band 1 fourth edition of Das Relativitatsprinzip ed Vieweg First edition 1911 second expanded edition 1913 third expanded edition 1919 a b Don Koks 2006 Explorations in Mathematical Physics Springer pp 235 269 ISBN 0387309438 Moller C 1955 1952 The theory of relativity Oxford Clarendon Press Moller 1952 eq 154 Misner amp Thorne amp Wheeler 1973 section 6 6 Munoz amp Jones 2010 eq 37 38 Pauli 1921 section 32 y Rindler 1966 p 1177 span cla, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.