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 also leads to more complex definitions of "acceleration". SR as the theory of flat Minkowski spacetime remains valid in the presence of accelerations, because general relativity (GR) is only required when there is curvature of spacetime caused by the energy–momentum tensor (which is mainly determined by mass). However, since the amount of spacetime curvature is not particularly high on Earth or its vicinity, SR remains valid for most practical purposes, such as experiments in particle accelerators.[1]

One can derive transformation formulas for ordinary accelerations in three spatial dimensions (three-acceleration or coordinate acceleration) as measured in an external inertial frame of reference, as well as for the special case of proper acceleration measured by a comoving accelerometer. Another useful formalism is four-acceleration, as its components can be connected in different inertial frames by a Lorentz transformation. Also equations of motion can be formulated which connect acceleration and force. Equations for several forms of acceleration of bodies and their curved world lines follow from these formulas by integration. Well known special cases are hyperbolic motion for constant longitudinal proper acceleration or uniform circular motion. Eventually, it is also possible to describe these phenomena in accelerated frames in the context of special relativity, see Proper reference frame (flat spacetime). In such frames, effects arise which are analogous to homogeneous gravitational fields, which have some formal similarities to the real, inhomogeneous gravitational fields of curved spacetime in general relativity. In the case of hyperbolic motion one can use Rindler coordinates, in the case of uniform circular motion one can use Born coordinates.

Concerning the historical development, relativistic equations containing accelerations can already be found in the early years of relativity, as summarized in early textbooks by Max von Laue (1911, 1921)[2] or Wolfgang Pauli (1921).[3] For instance, equations of motion and acceleration transformations were developed in the papers of Hendrik Antoon Lorentz (1899, 1904),[H 1][H 2] Henri Poincaré (1905),[H 3][H 4] Albert Einstein (1905),[H 5] Max Planck (1906),[H 6] and four-acceleration, proper acceleration, hyperbolic motion, accelerating reference frames, Born rigidity, have been analyzed by Einstein (1907),[H 7] Hermann Minkowski (1907, 1908),[H 8][H 9] Max Born (1909),[H 10] Gustav Herglotz (1909),[H 11][H 12] Arnold Sommerfeld (1910),[H 13][H 14] von Laue (1911),[H 15][H 16] Friedrich Kottler (1912, 1914),[H 17] see section on history.

Three-acceleration edit

In accordance with both Newtonian mechanics and SR, three-acceleration or coordinate acceleration   is the first derivative of velocity   with respect to coordinate time or the second derivative of the location   with respect to coordinate time:

 .

However, the theories sharply differ in their predictions in terms of the relation between three-accelerations measured in different inertial frames. In Newtonian mechanics, time is absolute by   in accordance with the Galilean transformation, therefore the three-acceleration derived from it is equal too in all inertial frames:[4]

 .

On the contrary in SR, both   and   depend on the Lorentz transformation, therefore also three-acceleration   and its components vary in different inertial frames. When the relative velocity between the frames is directed in the x-direction by   with   as Lorentz factor, the Lorentz transformation has the form

 

(1a)

or for arbitrary velocities   of magnitude  :[5]

 

(1b)

In order to find out the transformation of three-acceleration, one has to differentiate the spatial coordinates   and   of the Lorentz transformation with respect to   and  , from which the transformation of three-velocity (also called velocity-addition formula) between   and   follows, and eventually by another differentiation with respect to   and   the transformation of three-acceleration between   and   follows. Starting from (1a), this procedure gives the transformation where the accelerations are parallel (x-direction) or perpendicular (y-, z-direction) to the velocity:[6][7][8][9][H 4][H 15]

 

(1c)

or starting from (1b) this procedure gives the result for the general case of arbitrary directions of velocities and accelerations:[10][11]

 

(1d)

This means, if there are two inertial frames   and   with relative velocity  , then in   the acceleration   of an object with momentary velocity   is measured, while in   the same object has an acceleration   and has the momentary velocity  . As with the velocity addition formulas, also these acceleration transformations guarantee that the resultant speed of the accelerated object can never reach or surpass the speed of light.

Four-acceleration edit

If four-vectors are used instead of three-vectors, namely   as four-position and   as four-velocity, then the four-acceleration   of an object is obtained by differentiation with respect to proper time   instead of coordinate time:[12][13][14]

 

(2a)

where   is the object's three-acceleration and   its momentary three-velocity of magnitude   with the corresponding Lorentz factor  . If only the spatial part is considered, and when the velocity is directed in the x-direction by   and only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity are considered, the expression is reduced to:[15][16]

 

Unlike the three-acceleration previously discussed, it is not necessary to derive a new transformation for four-acceleration, because as with all four-vectors, the components of   and   in two inertial frames with relative speed   are connected by a Lorentz transformation analogous to (1a, 1b). Another property of four-vectors is the invariance of the inner product   or its magnitude  , which gives in this case:[16][13][17]

 .

(2b)

Proper acceleration edit

In infinitesimal small durations there is always one inertial frame, which momentarily has the same velocity as the accelerated body, and in which the Lorentz transformation holds. The corresponding three-acceleration   in these frames can be directly measured by an accelerometer, and is called proper acceleration[18][H 14] or rest acceleration.[19][H 12] The relation of   in a momentary inertial frame   and   measured in an external inertial frame   follows from (1c, 1d) with  ,  ,   and  . So in terms of (1c), when the velocity is directed in the x-direction by   and when only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity are considered, it follows:[12][19][18][H 1][H 2][H 14][H 12]

 

(3a)

Generalized by (1d) for arbitrary directions of   of magnitude  :[20][21][17]

 

There is also a close relationship to the magnitude of four-acceleration: As it is invariant, it can be determined in the momentary inertial frame  , in which   and by   it follows  :[19][12][22][H 16]

 .

(3b)

Thus the magnitude of four-acceleration corresponds to the magnitude of proper acceleration. By combining this with (2b), an alternative method for the determination of the connection between   in   and   in   is given, namely[13][17]

 

from which (3a) follows again when the velocity is directed in the x-direction by   and only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity are considered.

Acceleration and force edit

Assuming constant mass  , the four-force   as a function of three-force   is related to four-acceleration (2a) by  , thus:[23][24]

 

(4a)

The relation between three-force and three-acceleration for arbitrary directions of the velocity is thus[25][26][23]

 

(4b)

When the velocity is directed in the x-direction by   and only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity are considered[27][26][23][H 2][H 6]

 

(4c)

Therefore, the Newtonian definition of mass as the ratio of three-force and three-acceleration is disadvantageous in SR, because such a mass would depend both on velocity and direction. Consequently, the following mass definitions used in older textbooks are not used anymore:[27][28][H 2]

  as "longitudinal mass",
  as "transverse mass".

The relation (4b) between three-acceleration and three-force can also be obtained from the equation of motion[29][25][H 2][H 6]

 

(4d)

where   is the three-momentum. The corresponding transformation of three-force between   in   and   in   (when the relative velocity between the frames is directed in the x-direction by   and only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity are considered) follows by substitution of the relevant transformation formulas for  ,  ,  ,  , or from the Lorentz transformed components of four-force, with the result:[29][30][24][H 3][H 15]

 

(4e)

Or generalized for arbitrary directions of  , as well as   with magnitude  :[31][32]

 

(4f)

Proper acceleration and proper force edit

The force   in a momentary inertial frame measured by a comoving spring balance can be called proper force.[33][34] It follows from (4e, 4f) by setting   and   as well as   and  . Thus by (4e) where only accelerations parallel (x-direction) or perpendicular (y-, z-direction) to the velocity   are considered:[35][33][34]

 

(5a)

Generalized by (4f) for arbitrary directions of   of magnitude  :[35][36]

 

Since in momentary inertial frames one has four-force   and four-acceleration  , equation (4a) produces the Newtonian relation  , therefore (3a, 4c, 5a) can be summarized[37]

 

(5b)

By that, the apparent contradiction in the historical definitions of transverse mass   can be explained.[38] Einstein (1905) described the relation between three-acceleration and proper force[H 5]

 ,

while Lorentz (1899, 1904) and Planck (1906) described the relation between three-acceleration and three-force[H 2]

 .

Curved world lines edit

By integration of the equations of motion one obtains the curved world lines of accelerated bodies corresponding to a sequence of momentary inertial frames (here, the expression "curved" is related to the form of the worldlines in Minkowski diagrams, which should not be confused with "curved" spacetime of general relativity). In connection with this, the so-called clock hypothesis of clock postulate has to be considered:[39][40] The proper time of comoving clocks is independent of acceleration, that is, the time dilation of these clocks as seen in an external inertial frame only depends on its relative velocity with respect to that frame. Two simple cases of curved world lines are now provided by integration of equation (3a) for proper acceleration:

a) Hyperbolic motion: The constant, longitudinal proper acceleration   by (3a) leads to the world line[12][18][19][25][41][42][H 10][H 15]

 

(6a)

The worldline corresponds to the hyperbolic equation  , from which the name hyperbolic motion is derived. These equations are often used for the calculation of various scenarios of the twin paradox or Bell's spaceship paradox, or in relation to space travel using constant acceleration.

b) The constant, transverse proper acceleration   by (3a) can be seen as a centripetal acceleration,[13] leading to the worldline of a body in uniform rotation[43][44]

 

(6b)

where   is the tangential speed,   is the orbital radius,   is the angular velocity as a function of coordinate time, and   as the proper angular velocity.

A classification of curved worldlines can be obtained by using the differential geometry of triple curves, which can be expressed by spacetime Frenet-Serret formulas.[45] In particular, it can be shown that hyperbolic motion and uniform circular motion are special cases of motions having constant curvatures and torsions,[46] satisfying the condition of Born rigidity.[H 11][H 17] A body is called Born rigid if the spacetime distance between its infinitesimally separated worldlines or points remains constant during acceleration.

Accelerated reference frames edit

Instead of inertial frames, these accelerated motions and curved worldlines can also be described using accelerated or curvilinear coordinates. The proper reference frame established that way is closely related to Fermi coordinates.[47][48] For instance, the coordinates for an hyperbolically accelerated reference frame are sometimes called Rindler coordinates, or those of a uniformly rotating reference frame are called rotating cylindrical coordinates (or sometimes Born coordinates). In terms of the equivalence principle, the effects arising in these accelerated frames are analogous to effects in a homogeneous, fictitious gravitational field. In this way it can be seen, that the employment of accelerating frames in SR produces important mathematical relations, which (when further developed) play a fundamental role in the description of real, inhomogeneous gravitational fields in terms of curved spacetime in general relativity.

History edit

For further information see von Laue,[2] Pauli,[3] Miller,[49] Zahar,[50] Gourgoulhon,[48] and the historical sources in history of special relativity.

1899:
Hendrik Lorentz[H 1] derived the correct (up to a certain factor  ) relations for accelerations, forces and masses between a resting electrostatic systems of particles   (in a stationary aether), and a system   emerging from it by adding a translation, with   as the Lorentz factor:
 ,  ,   for   by (5a);
 ,  ,   for   by (3a);
 ,  ,   for  , thus longitudinal and transverse mass by (4c);
Lorentz explained that he has no means of determining the value of  . If he had set  , his expressions would have assumed the exact relativistic form.
1904:
Lorentz[H 2] derived the previous relations in a more detailed way, namely with respect to the properties of particles resting in the system   and the moving system  , with the new auxiliary variable   equal to   compared to the one in 1899, thus:
  for   as a function of   by (5a);
  for   as a function of   by (5b);
  for   as a function of   by (3a);
  for longitudinal and transverse mass as a function of the rest mass by (4c, 5b).
This time, Lorentz could show that  , by which his formulas assume the exact relativistic form. He also formulated the equation of motion
  with  
which corresponds to (4d) with  , with  ,  ,  ,  ,  , and   as electromagnetic rest mass. Furthermore, he argued, that these formulas should not only hold for forces and masses of electrically charged particles, but for other processes as well so that the earth's motion through the aether remains undetectable.
1905:
Henri Poincaré[H 3] introduced the transformation of three-force (4e):
 
with  , and   as the Lorentz factor,   the charge density. Or in modern notation:  ,  ,  , and  . As Lorentz, he set  .
1905:
Albert Einstein[H 5] derived the equations of motions on the basis of his special theory of relativity, which represent the relation between equally valid inertial frames without the action of a mechanical aether. Einstein concluded, that in a momentary inertial frame   the equations of motion retain their Newtonian form:
 .
This corresponds to  , because   and   and  . By transformation into a relatively moving system   he obtained the equations for the electrical and magnetic components observed in that frame:
 .
This corresponds to (4c) with  , because   and   and   and  . Consequently, Einstein determined the longitudinal and transverse mass, even though he related it to the force   in the momentary rest frame measured by a comoving spring balance, and to the three-acceleration   in system  :[38]
 
This corresponds to (5b) with  .
1905:
Poincaré[H 4] introduces the transformation of three-acceleration (1c):
 
where   as well as   and   and  .
Furthermore, he introduced the four-force in the form:
 
where   and   and  .
1906:
Max Planck[H 6] derived the equation of motion
 
with
  and  
and
 
The equations correspond to (4d) with
 , with   and   and  , in agreement with those given by Lorentz (1904).
1907:
Einstein[H 7] analyzed a uniformly accelerated reference frame and obtained formulas for coordinate dependent time dilation and speed of light, analogous to those given by Kottler-Møller-Rindler coordinates.
1907:
Hermann Minkowski[H 9] defined the relation between the four-force (which he called the moving force) and the four acceleration
 
corresponding to  .
1908:
Minkowski[H 8] denotes the second derivative   with respect to proper time as "acceleration vector" (four-acceleration). He showed, that its magnitude at an arbitrary point   of the worldline is  , where   is the magnitude of a vector directed from the center of the corresponding "curvature hyperbola" (German: Krümmungshyperbel) to  .
1909:
Max Born[H 10] denotes the motion with constant magnitude of Minkowski's acceleration vector as "hyperbolic motion" (German: Hyperbelbewegung), in the course of his study of rigidly accelerated motion. He set   (now called proper velocity) and   as Lorentz factor and   as proper time, with the transformation equations
 .
which corresponds to (6a) with   and  . Eliminating   Born derived the hyperbolic equation  , and defined the magnitude of acceleration as  . He also noticed that his transformation can be used to transform into a "hyperbolically accelerated reference system" (German: hyperbolisch beschleunigtes Bezugsystem).
1909:
Gustav Herglotz[H 11] extends Born's investigation to all possible cases of rigidly accelerated motion, including uniform rotation.
1910:
Arnold Sommerfeld[H 13] brought Born's formulas for hyperbolic motion in a more concise form with   as the imaginary time variable and   as an imaginary angle:
 
He noted that when   are variable and   is constant, they describe the worldline of a charged body in hyperbolic motion. But if   are constant and   is variable, they denote the transformation into its rest frame.
1911:
Sommerfeld[H 14] explicitly used the expression "proper acceleration" (German: Eigenbeschleunigung) for the quantity   in  , which corresponds to (3a), as the acceleration in the momentary inertial frame.
1911:
Herglotz[H 12] explicitly used the expression "rest acceleration" (German: Ruhbeschleunigung) instead of proper acceleration. He wrote it in the form   and   which corresponds to (3a), where   is the Lorentz factor and   or   are the longitudinal and transverse components of rest acceleration.
1911:
Max von Laue[H 15] derived in the first edition of his monograph "Das Relativitätsprinzip" the transformation for three-acceleration by differentiation of the velocity addition
 
equivalent to (1c) as well as to Poincaré (1905/6). From that he derived the transformation of rest acceleration (equivalent to 3a), and eventually the formulas for hyperbolic motion which corresponds to (6a):
 
thus
 ,
and the transformation into a hyperbolic reference system with imaginary angle  :
 .
He also wrote the transformation of three-force as
 
equivalent to (4e) as well as to Poincaré (1905).
1912–1914:
Friedrich Kottler[H 17] obtained general covariance of Maxwell's equations, and used four-dimensional Frenet-Serret formulas to analyze the Born rigid motions given by Herglotz (1909). He also obtained the proper reference frames for hyperbolic motion and uniform circular motion.
1913:
von Laue[H 16] replaced in the second edition of his book the transformation of three-acceleration by Minkowski's acceleration vector for which he coined the name "four-acceleration" (German: Viererbeschleunigung), defined by   with   as four-velocity. He showed, that the magnitude of four-acceleration corresponds to the rest acceleration   by
 ,
which corresponds to (3b). Subsequently, he derived the same formulas as in 1911 for the transformation of rest acceleration and hyperbolic motion, and the hyperbolic reference frame.

References edit

  1. ^ Misner & Thorne & Wheeler (1973), p. 163: "Accelerated motion and accelerated observers can be analyzed using special relativity."
  2. ^ a b von Laue (1921)
  3. ^ a b Pauli (1921)
  4. ^ Sexl & Schmidt (1979), p. 116
  5. ^ Møller (1955), p. 41
  6. ^ Tolman (1917), p. 48
  7. ^ French (1968), p. 148
  8. ^ Zahar (1989), p. 232
  9. ^ Freund (2008), p. 96
  10. ^ Kopeikin & Efroimsky & Kaplan (2011), p. 141
  11. ^ Rahaman (2014), p. 77
  12. ^ a b c d Pauli (1921), p. 627
  13. ^ a b c d Freund (2008), pp. 267-268
  14. ^ Ashtekar & Petkov (2014), p. 53
  15. ^ Sexl & Schmidt (1979), p. 198, Solution to example 16.1
  16. ^ a b Ferraro (2007), p. 178
  17. ^ a b c Kopeikin & Efroimsky & Kaplan (2011), p. 137
  18. ^ a b c Rindler (1977), pp. 49-50
  19. ^ a b c d von Laue (1921), pp. 88-89
  20. ^ Rebhan (1999), p. 775
  21. ^ Nikolić (2000), eq. 10
  22. ^ Rindler (1977), p. 67
  23. ^ a b c Sexl & Schmidt (1979), solution of example 16.2, p. 198
  24. ^ a b Freund (2008), p. 276
  25. ^ a b c Møller (1955), pp. 74-75
  26. ^ a b Rindler (1977), pp. 89-90
  27. ^ a b von Laue (1921), p. 210
  28. ^ Pauli (1921), p. 635
  29. ^ a b Tolman (1917), pp. 73-74
  30. ^ von Laue (1921), p. 113
  31. ^ Møller (1955), p. 73
  32. ^ Kopeikin & Efroimsky & Kaplan (2011), p. 173
  33. ^ a b Shadowitz (1968), p. 101
  34. ^ a b Pfeffer & Nir (2012), p. 115, "In the special case in which the particle is momentarily at rest relative to the observer S, the force he measures will be the proper force".
  35. ^ a b Møller (1955), p. 74
  36. ^ Rebhan (1999), p. 818
  37. ^ see Lorentz's 1904-equations and Einstein's 1905-equations in section on history
  38. ^ a b Mathpages (see external links), "Transverse Mass in Einstein's Electrodynamics", eq. 2,3
  39. ^ Rindler (1977), p. 43
  40. ^ Koks (2006), section 7.1
  41. ^ Fraundorf (2012), section IV-B
  42. ^ PhysicsFAQ (2016), see external links.
  43. ^ Pauri & Vallisneri (2000), eq. 13
  44. ^ Bini & Lusanna & Mashhoon (2005), eq. 28,29
  45. ^ Synge (1966)
  46. ^ Pauri & Vallisneri (2000), Appendix A
  47. ^ Misner & Thorne & Wheeler (1973), Section 6
  48. ^ a b Gourgoulhon (2013), entire book
  49. ^ Miller (1981)
  50. ^ Zahar (1989)

Bibliography edit

  • Ashtekar, A.; Petkov, V. (2014). Springer Handbook of Spacetime. Springer. ISBN 978-3642419928.
  • Bini, D.; Lusanna, L.; Mashhoon, B. (2005). "Limitations of radar coordinates". International Journal of Modern Physics D. 14 (8): 1413–1429. arXiv:gr-qc/0409052. Bibcode:2005IJMPD..14.1413B. doi:10.1142/S0218271805006961. S2CID 17909223.
  • Ferraro, R. (2007). Einstein's Space-Time: An Introduction to Special and General Relativity. Spektrum. ISBN 978-0387699462.
  • Fraundorf, P. (2012). "A traveler-centered intro to kinematics". IV-B. arXiv:1206.2877 [physics.pop-ph].
  • French, A.P. (1968). Special Relativity. CRC Press. ISBN 1420074814.
  • Freund, J. (2008). Special Relativity for Beginners: A Textbook for Undergraduates. World Scientific. ISBN 978-9812771599.
  • Gourgoulhon, E. (2013). Special Relativity in General Frames: From Particles to Astrophysics. Springer. ISBN 978-3642372766.
  • 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.
  • Koks, D. (2006). Explorations in Mathematical Physics. Springer. ISBN 0387309438.
  • Kopeikin, S.; Efroimsky, M.; Kaplan, G. (2011). Relativistic Celestial Mechanics of the Solar System. John Wiley & Sons. ISBN 978-3527408566.
  • Miller, Arthur I. (1981). Albert Einstein's special theory of relativity. Emergence (1905) and early interpretation (1905–1911). Reading: Addison–Wesley. ISBN 0-201-04679-2.
  • Misner, C. W.; Thorne, K. S.; Wheeler, J. A. (1973). Gravitation. Freeman. ISBN 0716703440.
  • Møller, C. (1955) [1952]. The theory of relativity. Oxford Clarendon Press.
  • Nikolić, H. (2000). "Relativistic contraction and related effects in noninertial frames". Physical Review A. 61 (3): 032109. arXiv:gr-qc/9904078. Bibcode:2000PhRvA..61c2109N. doi:10.1103/PhysRevA.61.032109. S2CID 5783649.
  • 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)

Historical papers edit

  1. ^ a b c Lorentz, Hendrik Antoon (1899). "Simplified Theory of Electrical and Optical Phenomena in Moving Systems" . Proceedings of the Royal Netherlands Academy of Arts and Sciences. 1: 427–442. Bibcode:1898KNAB....1..427L.
  2. ^ a b c d e f g Lorentz, Hendrik Antoon (1904). "Electromagnetic phenomena in a system moving with any velocity smaller than that of light" . Proceedings of the Royal Netherlands Academy of Arts and Sciences. 6: 809–831. Bibcode:1903KNAB....6..809L.
  3. ^ a b c Poincaré, Henri (1905). "Sur la dynamique de l'électron"  [Wikisource translation: On the Dynamics of the Electron]. Comptes rendus hebdomadaires des séances de l'Académie des sciences. 140: 1504–1508.
  4. ^ a b c Poincaré, Henri (1906) [1905]. "Sur la dynamique de l'électron"  [Wikisource translation: On the Dynamics of the Electron]. Rendiconti del Circolo Matematico di Palermo. 21: 129–176. Bibcode:1906RCMP...21..129P. doi:10.1007/BF03013466. hdl:2027/uiug.30112063899089. S2CID 120211823.
  5. ^ a b c Einstein, Albert (1905). "Zur Elektrodynamik bewegter Körper". Annalen der Physik. 322 (10): 891–921. Bibcode:1905AnP...322..891E. doi:10.1002/andp.19053221004.; See also: English translation.
  6. ^ a b c d Planck, Max (1906). "Das Prinzip der Relativität und die Grundgleichungen der Mechanik" [Wikisource translation: The Principle of Relativity and the Fundamental Equations of Mechanics]. Verhandlungen Deutsche Physikalische Gesellschaft. 8: 136–141.
  7. ^ a b Einstein, Albert (1908) [1907], "Über das Relativitätsprinzip und die aus demselben gezogenen Folgerungen" (PDF), Jahrbuch der Radioaktivität und Elektronik, 4: 411–462, Bibcode:1908JRE.....4..411E; English translation On the relativity principle and the conclusions drawn from it at Einstein paper project.
  8. ^ a b Minkowski, Hermann (1909) [1908]. "Raum und Zeit. Vortrag, gehalten auf der 80. Naturforscher-Versammlung zu Köln am 21. September 1908"  [Wikisource translation: Space and Time]. Jahresbericht der Deutschen Mathematiker-Vereinigung. Leipzig.
  9. ^ a b Minkowski, Hermann (1908) [1907], "Die Grundgleichungen für die elektromagnetischen Vorgänge in bewegten Körpern"  [Wikisource translation: The Fundamental Equations for Electromagnetic Processes in Moving Bodies], Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse: 53–111
  10. ^ a b c Born, Max (1909). "Die Theorie des starren Elektrons in der Kinematik des Relativitätsprinzips" [Wikisource translation: The Theory of the Rigid Electron in the Kinematics of the Principle of Relativity]. Annalen der Physik. 335 (11): 1–56. Bibcode:1909AnP...335....1B. doi:10.1002/andp.19093351102.
  11. ^ a b c Herglotz, G (1910) [1909]. "Über den vom Standpunkt des Relativitätsprinzips aus als starr zu bezeichnenden Körper" [Wikisource translation: On bodies that are to be designated as "rigid" from the standpoint of the relativity principle]. Annalen der Physik. 336 (2): 393–415. Bibcode:1910AnP...336..393H. doi:10.1002/andp.19103360208.
  12. ^ a b c d Herglotz, G. (1911). "Über die Mechanik des deformierbaren Körpers vom Standpunkte der Relativitätstheorie". Annalen der Physik. 341 (13): 493–533. Bibcode:1911AnP...341..493H. doi:10.1002/andp.19113411303.
  13. ^ a b Sommerfeld, Arnold (1910). "Zur Relativitätstheorie II: Vierdimensionale Vektoranalysis" [Wikisource translation: On the Theory of Relativity II: Four-dimensional Vector Analysis]. Annalen der Physik. 338 (14): 649–689. Bibcode:1910AnP...338..649S. doi:10.1002/andp.19103381402.
  14. ^ a b c d Sommerfeld, Arnold (1911). "Über die Struktur der gamma-Strahlen". Sitzungsberichte der Mathematematisch-physikalischen Klasse der K. B. Akademie der Wissenschaften zu München (1): 1–60.
  15. ^ a b c d e Laue, Max von (1911). Das Relativitätsprinzip. Braunschweig: Vieweg.
  16. ^ a b c Laue, Max von (1913). Das Relativitätsprinzip (2. Ausgabe ed.). Braunschweig: Vieweg.
  17. ^ a b c Kottler, Friedrich (1912). "Über die Raumzeitlinien der Minkowski'schen Welt" [Wikisource translation: On the spacetime lines of a Minkowski world]. Wiener Sitzungsberichte 2a. 121: 1659–1759. hdl:2027/mdp.39015051107277. Kottler, Friedrich (1914a). "Relativitätsprinzip und beschleunigte Bewegung". Annalen der Physik. 349 (13): 701–748. Bibcode:1914AnP...349..701K. doi:10.1002/andp.19143491303. Kottler, Friedrich (1914b). "Fallende Bezugssysteme vom Standpunkte des Relativitätsprinzips". Annalen der Physik. 350 (20): 481–516. Bibcode:1914AnP...350..481K. doi:10.1002/andp.19143502003.

External links edit