Вы находитесь на странице: 1из 6

The Lorentz transformation: Simplication through complexication

Ezra T. Newman
Department of Physics and Astronomy, University of Pittsburgh, Pennsylvania 15260

Richard H. Price
Department of Physics and Astronomy and Center for Gravitational Wave Astronomy, University of Texas at Brownsville, Brownsville, Texas 78520

Received 9 May 2009; accepted 11 August 2009 The general Lorentz transformation, the relation between observations made in different frames, involves the three components of the relative velocity of the two frames and three parameters describing the relative orientation of the frames. A simpler description is known in which the six parameters are encoded in four complex numbers subject to a single algebraic constraint. This description is tied to the transformation of the directions of light propagation and underscores how closely related special relativity is to light propagation. This form of the transformation, which is particularly useful for dealing with the aberration of light is usually introduced through advanced mathematical formulations. We show that the basic idea, in addition to being useful, is quite simple. 2010 American Association of Physics Teachers. DOI: 10.1119/1.3216468 I. INTRODUCTION The Lorentz transformation is the mathematical heart of special relativity. In addition to its importance, it is algebraically simple in the context in which it is usually rst presented: A boost relation of relatively moving reference frames that is aligned with one of the spatial coordinate axes. This simplicity disappears for a more general Lorentz transformation, one in which the two reference frames are rotated about each other around an axis that is not parallel to a coordinate axis, or in which the boost relating the reference frames is not aligned with a coordinate axis. It is often the case that a change in formalism leads to a considerable change in the difculty of performing calculations and sometimes to a useful alternative way of thinking about a problem. An example is the use of Laplace transforms for solving problems with linear passive electrical circuits. In somewhat the same spirit, we give here a new view of the Lorentz transformation, one that shares with the Laplace transform the seemingly magical simplication that complex variables can afford. The new here is in quotes because this formalism is not at all new. For several decades experts in mathematical relativity have known that the special relativistic Lorentz transformation can be put in a form known variously as a linear fractional transformation or as a Mbius transformation.13 Discussions of this topic have been conned to advanced mathematical treatments. Our goal here is to show the central ideas of this alternative formalism in a straightforward and self-contained way. To do this we have stripped away the mathematical connections and vocabulary that are typically found with this form of the Lorentz transformation. For readers who might be curious about the mathematics at a deeper level we provide a brief sketch in the Appendix. This paper introduces two new formalisms for Lorentz transformations. The rst of these uses 2 2 matrices and is explained in Sec. II and is contrasted with the standard way of working with Lorentz transformations. Although this formalism has calculational advantages, we consider it prima14 Am. J. Phys. 78 1 , January 2010 http://aapt.org/ajp

rily to be a useful intermediate step in the introduction of the aberration transform in Sec. III. We give a brief summary in Sec. IV. Throughout the paper we follow the common convention of treating the velocity of light as unity so that we may refer to the dimensionless parameter = v / c as the velocity and claim that t2 x2 y 2 z2 is left invariant by a Lorentz transformation. To denote matrices, column, row, or square, whether for two or four dimensions, we use blackboard indicates complex conbold fonts, as in X. An asterisk jugate, and a dagger on a matrix; for example, X indicates the Hermitian conjugate the transpose of the matrix with all elements replaced by their complex conjugates . II. 4 4 MATRICES AND 2 2 MATRICES Lorentz transformations consist of linear transformations of the real coordinates t , x , y , z of an event to new coordinates t , x , y , z representing measurements made in a different special relativistic reference frame. Typical examples are a boost by a velocity in the z direction,4 t = x = x, a rotation by t = t, t z , y = y, z = z t 1/ 1
2

about the z axis,

x = cos x + sin y z = z, x = x, y = y, y = y. 2 3 4 5

y = sin x + cos y, t = t, t = t, t = t, z = y, z = z, z = z, y = z, x = x, x = x,

a rotation by 90 about the x axis, an inversion through the z = 0 plane, and time reversal, The last two examples, which involve inversions, are called improper.
2010 American Association of Physics Teachers 14

The dening property of a Lorentz transformation is that it is a linear transformation for which t 2 x 2 y 2 z 2 = t2 x2 y 2 z2 . 6

t2 x2 y 2 z2 = Det X ,

12

so that the Lorentz transformation A requires that the determinants of X and X are equal. We have Det X = Det A Det X Det A = Det A 2Det X , 13 so the condition for Eq. 10 to be a Lorentz transformation is that Det A = 1. For a matrix A that has Det A = 1 but Det A 1, a trivial multiplication of each of its matrix elements by some unit magnitude number ei reduces A to a matrix with unit determinant and does not affect the transformation in Eq. 10 . With no loss of generality we can therefore conne our attention to matrices A with Det A = 1, that is, with ad bc = 1. 14

It is easy to check that the examples in Eqs. 1 5 satisfy this condition. These transformations can be applied multiple times or one after another. As an example, applying Eq. 3 three times in a row gives a rotation about the x axis by 270 or by 90. As another example, if we apply Eq. 3 followed by Eq. 2 and then by a rotation about x by 90, we have a rotation about y by the angle . We can arrive at any Lorentz transformation by performing such basic rotations and boosts one after the other. Such calculations with Lorentz transformations are handled best by arranging the original t , x , y , z and the transformed t , x , y , z coordinates as column vectors X and X and viewing the transformation as the matrix equation X = LX, 7

in which the matrix L contains the details of the Lorentz transformation. For the special case of a boost in the z direction, the case explicitly given in Eq. 1 , the matrix representation is t x y z 0 0 = 0 0 1 0 0 1 0 0 0 0 t x y z

The six parameters of the general Lorentz transformation are thus encoded in the four complex numbers a, b, c, and d constrained by the two conditions real and imaginary parts in Eq. 14 . The advantages of this 2 2 approach may be seen in the matrix elements that correspond to the transformations in Eqs. 1 3 . The boost by in the z direction is represented by 0 0 1/ ,
1/4

15 or equivalently = 1 4 / 1 about the z axis is represented by 16

The matrix for this simple boost along the z direction is simple, but in general L is notoriously complicated5 because it embodies the six pieces of information that describe the general Lorentz transformation: The three parameters that describe the general rotation and the three that describe the velocity of the general boost. An alternative version of the Lorentz transformation starts with the packing of the coordinates t , x , y , z of a spacetime event into a 2 2 matrix in the pattern6 X= t+z x iy x + iy tz . 9

with = 1 / 1 + + 4 ; the rotation by ei 0


/2

0 e
i /2

and a rotation by 90 about the x axis is represented by 1/ 2 i/ 2 The matrix 1/ 2 i/ 2 i/ 2 1/ 2 18 i/ 2 1/ 2 . 17

Note that this matrix is Hermitian. Given any coordinates t , x , y , z we can arrange them in this form. Conversely given any 2 2 Hermitian matrix7 we can read off the values of the coordinates. For example, the t coordinate is equal to half the sum of the diagonal elements. A proper Lorentz transformation consists of any linear transformation from t , x , y , z to t , x , y , z for which t +z x iy x + iy t z = a b c d a b c d t+z x iy . x + iy tz

is the inverse of the matrix in Eq. 17 , and hence, it represents a rotation about x by 90. To nd the 2 2 matrix representing any Lorentz transformations, we can multiply the matrices in order as we can with 4 4 matrices. Thus 1/ 2 i/ 2 ei 0
/2

0 e
i /2

1/ 2 i/ 2

i/ 2 1/ 2 19

10

i/ 2 1/ 2 = cos /2 sin /2

We can write Eq. 10 in matrix notation as X = AXA . 11

sin /2 cos /2

It is easy to check that X is Hermitian if X is Hermitian, but we must ask whether Eq. 10 satises the condition of Eq. 6 . Notice that
15 Am. J. Phys., Vol. 78, No. 1, January 2010

tells us that a 90 rotation about x, followed by a rotation about z, and followed by a 90 rotation about x results in a rotation about y. As another example, with = 1 / 1 + 1/4 the transformation
Ezra T. Newman and Richard H. Price 15

1/ 2 i/ 2 i/ 2 1/ 2 1 2 i 2 + 1/ 1/

0 0 1/ i 2

1/ 2 i/ 2 1/

i/ 2 1/ 2

It is easily checked that all the coefcients in these formulas are real. III. THE ABERRATION TRANSFORM 20 We now focus attention on a particular class of spacetime intervals, those that are null, and relate two spacetime points that can lie on a null line or, from a physical point of view, on the worldline of a photon. The two points are the origin and the point t , x , y , z . The condition for the two points to have a null separation is t2 x2 y 2 z2 = 0. For denitiveness we also add the condition that the point t , x , y , z is to the future of the origin so that t z and t + z are both nonnegative. Under these restrictions a remarkable simplication applies, and we can express the Lorentz transformation in a way very different from that of 4 4 or 2 2 matrices. In principle we can do so as a direct consequence only of the relations in Eq. 22 , but the algebraic manipulations would be too much of an unmotivated miracle. Instead we start by noticing8 that the matrix X in Eq. 9 can be written as the product of a two-component column vector and the two-component row vector that is its Hermitian conjugate, X t+z x iy x + iy tz x + iy = tz tz x iy tz tz , 23

1 2

+ 1/

is a boost by in the negative y direction. As a more complicated example, we can use the 2 2 matrices to nd a boost in z = 0 plane at an angle from the negative y axis by combining rotations of about the z axis with the result in Eq. 20 ,
i /2

0 ei ei 0 1 2 i 2
/2

1 2 i 2 0 e
i /2

+ 1/ 1/

i 2

1/ + 1/

1 2

/2

+ 1/ 1/ ei

i 2 1 2

1/ + 1/

21

In a similar manner, any Lorentz transformation can be constructed from the basic building blocks in Eqs. 15 18 . Because the 2 2 matrices represent any Lorentz transformation, there must exist a decoding of those matrices that produces the elements of the 4 4 matrices, which give more familiar transformations of coordinates. These are found from straightforward substitution to be x =
1 2

and similarly for X . Note that we are using the null condition x2 + y 2 = t2 z2 = t z t + z to replace x2 + y 2 / t z by t + z. With this decomposition of X we can write the Lorentz transformation in Eq. 11 as x + iy t z t z x iy t z x iy tz x + iy t z =A tz tz t z A , 24

ad + cb + bc + da x

i bc + da ad cb y + ac + ca bd db z + ac + ca + bd + db t , y =
1 2

from which we can infer 22a x + iy t z t z =A x + iy tz = tz c a x + iy tz x + iy tz +b tz . +d tz 25

i cb + da ad bc x

cb + bc ad da y + i ca + bd ac db z + i ca + db ac bd t , z =
1 2

22b

This matrix relation contains two equations. The ratio of the top equation to the bottom equation is a x + iy / t z + b x + iy = . t z c x + iy / t z + d 26

ab + ba cd dc x

i ba + cd ab dc y + aa + dd cc bb z + aa + bb cc dd t , t =
1 2

22c

We next notice that for a photon whose worldline connects the origin and t , x , y , z , the ratios x / t, y / t, and z / t are the direction cosines for the photons path through space. To put Eq. 26 into the most useful form, we dene x/t + i y/t x + iy = tz 1 z/t = sin cos + i sin 1 cos 27
16

ab + cd + ba + dc x

i ba + dc ab cd y + aa + cc bb dd z + aa + bb + cc + dd t .
16 Am. J. Phys., Vol. 78, No. 1, January 2010

22d

= ei cot /2 .
Ezra T. Newman and Richard H. Price

The trigonometric forms follow from the introduction of spherical coordinates , , which are related to x , y , z in the usual way. The last expression in Eq. 27 follows simply from the trigonometric half angle formulas. In the Appendix it is shown how this particular trigonometric expression is related to stereographic projection. With this simplied, geometrically related notation, Eq. 26 takes the form = a +b c +d 28

+ 1/ = 1/

+1

34d

The imaginary part and the absolute value in Eq. 32 give us, respectively, tan cos = I/R, = R2 + I2 + D2 . R2 + I2 D2 35a 35b

and represents a relation between the angular direction , observed for photon propagation and the angular direction , observed in another frame. The redistribution of photon directions is the familiar aberration of light, or headlight effect.9 As a simple example, we consider the primed observing frame to be moving in the +z direction with speed . Equations 15 , 27 , and 28 give ei sin 1 cos sin = 2e i . 1 cos 29

The lengthy and complicated nature of the results in Eqs. 34 and 35 , compared with the simplicity of Eq. 32 , is instructive. It makes a case for using the complex variable rather than the spherical angles and for encoding directional information. This viewpoint has been very useful in mathematical relativity.10

IV. CONCLUSIONS In Sec. II we introduced complex 2 2 matrices as an alternative to the standard way of dealing with Lorentz transformations. This treatment was followed in Sec. III by the aberration formalism in which a proper Lorentz transformation was shown to carry the same information as a special complex transformation of the complex variable , a transformation which can be interpreted as the way in which photon propagation directions are related in two inertial frames. We found in both the 2 2 formalism and the aberration formalism that the six parameters that specify the general proper Lorentz transformation can be represented by four complex numbers that satisfy one simple relation the condition ad bc = 1 . We can therefore think of the Lorentz transformation either as a transformation of spacetime points, as described by Eq. 11 , or as a transformation of photon directions, as described by Eqs. 27 and 28 . A curious reader may notice some curious features of our presentation in Secs. II and III. One such feature is that Eq. 22 is quadratic in the parameters a , b , c , d so that reversing the sign of all the complex parameters a , b , c , d gives the same Lorentz transformation and continues to satisfy the ad bc = 1 constraint. Thus the a , b , c , d parameters for a Lorentz transformation are not unique, and therefore the 2 2 matrix for a specic Lorentz transformation is not unique. From a mathematical point of view it might seem curious that the particular class of complex functions of on the right-hand side of Eq. 28 is broad enough and narrow enough to represent only and all proper Lorentz transforms. From a physical point of view it might seem curious that the aberration transformation completely describes a proper Lorentz transformation. That is, if we know the relation between the directions11 and observed for photon motion in one frame and the directions and observed in another frame for the same photons, then we can infer all the parameters a , b , c , d , and hence, we have determined the Lorentz transformation of any interval, null, spacelike, or timelike. The transformation of photon directions xes the transformation of everything. This might seem more plausible when we realize that clocks and measuring rods can be constructed, in principle, using light rays. Transforming all light rays correctly, therefore, is acceptable as the only test that the Lorentz transformation needs to satisfy. Perhaps the aberraEzra T. Newman and Richard H. Price 17

Equation 29 shows that ei / ei must be real and positive, and thus we conclude that = , in agreement with the symmetry of the conguration. Next, we square the magnitude of the two sides and simplify to give 1 + cos 1 cos =
41

+ cos . 1 cos

30 leads to the familiar result = cos . 1 cos 31

Finally, solving Eq. 30 for cos


4

cos

1+ +1+

4 4

+ 1 cos 1 cos

The advantages of working with become more apparent for more complicated relations of coordinate frames. An example is the boost described in the matrix on the right-hand side of Eq. 21 . For this case = + 1/ 1/ i ei 1/ ei . + + 1/ 32

To nd the relation of spherical coordinates in the two frames, we rewrite this result as ei cot where R= +
2

/2 =

R + iI , D

33

cot /2 cos 1 + cot2 /2 sin cot /2 cos 2 , 34a

I=

cot /2 sin 1 + cot2 /2 cos cot /2 sin 2 , 34b

D=
17

2 cot /2 + cot2 /2 ,

34c

Am. J. Phys., Vol. 78, No. 1, January 2010

tion transform is the more fundamental way of describing the relation of measurements in two different inertial frames. These curious features may lead readers to look more carefully into the rich mathematical background of the representations of the Lorentz transformations. A reader looking only for a brief commentary on the background might be satised with the sketchy treatment in the Appendix. ACKNOWLEDGMENT One of the authors R.H.P. acknowledges support by the National Science Foundation under Grant No. PHY0554367. APPENDIX: THE MATHEMATICAL BACKGROUND Transformations typically form a group, which means that if we make one transformation and then a second transformation, the combined result is a transformation. The Lorentz transformations form a group12 the Lorentz group and multiplication of the 2 2 matrices in Sec. II also form a group: SL 2,C for special unit determinant , linear matrix , 2 2 2 , and C complex .6,13 Groups can be viewed in an abstract way in which one concentrates on the rules for the group unencumbered by physical meaning or visual pictures. In this way it can be shown that the Lorentz group consists of disjoint sets, depending on whether they preserve or reverse handedness and the direction of time. We conne ourselves to proper Lorentz transformations, those that preserve handedness and the direction of time. In the abstract view of groups one can show formally what we have shown by construction in Sec. II: Any proper Lorentz transformation can be represented by an SL 2,C matrix, and calculations can be done with the SL 2,C matrices. The connection between the two groups is not isomorphic, that is, one to one. Rather there are two SL 2,C matrices corresponding to every proper Lorentz transform. If a SL 2,C matrix A corresponds to a particular Lorentz transformation, then so does A. The 2:1 feature does not apply to the relation between proper Lorentz transformations and the aberration transformation in Sec. III. If we change the sign of each of the parameters a , b , c , d , the right-hand side of Eq. 28 does not change. This invariance of the transformation under sign change suggests that the proper Lorentz group is isomorphic to the aberration transformations, and this can formally shown to be true.3 That is, for each proper Lorentz transformation, there is one and only one complex function with the form of the last expression in Eq. 28 and with ad bc = 1 ; the relation is one to one. The 2 2 matrices might seem familiar from another context. The common sense representation of the Lorentz group in terms of 4 4 matrices is thought of as a vector representation because the matrices act on four-dimensional column vectors equivalent to the four-vectors of relativity. By contrast the 2 2 matrices act on two-component columns, called spinors, that are mathematical constructs not directly related to any physical vectors. Because the rotation group is a subgroup of the proper Lorentz group, the twocomponent spinors serve also to give a representation of the rotation group; we need only ignore the t in Eq. 9 . It turns out that the subgroup of SL 2,C that represents rotations, but not boosts, is the group of matrices that, in addition to being in SL 2,C , are unitary. This is the famous SU 2
18 Am. J. Phys., Vol. 78, No. 1, January 2010

z
P 1 (x,y,z) plane P (X,Y)
2

y,Y

x,X
Fig. 1. The stereographic projection of the unit sphere onto the complex plane. The XY plane passes through the center of the unit sphere. Point P1 on the sphere is mapped to point P2 on the XY plane by a line starting at the top of the sphere. It is straightforward to show that the spherical polar coordinates , of P1 are related to the point X , Y in the plane by ei cot / 2 = X + iY .

group special, unitary 2 2 , so important for dealing with half-integer spin in quantum mechanics. Another way of looking at the transformation in Eq. 28 is in terms of functions of a complex variable. This viewpoint turns out though not obviously to be connected to the fact that the proper Lorentz group can be shown to be isomorphic to the conformal group of the two-sphere,3 which is the group of transformations of points on the surface of a unit sphere that leave unchanged the angles between any two curves on the two-sphere. Such an angle-preserving conformal transformation is not as easy to picture as a symmetry transformation such as a rotation of the two-sphere, so it is best to revert to the relatively familiar mathematics of complex variable theory. It is known2 that the class of complex functions that conformally map the complex plane, including innity, is the class of the form F = A + B / C + D , where A, B, C, and D are complex constants.2 Note that there are really only three independent constants in this expression because we can always rescale the numerator and denominator by the same complex factor to make any of the nonzero constants equal to unity. We can make this three-parameter dependence explicit by choosing to write the class of everywhere-conformal functions as F= a +b / c +d , A1 where ad bc = 1. That takes care of conformal transformations in the complex plane. Next we consider the connection between a twosphere and the complex plane via the stereographic plane projection14 as pictured in Fig. 1: The complex passes through the center of a unit sphere; a line starting at the top of the unit sphere passes through the sphere connecting the point P1 on the sphere to the point P2 in the plane. This connection is well known to be conformal so that the function F in Eq. A1 induces a conformal transformation of the points on the unit sphere to points on the plane. Simple trigonometry shows that by stereographic projection, the spherical coordinates and on the unit sphere are connected symbolized by to the points on the complex plane by ei cot /2 X + iY = . A2 It follows that the conformal transformation of given by the function F induces a conformal mapping of points , on the unit sphere. We can add one more connection to Sec. III by introducing Cartesian coordinates x, y, and z centered on the sphere, with the x and y coordinates coincident with the X and Y
Ezra T. Newman and Richard H. Price 18

coordinates of the complex plane. It is straightforward to show that the Cartesian coordinates x , y , z of the point P1 on the unit sphere are related to the spherical coordinates , of P1 by ei cot /2 = x + iy . 1z A3

To summarize, we have found that the transformation dened by Eqs. 27 and 28 is equivalent to a proper Lorentz transformation by using three properties: i The group of proper Lorentz transformations and the conformal group of the sphere are isomorphic; ii functions of the form A1 are conformal everywhere; and iii the stereographic projection is conformal.
1

A. Wheeler, Gravitation W. H. Freeman, San Francisco, 1973 , Chap. 41. 7 The condition that X be Hermitian is removed if the coordinates t , x , y , z are taken to be complex. In this case the transformation in Eq. 11 can be replaced by X = AXB, where it is only required that Det A Det B = 1, and hence, complex Lorentz transformations can be considered. Complex spacetime coordinates are the basis of some important techniques in mathematical physics. See H. Bateman, The Mathematical Analysis of Electric and Optical Wave-Motion on the Basis of Maxwells Equations Dover, New York, 1955 ; E. T. Newman, Maxwell elds and shear-free null geodesic congruences, Class. Quantum Grav. 21, 31973221 2004 ; R. O. Hansen and E. T. Newman, A complex Minkowski space approach to twistors, Gen. Relativ. Gravit. 6 4 , 361385 1975 ; R. Penrose and W. Rindler, Spinors and Space-Time: Spinor and Twistor Methods in Space-Time Geometry Cambridge U. P., Cambridge, 2008 , Vol. 2. 8 Here we are showing a special case of a general property. A 2 2 Hermitian zero-determinant matrix, with nonnegative elements on its AA AB diagonal, can always be written in the form BA BB , and therefore as the product of a column vector with elements B and the row matrix A B that is the Hermitian conjugate of that column matrix. The validity of this claim can be demonstrated by writing A = A ei and B = B ei . The values of A and B follow immediately from the values of the diagonal elements in the 2 2 matrix, and the value of is the phase of the upper right-hand element of the matrix. A phase angle can be added to both A and B without changing the 2 2 matrix. Only the value of is xed. 9 See, for example, Ref. 4, Chap. 1, Exer. 22 and Lightman et al. Ref. 5 , Prob. 1.8. 10 The use of in place of , and the connection to the edth operator can be found, for example, in Ref. 3 E. T. Newman and R. Penrose, Note on the BondiMetznerSachs group, J. Math. Phys. 7 5 , 863870 1966 ; J. N. Goldberg, A. J. Macfarlane, E. T. Newman, F. Rohrlich, and E. C. G. Sudarshan, Spin-s spherical harmonics and , ibid. 8 11 , 2155 2161 1967 . 11 In principle it is necessary to know the relation between the directions of three photons in two different frames. Along with the condition ad bc = 1, this relation xes a , b , c , d . If we have prior knowledge that the two frames are related by a pure boost or a pure rotation, then only two photons are needed. 12 For more on the Lorentz group, see Ungar in Ref. 5. 13 The two-component column vectors are a simplied piece of the mathematical machinery of spinors More complete discussions of spinors can be found in F. A. E. Pirani, Introduction to gravitational radiation theory, in Lectures on General Relativity, edited by A. Trautman, F. A. E. Pirani, and H. Bondi Prentice-Hall, Englewood Cliffs, NJ, 1965 , and in Robert M. Wald, General Relativity U. of Chicago Press, Chicago, IL, 1984 , Chap. 13. 14 See en.wikipedia.org/wiki/Stereographic_projection .
A

Roger Penrose and Wolfgang Rindler, Spinors and Space-Time: TwoSpinor Calculus and Relativistic Fields Cambridge U. P., Cambridge, 1987 , Vol. 1. 2 See en.wikipedia.org/wiki/Mobius_transformation7 . 3 A. Held, E. T. Newman, and R. Posadas, The Lorentz group and the sphere, J. Math. Phys. 11, 31453154 1970 . 4 The simple boost transformation can be found in almost any introductory book on special relativity, such as that by Edwin F. Taylor and John A. Wheeler, Spacetime PhysicsIntroduction to Special Relativity, 2nd ed. Freeman, New York, 1992 , Chap. 1, Sec. 8. 5 The general Lorentz transformation is seldom if ever written out explicitly. The still-complicated case of the general boost can be found in John D. Jackson, Classical Electrodynamics, 3rd ed. Wiley, Hoboken, NJ, 1999 , Sec. 11.7. A pedagogical discussion of the general boost has also been given by Charles P. Frahm, Representing arbitrary boosts for undergraduates, Am. J. Phys. 47 10 , 870872 1979 . Special nontrivial cases can be found in Alan P. Lightman, William H. Press, Richard H. Price, and Saul A. Teukolsky, Problem Book in Relativity and Gravitation Princeton U. P., Princeton, NJ, 1975 , Problems 1.2, 1.26, and 1.28. A simple but relevant class of transformations is discussed by A. C. Hirshfeld and F. Metzger, A simple formula for combining rotations and Lorentz boosts, Am. J. Phys. 54 6 , 550552 1986 . The fully general case is discussed by A. L. Harvey, Lorentz transformation, ibid. 36 10 , 901904 1968 ; Abraham A. Ungar, The abstract Lorentz transformation group, ibid. 60 9 , 815828 1992 ; and H. L. Berk, K. Chaicherdsakul, and T. Udagawa, The proper Lorentz transformation operator eL = e S K: Wheres it going, whats the twist?, ibid. 69 9 , 9961009 2001 . 6 Somewhat similar notation and approaches can be found at a more advanced level in Ref. 1 and in Charles W. Misner, Kip S. Thorne, and John

19

Am. J. Phys., Vol. 78, No. 1, January 2010

Ezra T. Newman and Richard H. Price

19

Вам также может понравиться