Born coordinates

Born coordinates

/ ) are time-like curves with fixed "R".In relativistic physics, the Born coordinate chart is a coordinate chart for (part of) Minkowski spacetime, the flat spacetime of special relativity. It is often used to analyze the physical experience of observers who ride on a rigidly rotating ring or disk. This chart is often attributed to Max Born.

Langevin observers in the cylindrical chart

To motivate the Born chart, we first consider the family of Langevin observers represented in an ordinary cylindrical coordinate chart for Minkowski spacetime. The world lines of these observers form a timelike congruence which is "rigid" in the sense of having a vanishing expansion tensor. They represent observers who rotate rigidly around an axis of cylindrical symmetry.

From the line element: ds^2 = -dT^2 + dZ^2 + dR^2 + R^2 , dPhi^2, ; ; -infty < T, , Z < infty, ; 0 < R < infty, ; -pi < Phi < pi we can immediately read off a frame field representing the local Lorentz frames of stationary (inertial) observers: vec{e}_0 = partial_T, ; ; vec{e}_1 = partial_Z, ; ; vec{e}_2 = partial_R, ; ; vec{e}_3 = frac{1}{R} , partial_Phi Here, vec{e}_0 is a timelike unit vector field while the others are spacelike unit vector fields; at each event, all four are mutually orthogonal and determine the infinitesimal Lorentz frame of the static observer whose world line passes through that event.

Simultaneously boosting these frame fields in the vec{e}_3 direction, we obtain the desired frame field describing the physical experience of the Langevin observers, namely: vec{p}_0 = frac{1}{sqrt{1-omega^2 , R^2 , partial_T + frac{omega , R}{sqrt{1-omega^2 , R^2 ; frac{1}{R} partial_Phi : vec{p}_1 = partial_Z, ; ; vec{p}_2 = partial_R : vec{p}_3 = frac{1}{sqrt{1-omega^2 , R^2 ; frac{1}{R} , partial_Phi + frac{omega , R}{sqrt{1-omega^2 , R^2 , partial_TThis frame was apparently first introduced (implicitly) by Paul Langevin in 1935; its first "explicit" use appears to have been by T. A. Weber, as recently as 1997! It is defined on the region 0 < R < 1/ω; this limitation is fundamental, since near the outer boundary, the velocity of the Langevin observers approaches the speed of light.

Each integral curve of the timelike unit vector field vec{p}_0 appears in the cylindrical chart as a helix with constant radius (such as the red curve in the figure at right). Suppose we choose one Langevin observer and consider the other observers who ride on a "ring" of radius R which is rigidly rotating with angular velocity ω. Then if we take an integral curve (blue helical curve in the figure at right) of the spacelike basis vector vec{p}_3, we obtain a curve which we might hope can be interpreted as a "line of simultaneity" for the ring-riding observers. But as we see from the figure, ideal clocks carried by these ring-riding observers cannot be synchronized. This is our first hint that it is not as easy as one might expect to define a satisfactory notion of "spatial geometry" even for a "rotating ring", much less a "rotating disk!"

Computing the kinematic decomposition of the Langevin congruence, we find that the acceleration vector is: abla_{vec{p}_0} vec{p}_0 = frac{-omega^2 , R}{1- omega^2 , R^2} ; vec{p}_2 This points radially inward and it depends only on the (constant) radius of each helical world line. The "expansion tensor" vanishes identically, which means that nearby Langevin observers maintain constant distance from each other. The "vorticity vector" is: vec{Omega} = frac{omega}{1 - omega^2 , R^2} ; vec{p}_1 which is parallel to the axis of symmetry. This means that the world lines of the nearest neighbors of each Langevin observer are "twisting about its own world line", as suggested by the figure at right. This is a kind of "local notion" of "swirling" or vorticity.

In contrast, note that projecting the helices onto any one of the spatial hyperslices T=T_0 orthogonal to the world lines of the static observers gives a circle, which is of course a closed curve. Even better, the coordinate basis vector partial_Phi is a "spacelike" Killing vector field whose integral curves are closed spacelike curves (circles, in fact), which moreover degenerate to zero length closed curves on the axis R = 0. This expresses the fact that our spacetime exhibits "cylindrical symmetry", and also exhibits a kind of "global notion" of the rotation of our Langevin observers.

In the figure, the magenta curve shows how the spatial vectors vec{p}_2, ; vec{p}_3 are spinning about vec{p}_1 (which is suppressed in the figure since the Z coordinate is inessential). That is, the vectors vec{p}_2, ; vec{p}_3 are not Fermi-Walker transported along the world line, so the Langevin frame is "spinning" as well as non-inertial. In other words, in our straightforward derivation of the Langevin frame, we kept the frame aligned with the radial coordinate basis vector partial_R. By introducing a constant rate rotation of the frame carried by each Langevin observer about vec{p}_1, we could, if we wished "despin" our frame to obtain a gyrostabilized version.

Transforming to the Born chart

To obtain the Born chart, we straighten out the helical world lines of the Langevin observers using the simple coordinate transformation: t = T, ; ; z = Z, ; ; r = R, ; ; phi = Phi - omega , TThe new line element is: ds^2 = -left( 1- omega^2 , r^2 ight) , dt^2 + 2 , omega , r^2 , dt , dphi + dz^2 + dr^2 + r^2 , dphi^2 :-infty < t, , z < infty, 0 < r < frac{1}{omega}, ; -pi < phi < piNotice the "cross-terms" involving dt , dphi, which show that the Born chart is not an "orthogonal" coordinate chart. The Born coordinates are also sometimes referred to as "rotating cylindrical coordinates".

In the new chart, the world lines of the Langevin observers appear as vertical straight lines. Indeed, we can easily transform the four vector fields making up the Langevin frame into the new chart. We obtain: vec{p}_0 = frac{1}{sqrt{1-omega^2 , r^2 , partial_t: vec{p}_1 = partial_z, ; ; vec{p}_2 = partial_r : vec{p}_3 = frac{sqrt{1-omega^2 , r^2{r} , partial_phi + frac{omega , r}{sqrt{1-omega^2 , r^2 , partial_tThese are exactly the same vector fields as before--- they are now simply represented in a different coordinate chart!

Needless to say, in the process of "unwinding" the world lines of the Langevin observers, which appear as helices in the cylindrical chart, we "wound up" the world lines of the static observers, which now appear as helices in the Born chart! Note too that, like the Langevin frame, the Born chart is only defined on the region 0 < r < 1/ω.

If we recompute the "kinematic decomposition" of the Langevin observers, that is of the timelike congruence vec{p}_0 = frac{1}{sqrt{1-omega^2 , r^2 , partial_t, we will of course obtain the same answer that we did before, only expressed in terms of the new chart. Specifically, the acceleration vector is: abla_{vec{p}_0} , vec{p}_0 = frac{-omega^2 ,r}{1 - omega^2 , r^2} , vec{p}_2the expansion tensor vanishes, and the vorticity vector is: vec{Omega} = frac{omega}{1-omega^2 , r^2} ; vec{p}_1

The dual covector field of the timelike unit vector field in any frame field represents infinitesimal spatial hyperslices. However, the Frobenius integrability theorem gives a strong restriction on whether or not these spatial hyperplane elements can be "knit together" to form a family of spatial hypersurfaces which are everywhere orthogonal to the world lines of the congruence. Indeed, it turns out that this is possible, in which case we say the congruence is "hypersurface orthogonal", if and only if "the vorticity vector vanishes identically". Thus, while the static observers in the cylindrical chart admits a unique family of "orthogonal hyperslices" T=T_0, "the Langevin observers admit no such hyperslices". In particular, the spatial surfaces t=t_0 in the Born chart are "orthogonal to the static observers, not to the Langevin observers". This is our second (and much more pointed) indication that defining "the spatial geometry of a rotating disk" is not as simple as one might expect.

To better understand this crucial point, consider integral curves of the third Langevin frame vector:vec{p}_3 = sqrt{1-omega^2 , r^2} , frac{1}{r} , partial_phi + frac{omega , r}{ sqrt{1-omega^2 , r^2 , partial_twhich pass through the radius phi=0, , t=0. (For convenience, we will suppress the inessential coordinate z from our discussion.) These curves lie in the surface :phi + omega , t - frac{t}{omega , r^2} = 0, ; ; -pi < phi < pi shown in the figure. We would like to regard this as a "space at a time" for our Langevin observers. But two things go wrong.

First, the Frobenius theorem tells us that vec{p}_2, , vec{p}_3 are tangent to no spatial hyperslice whatever. Indeed, except on the initial radius, the vectors vec{p}_2 "do not lie in our slice". Thus, while we found a spatial hypersurface, it is orthogonal to the world lines of only "some" our Langevin observers. Because the obstruction from the Frobenius theorem can be understood in terms of the failure of the vector fields vec{p}_2, , vec{p}_3 to form a Lie algebra, this obstruction is differential, in fact Lie theoretic. That is, it is a kind of "infinitesimal obstruction" to the existence of a satisfactory notion of spatial hyperslices for our rotating observers.

Second, as the figure shows, our attempted hyperslice would lead to a "discontinuous" notion of "time" due to the "jumps" in the integral curves (shown as a coral colored discontinuity). Alternatively, we could try to use a multivalued time. Neither of these alternatives seems very attractive! This is evidently a "global obstruction". It is of course a consequence of our inability to synchronize the clocks of the Langevin observers riding even a single "ring"---say the rim of a disk--- much less an entire "disk".

The Sagnac effect

Imagine that we have fastened a fiber-optic cable around the circumference of a ring which is rotating with steady angular velocity ω. We wish to compute the round trip travel time, as measured by a ring-riding observer, for a laser pulse sent clockwise and counterclockwise around the cable. For simplicity, we will ignore the fact that light travels through a fiber optic cable at somewhat less than the speed of light in a vacuum, and will pretend that the world line of our laser pulse is a null curve (but certainly not a null "geodesic"!).

In the Born line element, let us put ds = dz = dr = 0. This gives: (1 - omega^2 , r_0^2) , dt^2 = 2 omega , r_0^2 , dt , dphi + r_0^2 , dphi^2 or: dt = frac{r_0 , dphi}{1 pm omega , r_0} We obtain for the round trip travel time: Delta t_+ = frac{2 pi r_0}{1 + omega , r_0}, ; ; Delta t_- = frac{2 pi r_0}{1 - omega , r_0} Putting delta = frac{Delta t_+ - Delta t_-}{2 , pi , r}, we find omega = frac{-1 + sqrt{1+delta^2{delta , r} so that the ring-riding observers can determine the angular velocity of the ring (as measured by a static observer) from the difference between clockwise and counterclockwise travel times. This is known as the Sagnac effect. It is evidently a "global effect".

Null Geodesics

We wish to compare the appearance of null geodesics in the cylindrical chart and the Born chart.

In the cylindrical chart, the geodesic equations read: ddot{T} = 0, ; ; ddot{Z} = 0, ; ; ddot{R} - R , dot{Phi}^2 = 0, ; ; ddot{Phi} + frac{2}{R} , dot{Phi} , dot{R} = 0We immediately obtain the first integrals:dot{T} = E,; ; dot{Z} = P, ; ; dot{Phi} = frac{L}{R^2} Plugging these into the expression obtained from the line element by setting ds^2 = 0, we obtain: dot{R}^2 = E^2 - P^2 - frac{L^2}{R^2} from which we see that the "minimal radius" of a null geodesic is given by:R_0 = frac{L}{sqrt{E^2-P^2

We can now solve to obtain the null geodesics as curves parameterized by an affine parameter, as follows::R = sqrt{(E^2-P^2) , s^2 + 2 , s , sqrt{ (E^2-P^2) ; R_0^2 - L^2 + R_0^2 :T = T_0 + E , s, ; ; Z = Z_0 + P :Phi = Phi_0 + operatorname{arctan} left( frac{ 2 , sqrt{(E^2-P^2), R_0^2-L^2} + 2 , (E^2-P^2) ,s}{2 , L} ight) : ; ; ; - operatorname{arctan} left( frac{ sqrt{(E^2-P^2), R_0^2-L^2} }{L} ight)

More useful for our purposes is the observation that the "trajectory" of a null geodesic (its projection into any spatial hyperslice T=T_0) is of course a straight line, given by: R = R_0 , sec (Phi-Phi_0 )To obtain the minimal radius of the line through two points (on the same side of the point of closest approach to the origin), we solve: R_1 = R_0 , sec (Phi_1-Phi_0 ), ; ; R_2 = R_0 , sec (Phi_2-Phi_0 )which gives: R_0 = frac{R_1 , R_2 , |sin (Phi_2-Phi_1){sqrt{R_1^2 - 2 , R_1 , R_2 , cos(Phi_2-Phi_1) + R_2^2

Now consider the simplest case, the "radial null geodesics". An outward bound radial null geodesic may be written in the form :T = s , ; Z = Z_0, ; R = s :Phi = omega , R_0 Transforming to the Born chart, we find that the trajectory can be written as: r = frac{phi}{omega} - r_0 Similarly for inward bound radial null geodesics. The tracks turn out to appear slightly bent in the Born chart (see figure at right). (We will see in a later section that in the Born chart, we cannot properly refer to these "tracks" as "projections", however.)

Notice that, just as a duck hunter would expect, to send a laser pulse toward the stationary observer at R = 0, the Langevin obsevers have to "aim slightly ahead" to correct for their own motion. Turning things around, to send a laser pulse toward a Langevin observer riding a counterclockwise rotating ring, the central observer has to aim, not at this observer's current position, but at the position at which he will arrive just in time to intercept the signal. These families of inward and outward bound radial null geodesics represent very different curves in spacetime, but their projections do agree.

Similarly, null geodesics between ring-riding Langevin observers appear slightly bent inward in the Born chart. To see this, write the equation of a null geodesic in the cylindrical chart in the form:T = R_0 , an (Phi):R = R_0 , sec(Phi) Transforming to Born coordinates, we obtain the equations: t = r_0 , an (phi + omega , t):r = r_0 , sec (phi + omega , t)Eliminating φ gives :r = sqrt{r_0^2 + t^2}which shows that the geodesic does indeed appear to bend inward. We also find that: phi = -omega , sqrt{r^2 - r_0^2} + operatorname{arctan}(t/r_0)This completes the description of the appearance of null geodesics in the Born chart, since every null geodesic is either radial or else has some point of closest approach to the axis of cylindrical symmetry.

Note (see figure) that a ring-riding observer trying to send a laser pulse to another ring-riding observer must aim slightly ahead of his angular coordinate as given in the Born chart, in order to compensate for the rotational motion of the target. Note too that the picture presented here is fully compatible with our expectation (see appearance of the night sky) that a moving observer will see the apparent position of other objects on his celestial sphere to be "displaced" toward the direction of his motion.

Radar distance in the large

Even in flat spacetime, it turns out that accelerating observers (even linearly accelerating observers; see Rindler coordinates) can employ various "distinct" but operationally significant notions of distance. Perhaps the simplest of these is "radar distance".

Consider how a static observer at R=0 might determine his distance to a ring riding observer at R = R0. At event C he sends a radar pulse toward the ring, which strikes the world line of a ring-riding observer at A′ and then returns to the central observer at event C″. (See the "right" hand diagram in the figure at right.) He then divides the elapsed time (as measured by an ideal clock he carries) by two. It is not hard to see that he obtains for this distance simply R0 (in the cylindrical chart), or r0 (in the Born chart).

Similarly, a ring-riding observer can determine his distance to the central observer by sending a radar pulse, at event A toward the central observer, which strikes his world line at event C′ and returns to the ring-riding observer at event A″. (See the left hand diagram in the figure at right.) It is not hard to see that he obtains for this distance frac{R_0}{sqrt{1-omega^2 , R_0^2 (in the cylindrical chart) or frac{r_0}{sqrt{1-omega^2 , r_0^2 (in the Born chart), a result which is somewhat larger than the one obtained by the central observer. Thus, while radar distance has a simple operational significance, "it is not even symmetric".

Just to drive home this crucial point, let us compare the radar distances obtained by two ring-riding observers with radial coordinate R = R0. In the left hand diagram at the figure to the left, we can write the coordinates of event A as :T = 0, ; Z = 0, ; X = R_0, Y=0and we can write the coordinates of event B′ as :T=2 , R_0 , sin left( frac{Phi}{2} ight), ; Z=0, : X = R_0 cos(Phi), ; Y=R_0 sin(Phi)Writing the unknown elapsed proper time as Delta s, we now write the coordinates of event A″ as:T=frac{Delta s}{sqrt{1-omega^2 , R_0^2, ; Z=0 :X = R_0 cos left( frac{omega , Delta s}{sqrt{1-omega^2 , R_0^2 ight), ; Y=R_0 sin left( frac{omega , Delta s}{sqrt{1-omega^2 , R_0^2 ight)By requiring that the line segments connecting these events be null, we obtain an equation which in principle we can solve for Δ s. It turns out that this procedure gives a rather complicated nonlinear equation, so we simply present some representative numerical results. With R0 = 1, Φ = π/2, and ω = 1/10, we find that the radar distance from A to B is about 1.308, while the distance from B to A is about 1.505. As ω tends to zero, both results tend toward sqrt{2} approx 1.414.

Despite these possibly discouraging discrepancies, it is by no means impossible to devise a coordinate chart which is adapted to describing the physical experience of a "single" Langevin observer, or even a single arbitrarily accelerating observer in Minkowski spacetime. Pauri and Vallisneri have adapted the "Märzke-Wheeler clock synchronization procedure" to devise adapted coordinates they call "Märzke-Wheeler coordinates" (see the paper cited below). In the case of steady circular motion, this chart is in fact very closely related to the notion of radar distance "in the large" from a given Langevin observer.

Radar distance in the small

As was mentioned above, for various reasons the family of Langevin observers admits no family of orthogonal hyperslices. Therefore "these observers simply cannot be associated with any slicing of spacetime into a family of successive "constant time slices"."

However, because the Langevin congruence is "stationary", we can imagine replacing each "world line" in this congruence by a "point". That is, we can consider the quotient space of Minkowski spacetime (or rather, the region 0 < R < 1/ω) by the Langevin congruence, which is a three-dimensional topological manifold. Even better, we can place a Riemannian metric on this quotient manifold, turning it into a three dimensional Riemannian manifold, in such a way that the metric has a simple operational significance.

To see this, consider the Born line element: ds^2 = -(1-omega^2 , r^2 ) , dt^2 + 2 , omega , r^2 , dt , dphi + dz^2 + dr^2 + r^2 , dphi^2 : -infty < t, z < infty, ; ; 0 < r < frac{1}{omega}, ; ; -pi < phi < piSetting ds2 = 0 and solving for dt we obtain: dt_+ = frac{-omega , r^2 + sqrt{(1-omega^2 , r^2) ; (dz^2+dr^2) + r^2 , dphi^2{1-omega^2 , r^2} : dt_- = frac{-omega , r^2 - sqrt{(1-omega^2 , r^2) ; (dz^2+dr^2) + r^2 , dphi^2{1-omega^2 , r^2} The elapsed "proper time" for a roundtrip radar blip emitted by a Langevin observer is then: sqrt{1 - omega^2 , r^2} ; frac{dt_+ - dt_-}{2} = sqrt{dz^2 + dr^2 + frac{r^2 , dphi^2}{1-omega^2 , r^2} }Therefore, in our quotient manifold, the Riemannian line element:dsigma^2 = dz^2 + dr^2 + frac{r^2 , dphi^2}{1-omega^2 , r^2}:-infty < t < infty, ; ; 0 < r < frac{1}{omega}, ; ; -pi < phi < picorresponds to distance between "infinitesimally nearby Langevin observers". We will call it the "Langevin-Landau-Lifschitz" metric, and we can call this notion of distance "radar distance "in the small".

This metric was first given by Langevin, but the interpretation in terms of radar distance "in the small" is due to Lev Landau and Evgeny Lifshitz, who generalized the construction to work for the quotient of any Lorentzian manifold by a "stationary" timelike congruence.

If we adopt the coframe: heta^hat{1} = dz, ; ; heta^hat{2} = dr, ; ; heta^hat{3} = frac{r , dphi}{sqrt{1- omega^2 , r^2we can easily compute the Riemannian curvature tensor of our three dimensional quotient manifold. It has only one independent nontrivial component,: {R^{hat{2}_{hat{3}hat{2}hat{3 = frac{-3 , omega^2}{(1 - omega^2 r^2)^2} = -3 , omega^2 ; left( 1 + 2 , omega^2 , r^2 ight) + O(omega^6 , r^6) Thus, in some sense, "the geometry of a rotating disk is curved", as Theodor Kaluza claimed (without proof) as early as 1910. In fact, to fourth order in ω it has the geometry of the hyperbolic plane, just as Kaluza claimed.

"Warning:" as we have seen, there are many possible notions of distance which can be employed by Langevin observers riding on a rigidly rotating disk, so statements referring to "the geometry of a rotating disk" always require careful qualification.

To drive home this important point, let us use the Landau-Lifschitz metric to compute the distance between a Langevin observer riding a ring with radius R0 and a central static observer. To do this, we need only integrate our line element over the appropriate null geodesic track. From our earlier work, we see that we must plug:dr = frac{dphi}{omega}into our line element and integrate. This gives:Delta = int_{r_0}^0 frac{dr}{sqrt{1-omega^2 , r^2 = frac{arcsin (omega r_0)}{omega} = r_0 + frac{omega^2 , r_0^3}{6} + O(r_0^4)Because we are now dealing with a Riemannian metric, this notion of distance is of course "symmetric" under interchanging the two observers, unlike radar distance "in the large". The values given by this notion are intermediate between the radar distances computed in the previous section. For example, for r0 = 1, ω = 1/2, we find approximately Δ = 1.047, which can be compared with 1.155 for the distance from the ring-riding observer to the central observer, or 1 for the central observer to the ring-riding observer. Also, because up to second order the Landau-Lifschitz metric agrees with radar distance "in the large", we see that the curvature tensor we just computed does have operational significance: while radar distance "in the large" between pairs of Langevin observers is certainly "not a Riemannian notion of distance", the distance between pairs of "nearby" Langevin observers does correspond to a Riemannian distance, given by the Langevin-Landau-Lifschitz metric. (In the felicitous phrase of Howard Percy Robertson, this is "kinematics im kleinem".)

One way to see that all reasonable notions of spatial distance for our Langevin observers agree for nearby observers is to show, following Nathan Rosen, that for any one Langevin observer, an "instantaneously comoving inertial observer" will also obtain the distances given by the Langevin-Landau-Lifschitz metric, for very small distances.

ummary

Observers riding on a rigidly rotating disk will conclude from measurements of "small distances" between themselves that the geometry of the disk is non-Euclidean. Regardless of which method they use, they will conclude that "the geometry is well approximated by a certain Riemannian metric", namely the Langevin-Landau-Lifschitz metric. This is in turn very well approximated by the geometry of the "hyperbolic plane" (with the constant negative curvature -3 ω2). But if these observers measure larger distances, they will obtain "different" results, depending upon which method of measurement they use! In all such cases, however, they will most likely obtain results which are "inconsistent with any Riemannian metric". In particular, if they use the simplest notion of distance, radar distance, owing to various effects such as the "asymmetry" already noted, they will conclude that "the "geometry" of the disk is not only non-Euclidean, it is non-Riemannian."

ee also

*Ehrenfest paradox, for a sometimes controversial topic often studied using the Born chart.
*Fibre optic gyroscope
*Rindler coordinates, for another useful coordinate chart adapted to another important family of accelerated observers in Minkowski spacetime; this article also emphasizes the existence of distinct notions of distance which may be employed by such observers.
*Sagnac effect

References

A few papers of historical interest:

*cite journal | author=Born, M. | authorlink=Max Born | title=Die Theorie des starren Elektrons in der Kinematik des Relativitäts-Prinzipes | journal=Ann. Phys. Lpz. | year=1909 | volume=30 | pages=1
*cite journal | author=Ehrenfest, P. | authorlink=Paul Ehrenfest | title=Gleichförmige Rotation starrer Körper und Relativitätstheorie | journal=Phys. Zeitschrift | year=1909 | volume=10 | pages=918
*cite journal | author=Planck, M. | authorlink=Max Planck | title=Gleichförmige Rotation und Lorentz-Kontraktion | journal=Phys. Zeitschrift | year=1910 | volume=11 | pages=294
*cite journal | author=Einstein, A. | authorlink=Albert Einstein | title=Zum Ehrenfesten Paradoxon | journal=Phys. Zeitschrift | year=1911 | volume=12 | pages=509
*cite journal | author=Sagnac, M. G. | authorlink=Georges Sagnac | title=L'éther lumineux démontré par l'effet du vent relatif d'éther dans un interféromètre en rotation uniforme | journal=C. R. Acad. Sci. Paris | year=1913 | volume=157 | pages=708
*cite journal | author=Langevin, P. | authorlink=Paul Langevin | title=Remarques au sujet de la Note de Prunier | journal=C. R. Acad. Sci. Paris | year=1935 | volume=200 | pages=48

A few classic references:
*cite journal | author=Grøn, Ø. | title=Relativistic description of a rotating disk | journal=Amer. J. Phys. | year=1975 | volume=43 | pages=869–876 | doi=10.1119/1.9969
*cite book | author=Landau, L. D. & Lifschitz, E. M. | title=The Classical Theory of Fields (4th ed.) | location=London | publisher=Butterworth-Heinemann | year=1980 | id=ISBN 0-7506-2768-9 See "Section 84" for the Landau-Lifschitz metric on the quotient of a Lorentzian manifold by a "stationary" congruence; see the problem at the end of "Section 89" for the application to Langevin observers.

Selected recent sources:
*cite book | author=Rizzi, G. ; & Ruggiero, M. L. | title=Relativity in Rotating Frames | location=Dordrecht | publisher= Kluwer | year=2004 | id=ISBN 1-4020-1805-3 This book contains a valuable historical survey by Øyvind Grøn and some other papers on the Ehrenfest paradox and related controversies and a paper by Lluis Bel discussing the Langevin congruence. Hundreds of additional references may be found in this book.
*cite journal | author=Pauri, Massimo; & Vallisneri, Michele | title=Märzke-Wheeler coordinates for accelerated observers in special relativity | journal=Found. Phys. Lett. | year=2000 | volume=13 | pages=401–425 | doi=10.1023/A:1007861914639 Studies a coordinate chart constructed using "radar distance "in the large" from a single Langevin observer. See also the [http://www.arxiv.org/abs/gr-qc/0006095 eprint version] .

External links

* [http://www2.corepower.com:8080/~relfaq/rigid_disk.html The Rigid Rotating Disk in Relativity] , by Michael Weiss (1995), from the "sci.physics FAQ".


Wikimedia Foundation. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Born-Oppenheimer approximation — In quantum chemistry, the computation of the energy and wavefunction of an average size molecule is a formidable task that is alleviated by the Born Oppenheimer (BO) approximation. For instance the benzene molecule consists of 12 nuclei and 42… …   Wikipedia

  • Rindler coordinates — In relativistic physics, the Rindler coordinate chart is an important and useful coordinate chart representing part of flat spacetime, also called the Minkowski vacuum. The Rindler chart was introduced by Wolfgang Rindler. The Rindler coordinate… …   Wikipedia

  • Ehrenfest paradox — The Ehrenfest paradox concerns the rotation of a rigid disc in the theory of relativity.In its original formulation as presented by Paul Ehrenfest 1909 in the Physikalische Zeitschrift, it discusses an ideally rigid cylinder that is made to… …   Wikipedia

  • Criticism of relativity theory — Criticism of Albert Einstein s theory of relativity was mainly expressed in the early years after its publication on a scientific, pseudoscientific, philosophical, or ideological basis. Reasons for criticism were, for example, alternative… …   Wikipedia

  • Coordenadas de Rindler — En física relativista, la carta de coordenadas de Rindler es una importante y útil carta de coordenadas representando parte del espaciotiempo plano, también llamado el vacío de Minkowski. La carta de Rindler fue introducida por Wolfgang Rindler.… …   Wikipedia Español

  • Sagnac effect — The Sagnac effect (also called Sagnac Interference), named after French physicist Georges Sagnac, is a phenomenon encountered in interferometry that is elicited by rotation. The Sagnac effect manifests itself in a setup called ring interferometry …   Wikipedia

  • Beveridge, Victoria — Beveridge Victoria Hunters Tryste Inn, 1845 …   Wikipedia

  • Paul Langevin — Infobox Scientist name = Paul Langevin |224px image width = 224px caption = Paul Langevin (1872 1946) birth date = birth date|1872|1|23|mf=y birth place = Paris, France death date = death date and age|df=yes|1946|12|19|1872|1|23 death place =… …   Wikipedia

  • Michelson–Gale–Pearson experiment — The Michelson–Gale–Pearson experiment (1925) is a modified version of the Michelson Morley experiment and the Sagnac Interferometer. It measured the Sagnac effect due to Earth s rotation, and thus tests the theories of special relativity and… …   Wikipedia

  • Cornwall, New York — Coordinates: 41°26′04″N 74°02′09″W / 41.43444°N 74.03583°W / 41.43444; 74.03583 …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”