Tuesday, August 29, 2023

The metric around a rotating disk

UPDATE September 3, 2023: Our observation today solves the mystery why the August 10, 2023 calculation had "radial corrections" but the calculation in this post does not have.

----

Let us use our Schwarzschild-based method to determine the metric perturbation from a lightweight, uniform, slowly rotating disk, in its plane, far away from the disk.

In this blog post we set c = 1, to simplify the notation.


                        M = static disk mass
                        K  = kinetic energy of disk
                        I   = moment of inertia of disk
                        J   = angular momentum of disk
                        R₀ = disk radius
                        ρ  = static disk mass / area

                   <------ ω angular velocity

      dm'''                                 dm''
    •                                       • 



                         center of the disk
                        ×     R = distance (×, dm)
                        |   \    α = angle (×, m) vs. (×, dm)
                        |      \   
                        |         \             ^   v
                                              /  
    •                                       • 
      dm'                                   dm

                                             r = distance (m, dm)




                        |    β = angle (m, ×) vs. (m, dm)
                        |  /
                        |/
                        • m observer (or a test mass)

                R_m = distance (m, ×)

  ^  y
  |
   -----> x


The diagram is much like the one on August 10, 2023, but we switched the x and y axes. We have a part of disk, dm, moving with a velocity vector v. The disk is very lightweight and rotates very slowly.

Mirror images: dm' is the mirror image of dm against the line (m, ×), dm''' is the mirror image against ×, and dm'' its image against (m, ×).

We want to determine the metric at an observer m, which is far away from the rotating disk, i.e., the distance (m, ×) is much larger than the disk radius R₀.

The Schwarzschild metric associated with dm is moving at a velocity v obliquely relative to the observer m. The calculations may be somewhat complicated.


Lorentz transformations



















The boost matrix B(v) is for a frame which moves with a velocity (v_x, v_y) relative to the laboratory frame.

Let us have a position 3-vector

                  t
        r  =    x
                  y

in the laboratory frame. The coordinates in the moving frame are

                  t'
        r'  =   x'   =  B(v) r,
                  y'

where the multiplication of B(v) and r is the matrix multiplication.

We obtain the inverse Lorentz transformation by flipping the signs of v_x and v_y.


The metric perturbation caused by a part dm


Let us have a part dm moving with a velocity vector (v_x, v_y). We write the line element of the Schwarzschild metric in the moving frame, in terms of coordinates of the laboratory frame. The metric in the moving frame is

        dτ² = (-1 + r_s / r') dt'² + (1 + r_s / r') dr'² + (r' dφ')²,

in polar coordinates, if we assume that r_s is small. We have

       dr'  =  -sin(β') dx'  -  cos(β') dy'.

The Minkowski metric maps onto itself in a Lorentz transformation. We can concentrate on how the perturbation is mapped. In the moving frame, the perturbation to the Minkowski metric is

       r_s / r'  *  dt'² 

    + r_s / r'  *  (- sin(β') dx'  -  cos(β') dy' )².

Let us try to write the perturbation in terms of dt, dx, dy. We can drop the terms ~ v³ because we assume that v is small. We can approximate γ = 1 + v² / 2. We get:

       dt'   = (1 + v² / 2) dt
               - v_x dx
               - v_y dy,

       dx'  =  v_x dt 
               + (1  +  v_x² / 2) dx
               + (v_x  v_y / 2) dy,

       dy'  =  v_y dt
               + (v_x  v_y / 2) dx
               + (1  +  v_y² / 2) dy.

Then

       dt'² =  dt²  +  v² dt²  +  v_x² dx²  +  v_y² dy²

                - 2 v_x  dt dx  -  2 v_y  dt dy

               + 2 v_x  v_y  dx dy,

and

       ( sin(β') dx'  +  cos(β') dy' )²
 
        =

          sin²(β')  *  (v_x² dt² + dx² + v_x² dx²

                         + 2 v_x dt dx  +  v_x v_y dx dy)

       + cos²(β')  *  (v_y² dt² + dy² + v_y² dy²

                         + 2 v_y dt dy  +  v_x v_y dx dy)

       + 2 sin(β') cos(β')

           * (  v_x  v_y dt² + v_x dt dy + v_y dt dx
   
               + v_x  v_y / 2 * dx²

               + (1 + v² / 2) * dx dy

               + v_x  v_y / 2 * dy²
               ).

We want to integrate, but what is the role of β' in the moving frame versus β in the laboratory frame? Also, r_s / r' is calculated in the moving frame, not in the laboratory frame.

The length contraction in the y direction is a factor

       1 - v_y² / 2,

and in the x direction

       1 - v_x² / 2.

Approximately,

       sin(β')  =  sin(β) (1 - v_x² + v_y²),

       cos(β') 
         = sqrt(1 - sin²(β) * (1 - v_x² + v_y²)²)
 
         = sqrt(1 - sin²(β) * (1 - 2 v_x² + 2 v_y²))

         = cos(β) + sin²(β) * (-v_x² + v_y²).

We can ignore the correction sin²(β) (-v_x² + v_y²) because

       r_s / r * sin²(β)  ~  1 / R_m³,

and we are only interested in metric perturbations which are ~ 1 / R_m or ~ 1 / R_m².


                           \  |  /
                             \|/
                      ------- • dm ---
                             /|\
                           /  |  \  squeezed field lines
  • m test mass
                            ---> v


How should we map r' to r? If dm is moving straight away from the observer m at a velocity v, then the moving frame, and the field of dm, appears squeezed in the direction of v. 

Simultaneousness is problematic. We are interested in the metric at the observer m at the laboratory time t. The laboratory observer thinks that dm is at a specific location X at that time t. But these events are not simultaneous in the moving frame.

Let us assume that dm carries a 1 meter long ruler with it, whose far end at the laboratory time t is at the observer m. The field at the end of the ruler is, of course, the Schwarzschild metric at the distance 1 meter, as seen by dm. In this case, we have

       r  =  (1 - v² / 2) r',

or

       r'  =  (1 + v² / 2) r.

If the ruler is oblique, then

       r'²  =  (1 + v²) r₁² + r₂²

             =  r² + v² r₁²,

where r₁ and r₂ are the components of r along v and normal to v.

We obtain:

       r'  =  r * sqrt(1 + v² r₁² / r²)

            =  r * ( 1 + 1/2 v² sin²(α + β) ).


Integrals of the type 1 / r over a circle: the generalized shell theorem


Below we need to integrate several terms of the type

       1 / r * dm

over a full circle, where r is the (large) distance to the observer m, and dm = ρ R₀ dα. The mass of the circle per a unit length of the circumference is ρ and its radius is R₀. We will study the sum of 1 / r for dm and dm'', as in the diagram at the start of this blog post.

Let us scale the diagram above to make calculations simpler. We keep ρ constant. If we scale by some factor X, then the length of the circle is scaled by X and 1 / r is scaled by 1 / X. The value of the integral does not change.

Let the coordinates of the observer be (0, 0), The radius of the circle is R₁ = R₀ / R_m, and its center is at (0, 1).

The coordinates of dm are (x, 1 - y), and the coordinates of dm'' are (x, 1 + y).

We will study the value of

       1 / r + 1 / r''  =  1 / sqrt(x² + 1 -  2y + y²)

                              + 1 / sqrt(x² + 1 + 2y + y²).

The Taylor series of 1 / sqrt(1 + a) is

       1  -  1/2 a  +  1 / 2! * 3/4 a²  +  ...

where a = x² +- 2y + y². We obtain

       1 / r + 1 / r'' = 2  -  1/2 (2 x² + 2 y²)

                                   +  3/8 (8 y²) + ...

where we dropped all terms with a power 3 or more. Let us integrate over the semicircle for which y < 0. The integral for the term 2 is

       2 π R₁  = 2 π R₀ * 1 / R_m,

while the integrals for other terms are 

       ~ 1 / R_m³.

We are only interested in terms ~ 1 / R_m and ~ 1 / R_m², and can ignore ~ 1 / R_m³ terms.

The integral can be interpreted this way: if we are far from the circle, we can pretend that all the parts of the circle are at the distance of the center of the circle. It is like the shell theorem of Newton.

What about an integral

       sin²(α) / r * dm ?

We have to integrate

       sin²(α) * 2

over the lower semicircle whose radius is R₀ / R_m. The integral is

       1/2 * 2 π R₀ * 1 / R_m.

Again, the integral is as if all the parts of the circle were at the distance of the center of the circle.

Generalized Shell Theorem. If we have to integrate

       f(α) / r,
or
       f(α) / r²

over a full circle, where the distance r is measured from a point P far away from the circle relative to the radius, and P is at the angle α = 0 relative to the center of the circle, and

       f(π - α)      =  f(α)        (symmetry),

       f(π + α)     =  f(-α)       (symmetry),

for any 0 ≤ α ≤ π / 2, and for any α, holds

       f(2 π + α)  =  f(α)       (periodicity),

then we can pretend that r = the distance to the center of the circle.

The error in the approximation is at most ~ 1 / r³. The conditions above say that the function f is symmetric relative to a diameter of the circle where the diameter is normal to the angle 0, and that f is periodic.



                             /  α = angle vs. (×, P)
                    ___ /_
                 /             \             r
                |       ×       | -----------------------------  P
                 \_______/

                  <------->
               symmetry


Proof. We can calculate as above in this section, that the integral of f(α) / r over the whole circle is

     π / 2
       ∫   f(α) * 2 * R₀ / R_m * dα.
   -π / 2

By the symmetry of f, this is equal to

       3/2 π
       ∫        f(α) R₀ / R_m * dα.
   -π / 2

The periodicity of f implies that the above is equal to

        2 π
        ∫   f(α) / R_m * R₀ dα,
      0

which is the integral of f(α) / R_m over the whole circle.

The proof for f(α) / r² is similar to f(α) / r. Q.E.D.


The integral of the dt'² term


           r_s / r   *   ( 1 + 1/2 v² sin²(α + β) )
     *
           ( dt² + v² dt² + v_x² dx² + v_y² dy²

                 - 2 v_x  dt dx  - 2 v_y  dt dy

                 + 2 v_x  v_y  dx dy),

      v      =  ω R,

      v_x  =  ω R cos(α),

      v_y  =  ω R sin(α),

      r  =  sqrt( R² sin²(α)  +  (R_m - R cos(α))² )

           =  R sin(α) / sin(β).

We must integrate the formula above for all dm. There should be a lot of canceling of different terms for dm and its mirror images dm' and dm'', just as we calculated on August 10, 2023.

The term for dt² is the attractive gravity force of the static mass of the disk. The mixed terms dt dx and dt dy are "magnetic" forces.

1.      r_s / r * dt² :   the integral is essentially 2 G M / R_m * dt², or the Schwarzschild / newtonian gravity attraction of the mass M of the static disk. We used the generalized shell theorem.

2.      r_s / r * v² dt² :   the integral is 2X the gravity of the kinetic energy K of the disk. We can use the generalized shell theorem.

3.      r_s / r * 1/2 v² sin²(α + β) dt² :   if R_m / R₀ is large, then β is small and the integral is 0.5X the gravity of the kinetic energy K of the disk. We can use the generalized shell theorem.

4.      r_s / r * v_x² dx² :    2 G K / R_m  *  dx².

5.      r_s / r * v_y² dy² :    2 G K / R_m  *  dy².

6.      r_s / r * -2 v_x  dt dx :    the symmetry between dm and dm'' almost cancels this term, but not entirely, because dm is closer.  This is the "dm moves to the right" term in our August 10, 2023 calculation. The sum is

       dt dx   *   -2 G dm / (R_m / cos(β))²

                   *    2 R * cos(α) cos(β) * 2 v cos(α).

The value of cos³(β) differs from 1 by ~ 1 / R_m². That makes ~ 1 / R_m⁴ and we can ignore cos³(β). The mass element is

       dm = ρ dR R dα.

The integral of cos²(α) over 0 < α < π / 2 is π / 4. We have to multiply by 2 to take into account the left side of the disk. We have to integrate 

       dt dx *

        R₀
       ∫    -2 G / R_m² * π / 2 * 2 R * 2 ω R ρ R dR
      0

      =  dt dx * -G / R_m² * ρ π R² * R² ω 

      =  dt dx * -G / R_m² * M * R² ω 

      =  dt dx * -G / R_m² * 2 I ω 

      =  dt dx * G / R_m² * -2 J.

7.      r_s / r * 2 v_y dt dy :    the symmetry between dm' and dm makes this zero.

8.      r_s / r * 2 v_x v_y dx dy :     the symmetry between dm' and dm makes this zero.


Summary: We have to add 2.5 K to M to get the gravity attraction. There is some stretching in the x and y metrics. There is a magnetic term which "corresponds to" -2 J.


The integral of the dx'² term


       r_s / r  *  sin²(β) (1 - 2 v_x² + 2 v_y²)

       * (v_x² dt² + dx² + v_x² dx²

           + 2 v_x dt dx + v_x v_y dx dy).

We can ignore this because

       1 / r * sin²(β) ~ 1 / R_m³.


The integral of the dy'² term


       r_s / r   *

       cos²(β)  *  (v_y² dt² + dy² + v_y² dy²

                       + 2 v_y dt dy + v_x v_y dx dy).

We have

       cos²(β) = 1 - sin²(β),

and 1 / r * sin²(β) ~ 1 / R_m³. We can assume that cos(β) = 1. Since v_y² has the same value for dm and dm'', the symmetry condition of the generalized shell theorem is satisfied and we can pretend that r = R_m.

We have

       v_y = ω R sin(α).

The integral of the v_y² term over the whole disk is
 
       2 G / R_m * ω²

            R₀  2 π
       *  ∫     ∫     R² sin²(α) ρ dR R dα
         0     0

       = 2 G / R_m * ω²  * ρ π R₀⁴ / 4

       = 1/2 G / R_m * ω² * 2 I

       = 4 G K / R_m.

1.      r_s / r * v_y² dt² :     4 G K / R_m * dt².

2.      r_s / r * dy² :             2 G M / R_m * dy².

3.      r_s / r * v_y² dy² :    4 G K / R_m * dy².

4.      r_s / r * 2 v_y dt dy :   the values for dm' and dm cancel each other out because the sign of v_y flips. We can ignore the term.

5.      r_s / r * v_x v_y dx dy : the sign of v_y flips for dm' and dm, and the term is canceled. We can ignore the term.


The integral of the dx' dy' term


        r_s / r 

          *  2 cos(β) sin(β) 

          *   (1 - v_x² + v_y²)

            * (  v_x  v_y dt² + v_x dt dy + v_y dt dx
   
               + v_x  v_y / 2 * dx²

               + (1 + v² / 2) * dx dy

               + v_x  v_y / 2 * dy²
               ).

We can assume that cos(β) = 1, because the difference from 1 only would bring a 1 / R_m⁴ term. Most terms above are ~ v³ or higher, and we can ignore them.

1.      r_s / r * 2 sin(β) v_x v_y dt²
    
       = 4 G dm / r * v² sin(β) cos(α) sin(α).

We have

       sin(β) = R / r * sin(α).

The integrand becomes

        4 G dm R / r² * v² cos(α) sin²(α).

The values for dm and dm'' cancel each other almost exactly:

       1 / (R_m - R)² - 1 / (R_m + R)²

       = 4 R_m R / (R_m² - R²)²

       ~ 1 / R_m³.

We can ignore the term.

2.      r_s / r * 2 sin(β) v_x dt dy

       = r_s / r * 2 v sin(β) cos(α) dt dy.

Again the values for dm and dm'' cancel each other almost exactly. We can ignore the term.

3.      r_s / r * 2 sin(β) v_y  dt dx

       = 4 G dm / r * R / r * sin²(α) v  dt dx

       = 4 G / R_m² * R sin²(α) ω R dm  dt dx.

The integral

       4 G / R_m² * ω

             R₀   2 π
       *   ∫      ∫     R² sin²(α) ρ dR R dα
          0       0

       =  G / R_m² * ω ρ π R₀⁴

       =  G / R_m² * 2 J.

The term is G / R_m² * 2 J  dt dx.

This term probably corresponds to the "dm' approaches faster" term of the August 10, 2023 blog post.

4.      r_s / r * sin(β) v_x v_y dx² :    we can ignore this like we did in case 1.

5.      r_s / r  * 2 sin(β) (1 + v² / 2) dx dy :    the values for dm' and dm cancel each other exactly. We can ignore the term.

6.      r_s / r * sin(β) v_x v_y dy² : we can ignore this like we did in case 1.

7.      r_s / r * 2 sin(β) * (- v_x² + v_y²) dx dy :    the values for dm' and dm cancel each other exactly. We can ignore the term.


Sanity checks


There probably are errors in the long calculations above.

Let us check if we can find the newtonian attraction term of the rotating mass. It is in the dt'² section above.

The dy² metric should be stretched like in the standard Schwarzschild metric. It is in the dy'² section above.

The dx² metric should be very little perturbed. There are no large dx² terms.

The "magnetic" component dt dx is zero. The magnetic term in dt'² is canceled by a term in dx' dy'. The canceling term probably is the "dm' approaches faster" term of August 10, 2023. It did cancel the magnetic component there, too.

On August 10, 2023 we saw that in the radial motion toward the test mass m, there is some positive frame dragging which comes from radial corrections. Let us try to spot that in this blog post.

It looks like that the frame dragging caused by the radial movement of dm' and dm toward / away from the observer m is "lost" in the metric. The reason is that the metric does not understand the handedness of the gravitomagnetic field?

Or is it so that we have to calculate the x derivatives of the metric explicitly?

There should probably be a non-zero x derivative of g_ty because if the observer m moves slightly to the positive x direction, then he is closer to the right side of the disk where mass is moving away from him.

The gravitomagnetic field caused by the x movement of dm, dm'' is preserved in the metric because even though they have opposite effects, the effect of dm'' is smaller because it farther away.


A metric cannot express the "handedness" rules for a gravitomagnetic field?

















On August 16, 2023 we claimed that the Einstein approximation formula loses "crucial information" about the movement of masses. We retracted the claim on August 20, 2023 when we realized that in the particular configuration, the metric g_ty (y is defined like in this blog post today), the value of g_ty varies with x.

But g_ty cannot vary with x for a rotating disk, since the metric has to have a rotational symmetry? No, since we are calculating with a fixed y axis which does not turn, there cannot be rotational symmetry. The metric for a neighboring observer should have the same y axis as the observer m, so that we can compare the values of g_ty.

If the neighbor uses his own y axis, then we must rotate his metric to use our own y axis. Let us calculate if we can get the right gravitomagnetic effect in that way.

Our present calculation method of the metric only calculates values for varying R_m. To get the values neighbors of the observer r, we have to use a rotation, or devise some other way.


A static metric around a rotating disk?


How can we model magnetism if there is no magnetic term?


This reminds us of the Ehrenfest paradox. One cannot define a fully consistent metric on a rotating disk itself?

Could it be that one cannot define a consistent metric around a rotating disk?

We have in earlier blog posts remarked that if a very heavy neutron star moves faster than what is the local speed of light inside it, then we cannot use static spatial coordinates at all. The neutron star must drag the coordinates along with it. There might be a similar thing happening around a rotating disk, even though the velocities are much less than the speed of light. Note that if we use rotating coordines, then the speed of the spatial coordinates does exceed the speed of light when we go to a great distance.

If one takes a fixed y coordinate line and adds the metric perturbations for each part of the rotating disk, then we end up in a metric which probably can reproduce the accelerations which we calculated on August 10, 2023. The left side of the disk is approaching the observer m and the right side receding from him. There is no rotational symmetry. The observer should see phenomena which are not symmetric.

But if one demands that the metric has to have rotational symmetry, then g_tx is the only component which can affect the x acceleration of a test mass m approaching along the y axis, and above we calculated that the dt dx, or g_tx is essentially zero.

It is as if we would be unable to define a metric which extends around the disk. This is like the Ehrenfest paradox.

We need to check carefully, if abandoning the rotational symmetry makes the metric sensible.


Conclusions


We will write a new blog post about this confusing situation. We may have made a calculation error on August 10, or in this blog post.

Since the left side of the disk is moving toward the observer m and the right receding,

       dg_ty / dx

should be non-zero. But if the metric is rotationally symmetric, the derivative must be zero.

The metric has to be given in rotating coordinates? But there should be a mapping to static coordinates.


The Gödel rotating metric contains closed time-like curves, which can be regarded as proving that the solution is not reasonable. There may be something similar going on around any rotating mass.

Monday, August 28, 2023

The gravity field of a (moving) cylinder

UPDATE February 23, 2024: The method below assumes that we can sum the (Schwarzschild) metric perturbations caused by each mass element dm at the test mass m. This is probably wrong. The summing is not linear at all. An example: a spherical shell of mass. The metric around it is the familiar Schwarzschild. But if we sum the metric perturbations for each element dm in the shell, we get a very different result, where the tangential metric is stretched a lot and the radial metric only a little.

In the case of the long cylinder, the stretching of the metric in the x direction (below) is probably very little and the stretching to the y direction is more than what we calculate below.

----

Let us calculate the approximate metric around a static, thin, lightweight uniform cylinder which is long, but not infinitely long. We start by approximating the metric around a static cylinder and then Lorentz transform the metric to obtain the metric for a moving cylinder.

We will also calculate the metrics using the Einstein approximation formula and compare to our results.


Summing Scwarzschild metrics using Schwarzschild coordinates



            cylinder length 2 L
                                               dm      ρ = dm / dx
  ======================•===
                                                   β₀ = angle to end
                                                           of cylinder
                                                   R = distance (m,
                                                               cylinder)


                           |     /           r = distance(m, dm)
                           |   / 
                           | / β = angle y axis vs. (m, dm)
                           •
                           m  test mass
    ^ y
    |
     ------> x


The test mass m is directly under the center of the cylinder. To obtain the metric at m, we sum the metric perturbation generated by each element dm of the cylinder. We use the Schwarzschild metric and coordinates for dm. The Schwarzschild radius:

       r_s = 2 G dm / c².

The mass dm has two effects at the test mass m:

1. it slows down clocks;

2. it squeezes rulers in the radial direction (m, dm).

Let us sum these effects for all parts dm of the cylinder, to obtain the metric at m.

The slowdown of time caused by dm is a perturbation to the Minkowski metric η:

             -1                   0                    0

              0                   1                    0

              0                   0                    1

The perturbation is

       dg_tt = r_s / r

                 = 2 G dm / c²  *  1 / (R / cos(β))

                 = 2 G / c² * 1 / R * cos(β) dm.

The stretching of the spatial metric by dm involves off-diagonal elements in the metric since a square stretched from an oblique angle is no longer a rectangle. However, these off-diagonal elements must cancel at the test mass m, because of the symmetry.

The perturbation to the diagonal elements of the metric is

       dg_xx = 2 G / c² * 1 / R * cos(β) sin(β) dm,

       dg_yy = 2 G / c² * 1 / R * cos(β) cos(β) dm.

We have to integrate over the whole cylinder. The mass element is

       dm = ρ R / cos²(β) * dβ.

We have:

       Δg_tt =
                     β₀
            2 *  ∫    2 G / c² * ρ / cos(β) * dβ
                  0
             
         =  4 G / c² * ρ 

             *  ln( 1 / cos(β₀)  +  tan(β₀) ).

If β₀ is almost π / 2 then the argument of ln is roughly 2 L / R. The derivative in that case is

       d(Δg_tt) / dy = 4 G / c² * ρ * 1 / R.

This corresponds to a gravity acceleration

       2 G ρ / R

toward the cylinder, which is the familiar newtonian formula.


The perturbations to the spatial metric:

       Δg_xx = 4 G / c² * ρ

                              β₀
                          * ∫   cos³(β) sin(β) dβ
                           0
     
                    = 4 G / c² * ρ

                           * -1/4 (cos⁴(β₀) - 1)

                    = G / c²  *  ρ (1  -  1/4 R⁴ / L⁴),

       Δg_yy  = 4 G / c² * ρ

                                 β₀
                             * ∫    cos⁴(β) dβ
                              0

                     = 4 G / c² * ρ 

                          * (3/8 β₀

                             + 1/4 sin(2 β₀) + 1/32 sin(4 β₀)

                    = 4 G / c² * ρ

                        * (3/16 π - 3/8 R / L
      
                           + 1/4 * 2 R / L + 1/32 * 4 R / L)

                    = G / c² * ρ * (3/4 π + 1/4 R / L).

We now have the metric perturbation near the center of a static cylinder:

     h = 

       G / c² * ρ *

         4 ln(2 L / R)  0                    0                     0

         0          1 - 1/4 R⁴ / L⁴         0                     0

         0                    0         3/4 π + 1/4 R / L      0

         0                    0                     0                     0


where we assumed that R / L is small.

What about the metric in the z direction in our diagram of the cylinder?

A short movement up from the screen to the z direction is tangential movement for all parts dm of the cylinder. The spatial metric is not stretched in the tangential direction. The perturbation in g_zz is zero. We add a fourth row and column of zeros to the matrix of h.


Orbits of the test mass m


If general relativity is Lorentz covariant, it does not matter if we calculate the orbit of a moving test mass against a static metric, or if we have a static test mass in the field of a moving cylinder.

If the test mass is initially static, then the above metric means that its y acceleration is the newtonian one.

What if the test mass m is moving to the positive x direction at a velocity v? The g_xx metric above is almost exactly constant close to the cylinder. The correcting term is only

       ~ R⁴.

Thus, the y acceleration for small R is essentially the newtonian one.


          ===================

                            ^  V
                            |
                            |
                            • --> v
                           m
      ^ y
      |
       ------> x


What about m moving to the positive x direction at a velocity v and toward the cylinder at a velocity V? We know from previous blog posts that the test mass gains more inertia as it approaches the cylinder. That is why the x velocity slows down. In the geodesic equation this is reflected by the proper time of m, τ slowing down. Let us check this using the geodesic equation and the metric which we calculated above. We will calculate

         d²x / dt².















We have μ = x. A relevant Christoffel symbol is

        Γ_tty =
              1/2 (dg_tt / dy + dg_ty / dt - dg_ty / dt)

                 = 1/2 dg_tt / dy.

It contributes to the x acceleration

        1/2 dg_tt / dy * V v.

The same contribution 

        1/2 dg_tt / dy * V v

comes from Γ_tyt. All the others are probably zero because of the symmetry and the diagonality of the metric.

Raising the time index in

       Γ^t_ty

flips the sign of the Christoffel symbol, since it is multiplied with 1 / g_tt = -1. We can assume that g_tt is almost exactly -1 if we make the mass of the cylinder small.

Suppose that the test mass moves closer to the cylinder Δy in a time Δt and gains the energy W. The test mass gains inertia worth 2 W / c². Let us calculate the x acceleration with our own methods.

The x velocity v slows down by a fraction

        2 W / c²   /    m.

The acceleration in the x direction is

        a_x = -2 W / (m c²)  *  v / Δt.

If dg_tt / dy = D, then

       W = 1/2 D Δy m c².

We get

        a_x = -D Δy m c² / (m c²)  *  v / Δt

               = -dg_tt / dy  * V v.

The acceleration agrees with the geodesic equation. Our own method is much easier to calculate in your head than the route through the geodesic equation.


Lorentz transformation to get the metric for a moving cylinder


We start from the static metric η + h and switch to a frame which moves at a velocity v to the negative direction of the x axis. The transformation Λ to the coordinates (marked with the prime ') in the moving frame is

       t'  =  γ (t + v x),

       x' =  γ (x + v t),

       y' =  y,

where

      γ = 1 / sqrt(1  -  v² / c²).

The inverse transformation is:

       t  =  γ (t' - v x'),

       x =  γ (x' - v t'),

       y =  y'.

We have

       R = sqrt(x² + y²),

if we put the origin at the center of the cylinder. Remember that our metric only hold close to the center of the cylinder.

The relevant part of the metric perturbation is:

h = 

    G / c² * ρ *

        4 ln(2 L / sqrt(x² + y²))    [ = T]               0

        0                     [X =]     1 - 1/4 (x² + y²)² / L⁴

or

       dτ² =

          -(1 + ρG/c² * 4 ln(2 L / sqrt(x² + y²)) dt²

        + (1 + ρG/c² * (1 - 1/4 (x² + y²)² / L⁴) dx².

Let us denote the tt element of h by T and the xx element by X. Using the formulae:

       dt  =  γ (dt' - v dx'),

       dx =  γ (dx' - v dt')
 
we obtain the transformation:

       T γ² (dt' - v dx')² + X γ² (dx' - v dt')²

       = γ² * ( T dt'² - 2 T v dt' dx' + T v² dx'²

                 + X dx'² - 2 X v dx' dt' + X v² dt'² ).

The relevant part of the perturbation matrix in the moving frame becomes:

   h_v =

       γ² *

           T + X v²                 -(T + X) v
        

          -(T + X) v                 X + T v²


Very close to the cylinder this is:

    h_v =

      γ² G / c² * ρ *

          4 ln(2 L / R) + v²      - (4 ln(2 L / R) + 1) v

       - (4 ln(2 L / R) + 1) v       1 + 4 ln(2 L / R) v²
            

      
           ==============   --> v

                         ^
                         |   V
                         | 
                         •  m


Let us use the geodesic equation to determine the x acceleration of a test mass m, which in the moving frame has v_y = V and v_x = 0. The relevant Christoffel symbols should be about g_xt and g_tx.



    









Since dx/dt = 0, the second term in the geodesic equation is zero. The relevant Christoffel symbols are:

       Γ_xty = 1/2 dg_xt / dy,

       Γ_xyt = 1/2 dg_xt / dy.

The derivative

       dg_xt / dy = -(dT + dX) v / dy

                         = -v dT / dy,

because X is essentially a constant. The x acceleration from the geodesic equation is:

       a_x = 1 / g_xx * v dT / dy * dy / dt * dt / dt

              = dg_tt / dy * V v


where we assumed that the mass of the cylinder is small, so that g_xx is essentially 1. Our result agrees with the calculation which we made in a frame where the cylinder is static. The flip of sign in a_y comes from the fact that in the last calculation x moves to the left relative to the cylinder.

General relativity should be Lorentz covariant. Then it does not matter if one lets the cylinder move or the test mass m move. The result should be the same. In our example this was true.

Close to the cylinder, and when v is small, the acceleration is
         
       a_x = V v 

        * d( γ² G / c² * ρ * (4 ln(2 L / R) + v²) ) / dy

            = 4 V v G / c² * ρ R * 1 / R²

            = V * 4 G / c² * v ρ / R.

This agrees with our calculation on August 14, 2023.


The Einstein approximation formula and the cylinder


We suspect that the Einstein approximation formula (1916) is incorrect. Let us use it to calculate the metric around the cylinder, and let us compare to our own, Schwarzschild-based results above.

















For a static cylinder, the "mapping" of the stress-energy tensor to the location of the test mass m gives:

                                            β₀
       γ'_tt  =  2 * 4 G / c²   ∫     ρ / cos(β) * dβ
                                         0

                =  8 G / c² * ρ 

                          *  ln( 1 / cos(β₀)  +  tan(β₀) ).

The metric perturbation is obtained by "trace reversing":

       h' =

          G / c² * ρ *

               4 ln(2 L / R)      0                          0

               0               4 ln(2 L / R)                 0

               0                        0        4 ln(2 L / R)

where we assumed that R / L is small. The tt component is the same as we calculated above from Schwarzschild metrics. But the xx and yy components are completely different.

For a moving cylinder (velocity v), the mass density per length ρ is multiplied by γ² because there is kinetic energy, and the cylinder is length contracted. The components tx and xt get a contribution from the movement of the mass:


   h_v' =

     γ² G / c² * ρ *

        4 ln(2 L / R)        4 ln(2 L / R)  v                0

        4 ln(2 L / R) v      4 ln(2 L / R)                   0

        0                                    0           4 ln(2 L / R)


This is very different from h_v.


Conclusions


Our own Schwarzschild-based method gives sensible and Lorentz covariant results for the cylinder. The Einstein approximation formula produces very different results, and the results are not Lorentz covariant. We believe that the Einstein approximation formula is erroneous. Gravitoelectromagnetism relies on the Einstein formula. This is the reason why results from gravitoelectromagnetism are not Lorentz covariant.

We will next proceed to calculate the approximate metric around a rotating ball, using the Schwarzschild-based method. We will check if orbits of the test mass agree with what we calculated on August 10, 2023.

The Einstein approximation: a strange acceleration term

The Einstein approximation formula is not simply about remapping the radial coordinate of the Schwarzschild metric (= bulging coordinates). In bulging coordinates, the metric has off-diagonal elements everywhere but on the x and y coordinate axes. This is demonstrated by the fact that the coordinate lines do not intersect orthogonally anywhere, but only on the coordinate axes.

















However, the Einstein approximate metric does not have off-diagonal elements anywhere for a static mass M.

Let us calculate what the Einstein approximate metric says about the acceleration if a test mass m moves tangentially relative to M.


Tangential movement in the Einstein approximate metric


In this post we have set c = 1.

In the Einstein approximation around a spherical mass M, the perturbation to the Minkowski metric η is:

     h =

        4 G *

              1/2 M / r             0                         0

               0                   1/2 M / r                  0

               0                        0            1/2 M / r


We will study a familiar configuration:


                             ●  M
                            /
                          /
                        /  r
                                      R = distance (m, M)


                              • --> v
                            m
   ^ y
   |
    -----> x

















We use the geodesic equation to determine the y acceleration of the test mass m. Let us multiply the geodesic equation (the first equation above) by dτ² / dt². Then we can assume that ds in the equation is dt.








Alternatively, we could start from the geodesic equation in the coordinate time (above). Since dy / dt is zero, the second term on the right is zero: we can use dt in the place of ds in the shorter form geodesic equation.

In the geodesic equation, μ = y, and αβ can be any of tt, tx, xt, xx. The combinations tx and xt do not contribute because the metric is diagonal.

For the location of the test mass m:

       dg_tt / dy    = 2 G M * 1 / R²,

       Γ^y_tt          = 1 / g_yy  * -G M / R².

Also,

       dg_xx / dy   =  2 G M * 1 / R²,

       Γ^y_xx          = 1 / g_yy * -G M / R².

We have

       Γ^y_tt * dt / dt * dt / dt      =  Γ^y_tt,

       Γ^y_xx * dx / dt * dx / dt   =  Γ^y_xx * v².

We can assume that g_yy = 1 by setting M small. The y acceleration is

       d²y / dt² = G M / R²

                       + v² G M / R².

The second acceleration term above is a strange extra term which does not appear if we use the Schwarzschild metric and coordinates.

We suspect that the extra term can lead to errors when we calculate gravitomagnetic effects.


Conclusions


The Einstein approximate metric is not obtained from the Schwarzschild metric by simply making the coordinates "bulging". Bulging would introduce off-diagonal elements, but the Einstein approximate metric is diagonal.

Anyway, the Einstein approximate metric introduces a strange radial acceleration term in a tangential movement.

We will next calculate what the Einstein approximation says about a moving cylinder. We know what the correct metric roughly is for a static cylinder. By Lorentz transforming we obtain the metric for a moving cylinder, and can compare to the Einstein approximate metric.

Saturday, August 26, 2023

The Einstein approximation Lorentz failure

UPDATE October 12, 2023: We forgot the γ M v² / r term T_xx from the stress-energy tensor T. We have to check if that changes anything.

----

UPDATE 2 August 27, 2023: The metric seems to understand that it moves with M, after all.

----

UPDATE August 27, 2023: The geodesic equation thinks that stretching of the spatial metric is "attached" to the coordinates, while it is attached to M. This may be a fundamental problem in general relativity.

We also corrected the error: the factor in the reverse Lorentz transformation is γ, not 1 / γ.

----

Let us analyze in detail why the Einstein approximation formula leads to a breach of Lorentz covariance when we switch to a moving frame. This is the simple configuration which we studied in our August 21, 2023 blog post. If the coordinates were used correctly, then Lorentz covariance would be preserved.















We showed in our August 15, 2023 post Summing metric perturbations is error-prone for wrong choices of coordinates that the Einstein-Thirring coordinates are obtained from Schwarzschild coordinates by defining a new radial coordinate R':

       R' = R  -  1/2 r_s,

where r_s is the Schwarzschild radius for the central mass M. If we draw coordinate lines for R' = 1, R' = 2, ... , the lines are r_s farther from M than the corresponding lines for the Schwarzschild coordinates. In this sense, the Einstein-Thirring coordinates are "bulging" relative to the Schwarzschild coordinates.  The pseudo-cartesian coordinates in the image above are bulging relative to standard cartesian coordinates.

Throughout this blog post we set the speed of light c = 1. We use the (- + +) sign convention.


The metric in the laboratory frame where the mass M and the test mass m are static


                               ●  M
                              /
                            /  r
                          /

                                         R = distance (m, M)


                               ^   d²y / dt²  acceleration
                               |
                               • m
   ^ y
   |
    -------> x


Let us start from a static mass M and a test mass m. The test mass is initially static but accelerates toward M. The coordinate distance (m, M) is R. We use (almost) cartesian coordinates. We want to calculate the y acceleration of m.








The Einstein approximation formula gives as the mapped stress-energy tensor T_μν / r at an arbitrary location at a distance r from M:

                M / r                  0                        0

                0                         0                        0

                0                         0                        0

Trace reversing gives us a metric perturbation of the Minkowski metric η:


     h₀ =

        4 G *

              1/2 M / r             0                         0

               0                   1/2 M / r                  0

               0                         0            1/2 M / r


The full metric is:

     η + h₀ =

              -1 + 2 G M / r     0                         0

               0              1 + 2 G M / r                0

               0                         0     1 + 2 G M / r















Let us calculate the y acceleration of the test mass m. Since m is initially static, only dt / dτ differs from zero in the geodesic equation:

       Γ_ytt  =  -1/2 dg_tt / dy

                  =  -1/2 d(4 G * 1/2  M / R ) / dy

                  =  -G M / R².

Then

       Γ^y_tt  =  -1 / g_yy  *  G M / R²,

where

       g_yy = 1  -  2 G M / R.

We get:

       d²y / dt²  *  dt² / dτ² 

       =  dt² / dτ²  *  1 / g_yy  *  G M / R²,

or

       d²y / dt²  =  1 / g_yy  *  G M / R².

If M is small, then we can assume that g_yy = 1. The acceleration is the familiar newtonian one.


Switch to moving coordinates: the Einstein approximation formula


                          M ● ----> v
                             /
                           /  r
                         /
       
                                         R = distance (m, M)


                               ^   d²y' / dt'²  acceleration
                               |
                           m • ----> v
   ^ y'
   |
    -------> x'


The new coordinates move to the left at a velocity v:

       |v| << 1.

We denote the new coordinates with the prime'. The Einstein approximation gives as the mapped stress-energy tensor T_μν / r at an arbitrary location:

               γ M / r         γ M v / r                  0

               γ M v / r            0                        0

               0                         0                        0

where

       γ = 1 / sqrt(1 - v²).

Trace reversing gives us the perturbation to the Minkowski metric:

     h₁ =

         2 G M / r *

               γ                       -2 γ v                     0

              -2 γ v                   γ                          0  

               0                         0                           γ

     =
     
        2 G M / r *

               1 + 1/2 v²       -2 v                         0

              -2 v               1 + 1/2 v²                  0

               0                       0              1 + 1/2 v²


where we used the fact that |v| is small and discarded ~ v³ terms.

The perturbation h₁ has to be summed to the Minkowski metric η to obtain the full metric:

     η =

              -1                        0                         0

               0                        1                         0

               0                        0                         1


Lorentz transformation Λ of the laboratory metric to moving coordinates


Let us compare the above result for h₁ to the Lorentz transformation of the metric η + h₀ in the laboratory frame into the moving frame. Let M move to the right so that its coordinates at a time t' are

       (x', y')  =  (v t',  R).

Then

       r  =  sqrt( (x' - v t')²  +  (y' - R)² ).

The coordinates of m are

       (x', y')  =  (v t',  0).

The transformation Λ to the coordinates (marked with the prime ') in the moving frame is

       t'   =  γ (t  +  v x),

       x'  =  γ (x  +  v t),

       y'  =  y.

The inverse transformation is:

       t   =  γ (t'  -  v x'),

       x  =  γ (x'  -  v t'),

       y  =  y'.

Let us first look at the transformation of the Minkowski metric η. If we write

      dt  =  γ  *  (dt'  -  v dx'),

      dx =  γ  *  (dx'  -  v dt'),

      dy =  dy',

we obtain

     Λ(η) = η =

            -1                          0                          0

             0                          1                          0

             0                          0                          1

as the metric in the moving frame. We have

       γ² * v²  =  (1 + v²)  *  v²  =  v²,

if we assume |v| << 1.

The transformation of the perturbation h₀ is

     Λ(h₀) = h₂ =

       2 G M / r * γ² *

              1 + v²                -2 v                       0

             -2 v                   1 + v²                     0

              0                          0                  1 * 1 / γ²


We write

       γ² = 1 + v².

We get:

     h₂ =

        2 G M / r *

              1 + 2 v²             -2 v                     0

             -2 v                  1 + 2 v²                 0

              0                         0                        1


where r = sqrt( (x' - v t')²  +  (y' - R)² ).

We dropped ~ v³ terms. We see that h₁ and h₂ differ in the g_t't' and g_x'x' elements.















We once again use the geodesic equation to determine the y' acceleration of the test mass m. Let us multiply the geodesic equation (the first equation above) by dτ² / dt'². Then we can assume that ds in the equation is dt'.

In the geodesic equation, μ = y', and αβ can be any of t't', t'x', x't', x'x'.

For the location of the test mass m:

       dg_t't' / dy'  =  2 G M (1 + 2 v²) * 1 / R²

       Γ^y'_t't'  =  1 / g_y'y'  *  -G M (1 + 2 v²) / R².

Also,

       dg_t'x' / dy'  =  -4 G M v / R²,

       Γ^y'_t'x'  =  1 / g_y'y' * 2 G M v / R²,

       Γ^y'_x't'  =  1 / g_y'y' * 2 G M v / R²,

       dg_x'x' / dy'  =  2 G M (1 + 2 v²) * 1 / R²,

       Γ^y'_x'x'  =  1 / g_y'y' * -G M (1 + 2 v²) / R².

Multiplying with v or v² where appropriate, and summing, we obtain the y' acceleration:

       d²y' / dt'²  =  1 / g_y'y' * (

                   G M (1 + 2 v²) / R²

                 - 2 G M v² / R²  

                 - 2 G M v² / R²

                 + G M v² / R²
                )
         
         =  1 / g_y'y' * (1 - v²) * G M / R²

         = (1 - v²) * G M / R²,

if M is small. We used the fact that g_y'y' goes to 1 when M goes to zero. We dropped terms ~ v⁴.

In the moving frame, the system (M, m) moves at the velocity v to the right. In the formula d²y / dt², the time dt is dilated by a factor 1 + 1/2 v² when looked at in the moving frame. Thus, the acceleration measured in the moving frame should be slower by a factor 1 - v². Our result matches the expectation!


What if we use the Schwarzschild coordinates? Does the geodesic equation work?


The above calculation uses the bulging Einstein-Thirring coordinates. If we instead use the Schwarzschild coordinates, then in the laboratory metric perturbation h₀, we have the metric x,x component zero.

Removing the contribution of the perturbation in the x,x metric causes the following changes to the formula of d²y' / dt'²:

1. G M (1 + 2 v²) / R² becomes G M (1 + v²) / R²,

2. both -2 G M v² / R² terms become -G M v² / R²,

3. + G M v² / R² drops off.


The y' acceleration stays the same, and is the correct one.


Analysis of the Einstein approximation error


The error was not in the bulging coordinates, after all, in this case. The test mass m moves directly below M, and the bulge in the y' coordinate lines does not generate a spurious y' acceleration to the movement of m.

The error simply is that the Einstein formula does not calculate the Lorentz transformation of the static metric which we had around M in the laboratory frame.

The Einstein formula does a Lorentz transformation of the stress-energy tensor, but the transformation for the (Schwarzschild) metric is more complicated, as we calculated above.


Conclusions


General relativity seems to be Lorentz covariant, though we have not yet proved that in detail.

The Einstein approximation formula does not calculate the Lorentz transformation of the metric correctly.

For other orbits of the test mass m, the bulging of the coordinate lines would produce spurious coordinate accelerations

       d²y' / dt'².

Also, the bulging coordinate lines may cause serious errors in the summing of perturbations for several masses M1, M2, etc. The bulging metric has "unnatural" stretching of tangential distances. For example, the metric for a long cylinder may be incorrect using the Einstein approximation. We will study this and other questions in subsequent blog posts.