Rubber ducky antenna: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Spinningspark
Reverted good faith edits by 15.211.169.107 (talk): Undo wikilink. The Narrowband article definition of the term is contrary to the usage here. Not saying they are rig...
 
en>Dsimic
→‎Origin of the name: Small cleanup
Line 1: Line 1:
Greetings! I am Marvella and I feel comfortable when individuals use the full title. California is exactly where I've always been living and I adore every day residing right here. My working day occupation is a librarian. Body building is one of the issues I love most.<br><br>Also visit my blog ... [http://raybana.com/chat/pg/profile/GIsaachse std testing at home]
The '''Komar mass''' (named after Arthur Komar<ref>A. Komar,'' Positive-Definite Energy Density and Global Consequences for General Relativity'', Physical Review, vol. 129, Issue 4, pp. 1873-1876 (1963)</ref>) of a system is one of several formal concepts of [[mass]] that are used in [[general relativity]].  The Komar mass can be defined in any [[stationary spacetime]], which is a [[spacetime]] in which all the [[metric tensor (general relativity)|metric]] can be written so that they are independent of time. Alternatively, a stationary spacetime can be defined as a spacetime which possesses a timelike [[Killing vector field]]. 
 
The following discussion is an expanded and simplified version of the motivational treatment in (Wald, 1984, pg 288).
 
== Motivation ==
 
Consider the [[Schwarzschild metric]].  Using the Schwarzschild basis, a [[Frame fields in general relativity#Example: Static observers in Schwarzschild vacuum|frame field]] for the Schwarzschild metric, one can find that the radial acceleration required to hold a test mass stationary at a Schwarzschild coordinate of r is:
 
<math>a^\hat{r} = \frac{m}{r^2 \sqrt{1-\frac{2m}{r c^2}}}</math>
 
Because the metric is static, there is a well-defined meaning to "holding a particle stationary".
 
Interpreting this acceleration as being due to a "gravitational force", we can then compute the integral of normal acceleration multiplied by area to get a "Gauss law" integral of:
 
:<math>\frac{4 \pi m}{\sqrt{1 - \frac{2m}{r c^2}}} </math>
 
While this approaches a constant as r approaches infinity, it is not a constant independent of r.  We are therefore motivated to introduce a correction factor to make the above integral independent of the radius r of the enclosing shell. For the Schwarzschild metric, this correction factor is just <math>\sqrt{g_{tt}}</math>, the "red-shift" or "time dilation" factor at distance r.  One may also view this factor as "correcting" the local force to the "force at infinity", the force that an observer at infinity would need to apply through a string to hold the particle stationary.  (Wald, 1984).
 
To proceed further, we will write down a line element for a static metric.
 
:<math> ds^2 = -g_{tt} \, dt^2 + \mathrm{quadratic\ form}(dx \, dy \, dz) \,</math>
 
where g<sub>tt</sub> and the quadratic form are functions only of the spatial coordinates x,y,z and are not functions of time.  In spite of our choices of variable names, it should not be assumed that our coordinate system is Cartesian.  The fact that none of the metric coefficients are functions of time make the metric stationary: the additional fact that there are no "cross terms" involving both time and space components (such as dx dt) make it static.
 
Because of the simplifying assumption that some of the metric coefficients are zero, some of our results in this motivational treatment will not be as general as they could be.
 
In flat space-time, the proper acceleration required to hold station is du/d tau, where u is the 4-velocity of our hovering particle and tau is the proper time.  In curved space-time, we must take the covariant derivative.  Thus we compute the acceleration vector as:
 
:<math> a^b = \nabla_u u^b = u^c \nabla_c u^b</math>
:<math>a_b = u^c \nabla_c u_b </math>
 
where u<sup>b</sup> is a unit time-like vector such that u<sup>b</sup> u<sub>b</sub> = -1.
 
The component of the acceleration vector normal to the surface is
:<math>a_{\mathrm{norm}}= N^b a_b \,</math> where N<sup>b</sup> is a unit vector normal to the surface.
 
In a Schwarzschild coordinate system, for example, we find that
:<math> N^b a_b = \left( \frac {\partial g_{tt}} {\partial r} c^2 \right) / \left( 2 g_{tt} \sqrt{g_{rr}} \right) = \frac{m}{r^2 \sqrt{1-\frac{2m}{r c^2}}}</math>
as expected - we have simply re-derived the previous results presented in a frame-field in a coordinate basis.
 
We define <math>a\mathrm{inf} = \sqrt{g_{tt}} a\,</math> so that in our Schwarzschild example <math>N^b a\mathrm{inf}_b = m/r^2\,</math>.
 
We can, if we desire, derive the accelerations a<sub>b</sub> and the adjusted "acceleration at infinity" ainf<sub>b</sub> from a scalar potential Z, though there is not necessarily any particular advantage in doing so. (Wald 1984, pg 158, problem 4)
 
<math>a_b = \nabla_b Z_1 \qquad Z_1 = \ln{gtt}</math>
<math>a\mathrm{inf}_b = \nabla_b Z_2 \qquad Z_2 = \sqrt{g_{tt}}</math>
 
We will demonstrate that integrating the normal component of the "acceleration at infinity" ainf over a bounding surface will give us a quantity that does not depend on the shape of the enclosing sphere, so that we can calculate the mass enclosed by a sphere by the integral
 
<math>m = -\frac{1}{4 \pi} \int_A N^b a\mathrm{inf}_b dA</math>
 
To make this demonstration, we need to express this surface integral as a volume integral.  In flat space-time, we would use [[Stokes theorem]] and integrate <math>-\nabla \cdot a\mathrm{inf}</math> over the volume.  In curved space-time, this approach needs to be modified slightly.
 
Using the formulas for [[Maxwell's equations in curved spacetime|electromagnetism in curved space-time]] as a guide, we write instead.
 
<!--sign - wrong here, or wrong later? -->
<math>F_{ab} = a\,\mathrm{inf}_a \,u_b - a\,\mathrm{inf}_b \,u_a \,</math>
 
where F plays a role similar to the "Faraday tensor", in that <math>a\mathrm{inf}_a = F_{ab} u^b\,</math> We can then find the value of "gravitational charge", i.e. mass, by evaluating
 
<math>\nabla^a F_{ab} u^b</math> and integrating it over the volume of our sphere.
 
An alternate approach would be to use [[Differential form#Integration of form|differential forms]], but the approach above is computationally more convenient as well as not requiring the reader to understand differential forms.
 
A  lengthly, but straightforward (with computer algebra) calculation from our assumed line element shows us that
 
<math> -u^b \nabla^a F_{ab} =  \sqrt{g_{tt}} R_{00} u^a u^b = \sqrt{g_{tt}} R_{ab} u^a u^b </math>
 
Thus we can write
 
<math>m = \frac {\sqrt{g_{tt}}} {4 \pi} \int_V R_{ab} u^a u^b </math>
 
In any vacuum region of space-time, all components of the Ricci tensor must be zero.  This demonstrates that enclosing any amount of vacuum will not change our volume integral.  It also means that our volume integral will be constant for any enclosing surface, as long as we enclose all of the gravitating mass inside our surface.   Because Stokes theorem guarantees that our surface integral is equal to the above volume integral, our surface integral will also be independent of the enclosing surface as long as the surface encloses all of the gravitating mass.
 
By using Einstein's Field Equations
 
:<math>G^u{}_v = R^u{}_v - \frac{1}{2} R I^u{}_v = 8 \pi T^u{}_v</math>
 
letting u=v and summing, we can show that R = -8 π T.
 
This allows us to rewrite our mass formula as a volume integral of the stress-energy tensor.
 
<math> m =  \int_V \sqrt{g_{tt}} \left( 2 T_{ab} - T g_{ab} \right) u^a u^b dV</math>
:where V is the volume being integrated over
:T<sub>ab</sub> is the [[Stress-energy tensor]]
:u<sup>a</sub> is a unit time-like vector such that u<sup>a</sup> u<sub>a</sub> = -1
 
== Komar mass as volume integral - general stationary metric ==
 
To make the formula for Komar mass work for a general stationary metric, regardless of the choice of coordinates, it must be modified slightly.  We will present the applicable result from (Wald, 1984 eq 11.2.10 ) without a formal proof.
 
<math> m =  \int_V \left( 2 T_{ab} - T g_{ab} \right) u^a \xi^b dV</math>
 
:where V is the volume being integrated over
:T<sub>ab</sub> is the [[Stress-energy tensor]]
:u<sup>a</sub> is a unit time-like vector such that u<sup>a</sup> u<sub>a</sub> = -1
:<math>\xi^b</math> is a [[Killing vector]], which expresses the time-translation symmetry of any [[stationary metric]].  The Killing vector is normalized so that it has a unit length at infinity, i.e. so that <math>\xi^a \xi_a = -1</math> at infinity.
 
Note that <math>\xi^b\,</math> replaces <math> \sqrt{g_{tt}} u^b\,</math> in our motivational result.
 
If none of the metric coefficients <math>g_{ab}\,</math> are functions of time, <math>\xi^a = \left( 1, 0, 0, 0 \right) </math>
 
While it is not ''necessary'' to choose coordinates for a stationary space-time such that the metric coefficients are independent of time, it is often ''convenient''. 
 
When we chose such coordinates, the time-like Killing vector for our system <math>\xi^a\,</math>  becomes a scalar multiple of a unit coordinate-time vector <math>u^a\,</math>, i.e.  <math>\xi^a = K u^a\,</math>.  When this is the case, we can rewrite our formula as
 
<math> m = \int_V \left( 2T_{00} - T g_{00} \right) K dV</math>
 
Because <math>u^a</math> is by definition a unit vector, K is just the length of <math>\xi^b</math>, i.e. K = <math>\sqrt{- \xi^a \xi_a}</math>.
 
Evaluating the "red-shift" factor K based on our knowledge of the components of <math>\xi^a</math>, we can see that  K = <math>\sqrt{g_{tt}}</math>.
 
If we chose our spatial coordinates so that we have a locally [[Minkowskian]] metric <math>g_{ab} = \eta_{ab}\,</math> we know that
 
:<math>g_{00}=-1, T = -T_{00}+ T_{11}+T_{22}+T_{33}\,</math>
 
With these coordinate choices, we can write our Komar integral as
 
:<math>m = \int_V \sqrt{-\xi^a \xi_a} \left( T_{00}+T_{11}+T_{22}+T_{33} \right) dV </math>
 
While we can't choose a coordinate system to make a curved space-time globally Minkowskian, the above formula provides some insight into the meaning of the Komar mass formula.  Essentially, both energy and pressure contribute to the Komar mass.  Furthermore, the contribution of local energy and mass to the system mass is multiplied by the local "red shift" factor <math>K = \sqrt{g_{tt}} = \sqrt{-\xi^a \xi_a}</math>
 
== Komar mass as surface integral - general stationary metric ==
 
We also wish to give the general result for expressing the Komar mass as a surface integral.
 
The formula for the Komar mass in terms of the metric and its Killing vector is (Wald, 1984, pg 289, formula 11.2.9)
 
<math>m = - \frac{1}{8 \pi} \int_S \epsilon_{abcd} \nabla^c \xi^d </math>
:where <math>\epsilon_{abcd}\,</math> are the [[Levi-Civita symbol|Levi-civita]] symbols
:<math>\xi^d</math> is the [[Killing vector]] of our [[stationary metric]], normalized so that <math>\xi^a \xi_a = -1</math> at infinity.
 
The surface integral above is interpreted as the [[Differential form#Integration of forms|"natural"]] integral of a two form over a manifold.
 
As mentioned previously, if none of the metric coefficients <math>g_{ab}\,</math> are functions of time, <math>\xi^a = \left( 1, 0, 0, 0 \right) </math>
 
==  See also ==
* [[Mass in general relativity]]
 
== Notes ==
{{reflist}}
 
== References ==
 
*{{cite book | title = General Relativity | last = Wald | first = Robert M | publisher = University of Chicago Press | year = 1984 | isbn=0-226-87033-2}}
 
*{{cite book | title=Gravitation | last = Misner, Thorne, Wheeler | year=1973 | publisher = W H Freeman and Company| isbn=0-7167-0344-0}}
 
[[Category:General relativity]]
[[Category:Mass]]

Revision as of 02:30, 28 October 2013

The Komar mass (named after Arthur Komar[1]) of a system is one of several formal concepts of mass that are used in general relativity. The Komar mass can be defined in any stationary spacetime, which is a spacetime in which all the metric can be written so that they are independent of time. Alternatively, a stationary spacetime can be defined as a spacetime which possesses a timelike Killing vector field.

The following discussion is an expanded and simplified version of the motivational treatment in (Wald, 1984, pg 288).

Motivation

Consider the Schwarzschild metric. Using the Schwarzschild basis, a frame field for the Schwarzschild metric, one can find that the radial acceleration required to hold a test mass stationary at a Schwarzschild coordinate of r is:

Because the metric is static, there is a well-defined meaning to "holding a particle stationary".

Interpreting this acceleration as being due to a "gravitational force", we can then compute the integral of normal acceleration multiplied by area to get a "Gauss law" integral of:

While this approaches a constant as r approaches infinity, it is not a constant independent of r. We are therefore motivated to introduce a correction factor to make the above integral independent of the radius r of the enclosing shell. For the Schwarzschild metric, this correction factor is just , the "red-shift" or "time dilation" factor at distance r. One may also view this factor as "correcting" the local force to the "force at infinity", the force that an observer at infinity would need to apply through a string to hold the particle stationary. (Wald, 1984).

To proceed further, we will write down a line element for a static metric.

where gtt and the quadratic form are functions only of the spatial coordinates x,y,z and are not functions of time. In spite of our choices of variable names, it should not be assumed that our coordinate system is Cartesian. The fact that none of the metric coefficients are functions of time make the metric stationary: the additional fact that there are no "cross terms" involving both time and space components (such as dx dt) make it static.

Because of the simplifying assumption that some of the metric coefficients are zero, some of our results in this motivational treatment will not be as general as they could be.

In flat space-time, the proper acceleration required to hold station is du/d tau, where u is the 4-velocity of our hovering particle and tau is the proper time. In curved space-time, we must take the covariant derivative. Thus we compute the acceleration vector as:

where ub is a unit time-like vector such that ub ub = -1.

The component of the acceleration vector normal to the surface is

where Nb is a unit vector normal to the surface.

In a Schwarzschild coordinate system, for example, we find that

as expected - we have simply re-derived the previous results presented in a frame-field in a coordinate basis.

We define so that in our Schwarzschild example .

We can, if we desire, derive the accelerations ab and the adjusted "acceleration at infinity" ainfb from a scalar potential Z, though there is not necessarily any particular advantage in doing so. (Wald 1984, pg 158, problem 4)

We will demonstrate that integrating the normal component of the "acceleration at infinity" ainf over a bounding surface will give us a quantity that does not depend on the shape of the enclosing sphere, so that we can calculate the mass enclosed by a sphere by the integral

To make this demonstration, we need to express this surface integral as a volume integral. In flat space-time, we would use Stokes theorem and integrate over the volume. In curved space-time, this approach needs to be modified slightly.

Using the formulas for electromagnetism in curved space-time as a guide, we write instead.

where F plays a role similar to the "Faraday tensor", in that We can then find the value of "gravitational charge", i.e. mass, by evaluating

and integrating it over the volume of our sphere.

An alternate approach would be to use differential forms, but the approach above is computationally more convenient as well as not requiring the reader to understand differential forms.

A lengthly, but straightforward (with computer algebra) calculation from our assumed line element shows us that

Thus we can write

In any vacuum region of space-time, all components of the Ricci tensor must be zero. This demonstrates that enclosing any amount of vacuum will not change our volume integral. It also means that our volume integral will be constant for any enclosing surface, as long as we enclose all of the gravitating mass inside our surface. Because Stokes theorem guarantees that our surface integral is equal to the above volume integral, our surface integral will also be independent of the enclosing surface as long as the surface encloses all of the gravitating mass.

By using Einstein's Field Equations

letting u=v and summing, we can show that R = -8 π T.

This allows us to rewrite our mass formula as a volume integral of the stress-energy tensor.

where V is the volume being integrated over
Tab is the Stress-energy tensor
ua is a unit time-like vector such that ua ua = -1

Komar mass as volume integral - general stationary metric

To make the formula for Komar mass work for a general stationary metric, regardless of the choice of coordinates, it must be modified slightly. We will present the applicable result from (Wald, 1984 eq 11.2.10 ) without a formal proof.

where V is the volume being integrated over
Tab is the Stress-energy tensor
ua is a unit time-like vector such that ua ua = -1
is a Killing vector, which expresses the time-translation symmetry of any stationary metric. The Killing vector is normalized so that it has a unit length at infinity, i.e. so that at infinity.

Note that replaces in our motivational result.

If none of the metric coefficients are functions of time,

While it is not necessary to choose coordinates for a stationary space-time such that the metric coefficients are independent of time, it is often convenient.

When we chose such coordinates, the time-like Killing vector for our system becomes a scalar multiple of a unit coordinate-time vector , i.e. . When this is the case, we can rewrite our formula as

Because is by definition a unit vector, K is just the length of , i.e. K = .

Evaluating the "red-shift" factor K based on our knowledge of the components of , we can see that K = .

If we chose our spatial coordinates so that we have a locally Minkowskian metric we know that

With these coordinate choices, we can write our Komar integral as

While we can't choose a coordinate system to make a curved space-time globally Minkowskian, the above formula provides some insight into the meaning of the Komar mass formula. Essentially, both energy and pressure contribute to the Komar mass. Furthermore, the contribution of local energy and mass to the system mass is multiplied by the local "red shift" factor

Komar mass as surface integral - general stationary metric

We also wish to give the general result for expressing the Komar mass as a surface integral.

The formula for the Komar mass in terms of the metric and its Killing vector is (Wald, 1984, pg 289, formula 11.2.9)

where are the Levi-civita symbols
is the Killing vector of our stationary metric, normalized so that at infinity.

The surface integral above is interpreted as the "natural" integral of a two form over a manifold.

As mentioned previously, if none of the metric coefficients are functions of time,

See also

Notes

43 year old Petroleum Engineer Harry from Deep River, usually spends time with hobbies and interests like renting movies, property developers in singapore new condominium and vehicle racing. Constantly enjoys going to destinations like Camino Real de Tierra Adentro.

References

  • 20 year-old Real Estate Agent Rusty from Saint-Paul, has hobbies and interests which includes monopoly, property developers in singapore and poker. Will soon undertake a contiki trip that may include going to the Lower Valley of the Omo.

    My blog: http://www.primaboinca.com/view_profile.php?userid=5889534
  • 20 year-old Real Estate Agent Rusty from Saint-Paul, has hobbies and interests which includes monopoly, property developers in singapore and poker. Will soon undertake a contiki trip that may include going to the Lower Valley of the Omo.

    My blog: http://www.primaboinca.com/view_profile.php?userid=5889534
  1. A. Komar, Positive-Definite Energy Density and Global Consequences for General Relativity, Physical Review, vol. 129, Issue 4, pp. 1873-1876 (1963)