Lorentz covariance

In physics, Lorentz symmetry, named for Hendrik Lorentz, is "the feature of nature that says experimental results are independent of the orientation or the boost velocity of the laboratory through space".[1] In everyday language, it means that the laws of physics stay the same for all observers that are moving with respect to one another with a uniform velocity. Lorentz covariance, a related concept, is a key property of spacetime following from the special theory of relativity. Lorentz covariance has two distinct, but closely related meanings:

  1. A physical quantity is said to be Lorentz covariant if it transforms under a given representation of the Lorentz group. According to the representation theory of the Lorentz group, these quantities are built out of scalars, four-vectors, four-tensors, and spinors. In particular, a Lorentz covariant scalar (e.g., the space-time interval) remains the same under Lorentz transformations and is said to be a Lorentz invariant (i.e., they transform under the trivial representation).
  2. An equation is said to be Lorentz covariant if it can be written in terms of Lorentz covariant quantities (confusingly, some use the term invariant here). The key property of such equations is that if they hold in one inertial frame, then they hold in any inertial frame; this follows from the result that if all the components of a tensor vanish in one frame, they vanish in every frame. This condition is a requirement according to the principle of relativity; i.e., all non-gravitational laws must make the same predictions for identical experiments taking place at the same spacetime event in two different inertial frames of reference.

This usage of the term covariant should not be confused with the related concept of a covariant vector. On manifolds, the words covariant and contravariant refer to how objects transform under general coordinate transformations. Confusingly, both covariant and contravariant four-vectors can be Lorentz covariant quantities.

Local Lorentz covariance, which follows from general relativity, refers to Lorentz covariance applying only locally in an infinitesimal region of spacetime at every point. There is a generalization of this concept to cover Poincaré covariance and Poincaré invariance.

Examples

In general, the nature of a Lorentz tensor can be identified by its tensor order, which is the number of free indices it has. No indices implies it is a scalar, one implies that it is a vector, etc. Furthermore, any number of new scalars, vectors ,etc. can be made by contracting or creating an outer product of any kinds of tensors together, but many of these may not have any real physical meaning. Some of those tensors that do have a physical interpretation are listed (by no means exhaustively) below.

Please note, the metric sign convention such that η = diag (1, −1, −1, −1) is used throughout the article.

Scalars

Spacetime interval
Proper time (for timelike intervals)
Proper distance (for spacelike intervals)
Rest mass
Electromagnetism invariants
D'Alembertian/wave operator

Four-vectors

4-displacement
4-position
4-gradient
which is the 4D partial derivative:
4-velocity
where
4-momentum
where
4-current
where

Four-tensors

Kronecker delta
Minkowski metric (the metric of flat space according to general relativity)
Levi-Civita symbol
Electromagnetic field tensor (using a metric signature of + − − − )
Dual electromagnetic field tensor

Lorentz violating models

In standard field theory, there are very strict and severe constraints on marginal and relevant Lorentz violating operators within both QED and the Standard Model. Irrelevant Lorentz violating operators may be suppressed by a high cutoff scale, but they typically induce marginal and relevant Lorentz violating operators via radiative corrections. So, we also have very strict and severe constraints on irrelevant Lorentz violating operators.

Since some approaches to quantum gravity lead to violations of Lorentz invariance,[2] these studies are part of Phenomenological Quantum Gravity.

Lorentz violating models typically fall into four classes:

Models belonging to the first two classes can be consistent with experiment if Lorentz breaking happens at Planck scale or beyond it, or even before it in suitable preonic models,[5] and if Lorentz symmetry violation is governed by a suitable energy-dependent parameter. One then has a class of models which deviate from Poincaré symmetry near the Planck scale but still flows towards an exact Poincaré group at very large length scales. This is also true for the third class, which is furthermore protected from radiative corrections as one still has an exact (quantum) symmetry.

Even though there is no evidence of the violation of Lorentz invariance, several experimental searches for such violations have been performed during recent years. A detailed summary of the results of these searches is given in the Data Tables for Lorentz and CPT Violation.[6]

See also

References

  1. "Framing Lorentz symmetry". CERN Courier. 2004-11-24. Retrieved 2013-05-26.
  2. Mattingly, David (2005). "Modern Tests of Lorentz Invariance". Living Reviews in Relativity. 8. arXiv:gr-qc/0502097Freely accessible. Bibcode:2005LRR.....8....5M. doi:10.12942/lrr-2005-5.
  3. Luis Gonzalez-Mestres (1995-05-25). "Properties of a possible class of particles able to travel faster than light".
  4. Luis Gonzalez-Mestres (1997-05-26). "Absence of Greisen-Zatsepin-Kuzmin Cutoff and Stability of Unstable Particles at Very High Energy, as a Consequence of Lorentz Symmetry Violation".
  5. Luis Gonzalez-Mestres (2014). "Ultra-high energy physics and standard basic principles. Do Planck units really make sense?" (PDF). EPJ Web of Conferences (ICNFP 2013 Conference). doi:10.1051/epjconf/20147100062.
  6. Kostelecky, V.A.; Russell, N. (2010). "Data Tables for Lorentz and CPT Violation". arXiv:0801.0287v3Freely accessible.
This article is issued from Wikipedia - version of the 9/22/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.