Skip to main content
Log in

Probing the Vacuum of Particle Physics with Precise Laser Interferometry

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

The discovery of the Higgs boson at LHC confirms that what we experience as empty space should actually be thought as a condensate of elementary quanta. This condensate characterizes the physically realized form of relativity and could play the role of preferred reference frame in a modern Lorentzian approach. This observation suggests a new interpretative scheme to understand the unexplained residuals in the old ether-drift experiments where light was still propagating in gaseous systems. Differently from present vacuum experiments, where anyhow deviations from Special Relativity are expected to be at the limit of visibility, these now acquire a crucial importance and become consistent with the Earth’s velocity of 370 km/s which characterizes the CMB anisotropy. In the same scheme, one can also understand the difference with the other experiments where light propagates in strongly bound systems such as solid or liquid transparent media. This non-trivial level of consistency motivates a new generation of precise laser interferometry experiments which explore the same particle physics vacuum and, in this sense, are complementary to those with high-energy accelerators.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. We ignore here the problem of vacuum degeneracy by assuming that any overlapping among equivalent vacua vanishes in the infinite-volume limit of quantum field theory (see e.g. S. Weinberg, The Quantum Theory of Fields, Cambridge University press, Vol. II, pp. 163–167).

  2. One could also argue that a satisfactory solution of the vacuum energy problem lies definitely beyond flat space. Nevertheless, in the absence of a consistent quantum theory of gravity, physical models of the vacuum in flat space can be useful to clarify a crucial point that, so far, remains obscure: the huge difference which is seen when comparing the typical vacuum-energy scales of particle physics with the value of the cosmological term needed in Einstein’s equations to fit the observations. In fact, the picture of the vacuum as a superfluid can naturally explain why there might be no non-trivial macroscopic curvature in the equilibrium state where any liquid is self-sustaining [3]. In any liquid, in fact, curvature requires deviations from the equilibrium state. In such representation of the lowest energy state, where the arbitrarily large condensation energy of the liquid plays no observable role, one can intuitively understand why curvature effects can be orders of magnitude smaller than those naively expected. In this perspective, ‘emergent-gravity’ approaches [1113], where gravity somehow arises from long-wavelength excitations of the same physical flat-space vacuum, may become natural and, to find the appropriate infinitesimal value of the cosmological term [14, 15], we are lead to sharpen our understanding of the vacuum structure and of its excitation mechanisms by starting from the picture of a superfluid medium in flat space.

  3. This picture reflects the basic Kolmogorov theory [34] of a fluid with vanishingly small viscosity.

  4. The dots for Michelson–Pease–Pearson (MPP) indicate that no error estimate is possible in this case. In fact, only one basic experimental session is explicitly reported in the literature. Since Miller’s extensive observations (see Fig. 22 of Ref. [25]), within their errors, gave fluctuations of the observable velocity in the wide range 4–14 km/s, a single observation giving \(v_\mathrm{obs} \sim \) 4 km/s cannot be interpreted as a refutation. This becomes even more true by noticing that the single MPP session explicitly reported was chosen, within a period of several months, to represent an example of extremely small ether-drift effect, see [8] for more details.

  5. One should appreciate how with such an explanation one can also reconcile Shankland’s criticism [45] with Miller’s claim for a non-zero ether-drift. In fact, by extracting the average velocity from the amplitude data there is a remarkable consistency among the various determinations of Table 1. Therefore, Shankland’s interpretation in terms of a temperature gradient is only acceptable provided this gradient represents a non-local effect as in our model.

  6. For instance, an important element to increase the overall stability and minimize systematic effects may consist in obtaining the two optical resonators from the same block of material as with the crossed optical cavity of Ref. [50].

  7. There is a subtle difference between our Eqs.(67) and (68) and the corresponding Eqs. (6) and (10) of Ref. [21] that has to do with the relativistic aberration of the angles. Namely, in Ref. [21], with the (wrong) motivation that the anisotropy is \(\mathcal{O}(\beta ^2)\), no attention was paid to the precise definition of the angle between the Earth’s velocity and the direction of the photon momentum. Thus the two-way speed of light in the \(S'\) frame was parameterized in terms of the angle \(\theta \equiv \theta _\Sigma \) as seen in the \(\Sigma \) frame. This can be explicitly checked by replacing in our Eqs. (67) and(68) the aberration relation \(\cos \theta _\mathrm{lab}=(-\beta + \cos \theta _\Sigma )/ (1-\beta \cos \theta _\Sigma )\) or equivalently by replacing \(\cos \theta _{\Sigma }=(\beta + \cos \theta _\mathrm{lab})/ (1+\beta \cos \theta _\mathrm{lab})\) in Eqs. (6) and (10) of Ref. [21]. However, the apparatus is at rest in the laboratory frame, so that the correct orthogonality condition of two optical cavities at angles \(\theta \) and \(\pi /2 + \theta \) is expressed in terms of \(\theta =\theta _\mathrm{lab}\) and not in terms of \(\theta =\theta _{\Sigma }\). This trivial remark produces however a non-trivial difference. In fact, the final anisotropy is now smaller by a factor of 3 than the one computed in Ref. [21] by adopting the wrong definition of orthogonality in terms of \(\theta =\theta _{\Sigma }\).

References

  1. ’t Hooft, G.: Search of the Ultimate Building Blocks. Cambridge University Press, Cambridge (1997)

    Google Scholar 

  2. Consoli, M., Stevenson, P.M.: Physical mechanisms generating spontaneous symmetry breaking and a hierarchy of scales. Int. J. Mod. Phys. A15, 133 (2000)

    ADS  Google Scholar 

  3. Volovik, G.E.: Superfluid analogies of cosmological phenomena. Phys. Rep. 351, 195 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Consoli, M., Pagano, A., Pappalardo, L.: Vacuum condensates and ether-drift experiments. Phys. Lett. A318, 292 (2003)

    Article  ADS  Google Scholar 

  5. Streater, R.F., Wightman, A.S.: PCT, Spin and Statistics, and All That. W. A. Benjamin, New York (1964)

    MATH  Google Scholar 

  6. Consoli, M., Costanzo, E.: Is the physical vacuum a preferred frame? Eur. Phys. J. C54, 285 (2008)

    Article  ADS  Google Scholar 

  7. Consoli, M., Costanzo, E.: Precision tests with a new class of dedicated ether-drift experiments. Eur. Phys. J. C55, 469 (2008)

    Article  ADS  Google Scholar 

  8. Consoli, M., Matheson, C., Pluchino, A.: The classical ether-drift experiments: a modern re-interpretation. Eur. Phys. J. Plus 128, 71 (2013)

    Article  Google Scholar 

  9. Zeldovich, Y.B.: The cosmological constant and the theory of elementary particles. Sov. Phys. Usp. 11, 381 (1968)

    Article  ADS  Google Scholar 

  10. Weinberg, S.: The cosmological constant problem. Rev. Mod. Phys. 61, 1 (1989)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Barcelo, C., Liberati, S., Visser, M.: Analog gravity from field theory normal modes? Class. Quantum Grav. 18, 3595 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Visser, M., Barcelo, C., Liberati, S.: Analogue models of and for gravity. Gen. Relat. Gravit. 34, 1719 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Consoli, M.: Ultraweak excitations of the quantum vacuum as physical models of gravity. Class. Quantum Gravity 26, 225008 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  14. Jannes, G., Volovik, G.E.: The cosmological constant: a lesson from the effective gravity of topological Weyl media. JETP Lett. 96, 215 (2012)

    Article  ADS  Google Scholar 

  15. Finazzi, S., Liberati, S., Sindoni, L.: Cosmological constant: a lesson from Bose–Einstein condensates. Phys. Rev. Lett. 108, 071101 (2012)

    Article  ADS  Google Scholar 

  16. Leggett, A.J.: Quantum Liquids, p. 102. Oxford University Press, New York (2006)

    Book  Google Scholar 

  17. Müller, H., et al.: Precision test of the isotropy of light propagation. Appl. Phys. B 77, 719 (2003)

    Article  ADS  Google Scholar 

  18. Hughes, V.W., Robinson, H.G., Beltran-Lopez, V.: Upper limit for the anisotropy of inertial mass from nuclear resonance experiments. Phys. Rev. Lett. 4, 342 (1960)

    Article  ADS  Google Scholar 

  19. Drever, R.W.P.: A search for anisotropy of inertial mass using a free precession technique. Philos. Mag. 6, 683 (1961)

    Article  ADS  Google Scholar 

  20. Will, C.M.: The confrontation between general relativity and experiment. arXiv:gr-qc/0510072

  21. Consoli, M., Costanzo, E.: From classical to modern ether-drift experiments: the narrow window for a preferred frame. Phys. Lett. A333, 355 (2004)

    Article  ADS  Google Scholar 

  22. Consoli, M., Costanzo, E.: Old and new ether-drift experiments: a sharp test for a preferred frame. N. Cim. 119B, 393 (2004)

    ADS  Google Scholar 

  23. Shamir, J., Fox, R.: A new experimental test of special relativity. N. Cim. 62B, 258 (1969)

    Article  ADS  Google Scholar 

  24. Hicks, W.M.: On the Michelson Morley experiment relating to the drift of ether. Philos. Mag. 3, 9 (1902)

    Article  MATH  Google Scholar 

  25. Miller, D.C.: The ether-drift experiment and the determination of the absolute motion of the earth. Rev. Mod. Phys. 5, 203 (1933)

    Article  ADS  MATH  Google Scholar 

  26. Nassau, J.J., Morse, P.M.: A study of solar motion by harmonic analysis. Astrophys. J. 65, 73 (1927)

    Article  ADS  Google Scholar 

  27. Herrmann, S., et al.: Test of the isotropy of the speed of light using a continuously rotating optical resonator. Phys. Rev. Lett. 95, 150401 (2005)

    Article  ADS  Google Scholar 

  28. Troshkin, O.V.: Wave properties of a turbulent fluid. Physica A168, 881 (1990)

    Article  ADS  Google Scholar 

  29. Puthoff, H.E.: Linearized turbulent flow as an analog model for linearized general relativity. arXiv:0808.3404

  30. Tsankov, T.D.: Classical electrodynamics and the turbulent Aether hypothesis (2009)

  31. Consoli, M., Pluchino, A., Rapisarda, A.: Basic randomness of nature and ether-drift experiments. Chaos, Solitons Fractals 44, 1089 (2011)

    Article  ADS  Google Scholar 

  32. Consoli, M.: A kinetic basis for space-time symmetries. Phys. Lett. A 376, 3377 (2012)

    Article  ADS  Google Scholar 

  33. Consoli, M., Pluchino, A., Rapisarda, A., Tudisco, S.: The vacuum as a form of turbulent fluid: motivations, experiments, implications. Physica A394, 61 (2014)

    Article  ADS  Google Scholar 

  34. Kolmogorov, A.N.: Dokl. Akad. Nauk SSSR 10, 4 (1940), English translation: The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. Proc. R. Soc. 434, 9 (1991)

  35. Landau, L.D., Lifshitz, E.M.: Fluid Mechanics, Chapt. III. Pergamon Press, London (1959)

    Google Scholar 

  36. Fung, J.C.H., et al.: Kinematical simulation of homogeneous turbulence by unsteady random Fourier modes. J. Fluid Mech. 236, 281 (1992)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Joos, G.: Die Jenaer Wiederholung des Michelsonversuchs. Ann. d. Phys. 7, 385 (1930)

    Article  ADS  Google Scholar 

  38. Jaseja, T.S., et al.: Test of special relativity or of the isotropy of space by use of infrared masers. Phys. Rev. 133, A1221 (1964)

    Article  ADS  Google Scholar 

  39. Brillet, A., Hall, J.L.: Improved laser test of the isotropy of space. Phys. Rev. Lett. 42, 549 (1979)

    Article  ADS  Google Scholar 

  40. Herrmann, S., et al.: Rotating optical cavity experiment testing Lorentz invariance at the \(10^{-17}\) level. Phys. Rev. D 80, 105011 (2009)

    Article  ADS  Google Scholar 

  41. Eisele, Ch., Newski, A., Schiller, S.: Laboratory test of the isotropy of light propagation at the \(10^{-17}\) level. Phys. Rev. Lett. 103, 090401 (2009)

    Article  ADS  Google Scholar 

  42. Müller, H., et al.: Modern Michelson–Morley experiment using cryogenic optical resonators. Phys. Rev. Lett. 91, 020401 (2003)

    Article  Google Scholar 

  43. Antonini, P., Okhapkin, M., Göklu, S., Schiller, S.: Test of constancy of speed of light with rotating cryogenic optical resonators. Phys. Rev. A71, 050101(R) (2005)

    Article  ADS  Google Scholar 

  44. Consoli, M., Pappalardo, L.: Emergent gravity and ether-drift experiments. Gen. Relat. Gravit. 42, 2585 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Shankland, R.S., et al.: New analysis of the interferometer observations of Dayton C. Miller. Rev. Mod. Phys. 27, 167 (1955)

    Article  ADS  Google Scholar 

  46. Colladay, D., Kostelecky, V.A.: CPT violation and the standard model. Phys. Rev. D 55, 6760 (1997)

    Article  ADS  Google Scholar 

  47. Kostelecky, V.A., Mewes, M.: Signals for Lorentz violation in electrodynamics. Phys. Rev. D 66, 056005 (2002)

    Article  ADS  Google Scholar 

  48. Kostelecky, V.A., Russell, N.: Data tables for Lorentz and CPT violation. Rev. Mod. Phys. 83, 11 (2011)

    Article  ADS  Google Scholar 

  49. Müller, H.: Testing Lorentz invariance by the use of vacuum and matter filled cavity resonators. Phys. Rev. D 71, 045004 (2005)

    Article  ADS  Google Scholar 

  50. Eisele, Ch., et al.: A crossed optical cavities apparatus for a precision measurement of the isotropy of light propagation. Opt. Commun. 281, 1189 (2008)

    Article  ADS  Google Scholar 

  51. De Abreu, R., Guerra, V.: Relativity-Einstein’s lost frame. Extra-Muros, Lisboa (2005)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Maurizio Consoli.

Appendices

Appendix 1

Let us consider light propagation in a gaseous system of refractive index \(\mathcal{N}\). By assuming isotropy, the time \(t\) spent by refracted light to cover some given distance \(L\) within the medium is \(t=\mathcal{N}L/c\). This can be expressed as the sum of \(t_0=L/c\) and \(t_1=(\mathcal{N}-1)L/c\) where \(t_0\) is the same time as in the vacuum and \(t_1\) represents the additional, average time by which refracted light is slowed down by the presence of matter. If there are convective currents, due to the motion of the laboratory with respect to a preferred reference frame \(\Sigma \), then \(t_1\) will be different in different directions, and there will be an anisotropy of the velocity of light proportional to \((\mathcal{N}-1)\). In fact, let us consider light propagating in a 2-dimensional plane and express \(t_1\) as

$$\begin{aligned} t_1={{L}\over {c}}f(\mathcal{N}, \theta , \beta ) \end{aligned}$$
(46)

with \(\beta =V/c\), \(V\) being (the projection on the considered plane of) the relevant velocity with respect to \(\Sigma \) where the isotropic form

$$\begin{aligned} f(\mathcal{N}, \theta , 0)=\mathcal{N}-1 \end{aligned}$$
(47)

is assumed. By expanding around \(\mathcal{N}=1\) where, whatever \(\beta \), \(f\) vanishes by definition, one finds for gaseous systems (where \(\mathcal{N}-1 \ll 1\)) the universal trend

$$\begin{aligned} f(\mathcal{N}, \theta ,\beta )\sim (\mathcal{N}-1)F(\theta ,\beta ) \end{aligned}$$
(48)

with

$$\begin{aligned} F(\theta ,\beta )\equiv (\partial f/\partial \mathcal{N})|_{ \mathcal{N}=1} \end{aligned}$$
(49)

and \(F(\theta ,0)=1\). Therefore, by introducing the one-way velocity of light

$$\begin{aligned} t(\mathcal{N},\theta ,\beta )= {{L}\over {c_\gamma (\mathcal{N},\theta ,\beta )}}\sim {{L}\over {c}}+ {{L}\over {c}}(\mathcal{N}-1)~F(\theta ,\beta ) \end{aligned}$$
(50)

one gets

$$\begin{aligned} c_\gamma (\mathcal{N},\theta ,\beta )\sim {{c}\over { \mathcal{N} }}~ \left[ 1- (\mathcal{N}-1) ~(F(\theta ,\beta ) -1)\right] \end{aligned}$$
(51)

Analogous relations hold for the two-way velocity \( \bar{c}_\gamma (\mathcal{N},\theta ,\beta )\)

$$\begin{aligned} \bar{c}_\gamma (\mathcal{N},\theta ,\beta )&= {{2~c_\gamma (\mathcal{N},\theta ,\beta )c_\gamma (\mathcal{N},\pi +\theta ,\beta )}\over {c_\gamma (\mathcal{N},\theta ,\beta ) +c_\gamma (\mathcal{N},\pi +\theta ,\beta )}}\nonumber \\&\,\,\sim {{c}\over { \mathcal{N} }} \left[ 1- (\mathcal{N}-1) ~\left( {{F(\theta ,\beta ) + F(\pi +\theta ,\beta )}\over {2}} -1\right) \right] \end{aligned}$$
(52)

A more explicit expression can be obtained by exploring some general properties of the function \(F(\theta ,\beta )\). By expanding in powers of \(\beta \)

$$\begin{aligned} F(\theta ,\beta )-1 = \beta F_1(\theta ) + \beta ^2 F_2(\theta )+ \cdots \end{aligned}$$
(53)

and taking into account that, by the very definition of two-way velocity, \(\bar{c}_\gamma (\mathcal{N},\theta ,\beta )= \bar{c}_\gamma (\mathcal{N},\theta ,-\beta )\), it follows that \(F_1(\theta )=-F_1(\pi + \theta )\). Therefore, to \(\mathcal{O}(\beta ^2)\), by expressing the combination \(F_2(\theta ) + F_2(\pi +\theta )\) as an infinite expansion of even-order Legendre polynomials, we get the general structure

$$\begin{aligned} \bar{c}_\gamma (\mathcal{N},\theta ,\beta ) \sim {{c}\over { \mathcal{N} }} \left[ 1- (\mathcal{N}-1)~\beta ^2 \sum ^\infty _{n=0}\zeta _{2n}P_{2n}(\cos \theta ) \right] \end{aligned}$$
(54)

which coincides with Eq. (17).

Appendix 2

To derive Eq. (18), let us first consider a dielectric medium of refractive index \(\mathcal{N}\) whose container is at rest in \(\Sigma \), the reference frame defined by imposing zero spatial momentum for the vacuum condensation phenomenon. For an observer at rest in this reference frame, light propagation within the medium is assumed to be isotropic and described by

$$\begin{aligned} \pi _\mu \pi _\nu \gamma ^{\mu \nu }= 0 \end{aligned}$$
(55)

where

$$\begin{aligned} \gamma ^{\mu \nu }=\mathrm{diag}(\mathcal{N}^2,-1,-1,-1) \end{aligned}$$
(56)

and \(\pi _\mu \) denotes the light 4-momentum vector for the \(\Sigma \) observer. Let us now consider that the container of the medium is moving with some velocity \(\mathbf{V}\) with respect to \(\Sigma \) and is at rest in some other frame \(S'\). By analogy, light propagation within the medium for the observer in \(S'\) will be described by

$$\begin{aligned} p_\mu p_\nu g^{\mu \nu }= 0 , \end{aligned}$$
(57)

where \(p_\mu \equiv (E/c,\mathbf{p})\) and \(g^{\mu \nu }\) denote respectively the light 4-momentum and the effective metric for \(S'\). On this basis, by introducing the \(S'\) dimensionless velocity 4-vector \(u^\mu \equiv (u^0,\mathbf{V}/c)\) (with \(u_\mu u^\mu =1\)), let us define \(A^{\mu }{_\nu }=A^{\mu }{_\nu }(u_\mu ,\mathcal{N}) \) to be the transformation matrix which produces \(g^{\mu \nu }\) from the reference isotropic metric \(\gamma ^{\mu \nu }\), i.e.

$$\begin{aligned} g^{\mu \nu }=A^{\mu }{_\sigma }A^{\nu }{_\rho }\gamma ^{\sigma \rho } \end{aligned}$$
(58)

In this context, requiring consistency between vacuum condensation and Special Relativity corresponds to place all reference frames on the same footing and assume \(g^{\mu \nu }=\gamma ^{\mu \nu }\) or (SR=Special Relativity)

$$\begin{aligned} \left[ A^{\mu }{_\nu }(u_\mu ,\mathcal{N})\right] _\mathrm{SR}=\delta ^{\mu }_{\nu } \end{aligned}$$
(59)

In Special Relativity, where there is no preferred reference frame, this identification is independent of the physical nature of the medium, being valid for gaseous systems as well as for liquid or solid transparent media.

On the other hand, let us now consider the case where the medium inside the container becomes more and more rarefied, i.e. let us take the \(\mathcal{N} \rightarrow 1\) limit. Here, the usual identification of the transformation matrix is instead

$$\begin{aligned} A^{\mu }{_\nu }(u_\mu ,\mathcal{N}=1)=\Lambda ^{\mu }_{\nu } \end{aligned}$$
(60)

\(\Lambda ^{\mu }_{\nu }\) being the Lorentz transformation matrix from \(\Sigma \) to \(S'\). In fact, this is the standard choice to solve non-trivially the equation \(g^{\mu \nu }=\gamma ^{\mu \nu }\) when light propagates in vacuum and \(\gamma ^{\mu \nu }\) reduces to the Minkowski tensor \(\eta ^{\mu \nu }\). But then, by continuity, it is conceivable that, to describe the case \(\mathcal{N}=1+ \epsilon \) of gaseous media, the right transformation matrix is closer to Eq. (60) rather than to Eq. (59). Otherwise, it would be very hard to justify the sudden change from Eq. (60) to

$$\begin{aligned} \left[ A^{\mu }{_\nu }(u_\mu ,\mathcal{N}=1+\epsilon )\right] _\mathrm{SR}=\delta ^{\mu }_{\nu } \end{aligned}$$
(61)

for any \(\epsilon > 0\) and for any value of the \(S'\) velocity.

Notice however that no contradiction arises once \(\mathcal{N}\) starts to differ substantially from unity. In this other limit, one can reconcile Eqs. (59) and (60). As discussed in Sects. 1 and 2, the possibility of two different regimes is in agreement with the idea of a tiny vacuum energy–momentum flow associated with a Lorentz non-invariant vacuum state. This flow, acting as an effective thermal gradient, could induce convective currents of the gas molecules along the light path and a non-zero anisotropy of refracted light. The point is that, differently from strongly bound systems, such as solid or liquid transparent media, where \(\mathcal{N}\) differs substantially from unity, the elementary constituents of such weakly bound matter can be set in motion by an arbitrarily small thermal gradient. This picture also agrees with the phenomenological pattern of experiments [21, 22] which indicates the existence of two different regimes. A former region of gaseous systems where \(\mathcal{N}\sim 1\) and a latter region where the difference of \(\mathcal{N}\) from unity is substantial, e.g. \(\mathcal{N}\sim 1.5\) as with perspex in the experiment by Shamir and Fox [23]. Although it would be difficult to describe in a fully quantitative way the transition between the two regimes, some simple arguments can be given along the lines suggested by de Abreu and Guerra (see pages 165–170 of Ref. [51]).

Thus we conclude that, for \(\mathcal{N}=1+ \epsilon \) and with a Lorentz non-invariant vacuum, it makes more sense to compute \(g^{\mu \nu }\) through the relation

$$\begin{aligned} g^{\mu \nu }=\Lambda ^{\mu }{_\sigma }\Lambda ^{\nu }{_\rho }\gamma ^{\sigma \rho } \end{aligned}$$
(62)

rather than setting \(g^{\mu \nu }= \gamma ^{\mu \nu }\). This gives the effective metric for \(S'\) (\(\kappa =\mathcal{N}^2 - 1\))

$$\begin{aligned} g^{\mu \nu }= \eta ^{\mu \nu } + \kappa u^\mu u^\nu \end{aligned}$$
(63)

In this way, Eq. (57) gives a photon energy (\(u^2_0=1 + \mathbf{V}^2/c^2\))

$$\begin{aligned} E(| {\mathbf {p}}| , \theta )= c~{{ -\kappa u_0 \zeta + \sqrt{ |{\mathbf {p}}|^2(1+\kappa u^2_0) - \kappa \zeta ^2 }}\over { 1 + \kappa u^2_0}} \end{aligned}$$
(64)

with

$$\begin{aligned} \zeta ={\mathbf {p}}\cdot {{{\mathbf {V}}}\over {c}}= |{\mathbf {p}}|\beta \cos \theta , \end{aligned}$$
(65)

where \(\beta ={{|{\mathbf {V}}|}\over {c}}\) and \(\theta \equiv \theta _\mathrm{lab}\) indicates the angle defined, in the laboratory \(S'\) frame, between the photon momentum and \({\mathbf {V}}\). By using the above relation, one gets the one-way velocity of light

$$\begin{aligned}&{{E(| {\mathbf {p}}| , \theta )}\over {|{\mathbf {p}}|}}= c_\gamma (\theta )= c~{{ -\kappa \beta \sqrt{1+\beta ^2} \cos \theta + \sqrt{ 1+ \kappa + \kappa \beta ^2 \sin ^2\theta } } \over {1+ \kappa (1+\beta ^2)}} . \end{aligned}$$
(66)

or to \(\mathcal{O}(\kappa )\) and \(\mathcal{O}(\beta ^2)\)

$$\begin{aligned} c_\gamma (\theta ) \sim {{c} \over {\mathcal{N}}}~\left[ 1- \kappa \beta \cos \theta - {{\kappa }\over {2}} \beta ^2(1+\cos ^2\theta )\right] \end{aligned}$$
(67)

From this one can compute the two-way velocity (\(\kappa \sim 2(\mathcal{N} - 1) \equiv 2\epsilon \))

$$\begin{aligned} \bar{c}_\gamma (\theta )&= {{ 2 c_\gamma (\theta ) c_\gamma (\pi + \theta ) }\over { c_\gamma (\theta ) + c_\gamma (\pi + \theta ) }} \nonumber \\&\sim {{c} \over {\mathcal{N}}}~\left[ 1-\epsilon \beta ^2\left( 2 - \sin ^2\theta \right) \right] \end{aligned}$$
(68)

which, as anticipated, is a special form of the more general Eq. (17).Footnote 7

One conceptual detail concerns the gas refractive index whose reported values are experimentally measured on Earth by two-way measurements. For instance, for air the most precise determinations are at the level \(10^{-7}\), say \(\mathcal{N}_\mathrm{air}=1.0002926..\) at STP (Standard Temperature and Pressure). By assuming a non-zero anisotropy in the Earth’s frame, one should interpret the isotropic value \(c/\mathcal{N_\mathrm{air}}\) as an angular average of Eq.(68), i.e.

$$\begin{aligned} {{c}\over { \mathcal{N}_\mathrm{air} }}\equiv \langle \bar{c}_\gamma (\bar{\mathcal{N}}_\mathrm{air},\theta ,\beta )\rangle _\theta = {{c}\over { \bar{\mathcal{N}} _\mathrm{air} }} ~[1-{{3}\over {2}} (\bar{\mathcal{N}}_\mathrm{air} -1)\beta ^2] \end{aligned}$$
(69)

From this relation, one can determine in principle the unknown value \(\bar{\mathcal{N}} _\mathrm{air} \equiv \mathcal{N}(\Sigma )\) (as if the gas were at rest in \(\Sigma \)), in terms of the experimentally known quantity \(\mathcal{N}_\mathrm{air}\equiv \mathcal{N}(Earth)\) and of \(V\). In practice, for the standard velocity values involved in most cosmic motions, say \( V \sim \) 300 km/s, the difference between \(\mathcal{N}(\Sigma )\) and \(\mathcal{N}(Earth)\) is at the level \(10^{-9}\) and thus completely negligible. The same holds true for the other gaseous systems at STP (say nitrogen, carbon dioxide, helium,..) for which the present experimental accuracy in the refractive index is, at best, at the level \(10^{-6}\). Finally, the isotropic two-way speed of light is better determined in the low-pressure limit where \((\mathcal{N}-1)\rightarrow 0\). In the same limit, for any given value of \(V\), the approximation \(\mathcal{N}(\Sigma )=\mathcal{N}(Earth)\) becomes better and better.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Consoli, M. Probing the Vacuum of Particle Physics with Precise Laser Interferometry. Found Phys 45, 22–43 (2015). https://doi.org/10.1007/s10701-014-9849-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-014-9849-2

Keywords

Navigation