zotero/storage/ZZ6AAADX/.zotero-ft-cache

3557 lines
152 KiB
Plaintext
Raw Blame History

This file contains ambiguous Unicode characters

This file contains Unicode characters that might be confused with other characters. If you think that this is intentional, you can safely ignore this warning. Use the Escape button to reveal them.

arXiv:2402.12421v1 [physics.gen-ph] 19 Feb 2024
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
International Journal of Modern Physics A © World Scientific Publishing Company
Michelson-Morley Experiments: at the crossroads of Relativity, Cosmology and Quantum Physics
Maurizio Consoli Istituto Nazionale di Fisica Nucleare, Sezione di Catania
Via S.Sofia 64, 95123, Catania, Italy maurizio.consoli@ct.infn.it
Alessandro Pluchino Dipartimento di Fisica e Astronomia E.Majorana, University of Catania
and Istituto Nazionale di Fisica Nucleare, Sezione di Catania Via S.Sofia 64, 95123, Catania, Italy alessandro.pluchino@ct.infn.it
Today, the original Michelson-Morley experiment and its early repetitions at the beginning of the 20th century are considered as a venerable historical chapter for which, at least from a physical point of view, there is nothing more to refine or clarify. The emphasis is now on the modern versions of these experiments, with lasers stabilized by optical cavities, that, apparently, have improved by many orders of magnitude on the limits placed by those original measurements. Though, in those old experiments light was propagating in gaseous systems (air or helium at atmospheric pressure) while now, in modern experiments, light propagates in a high vacuum or inside solid dielectrics. Therefore, in principle, the difference might not depend on the technological progress only but also on the different media that are tested by preventing a straightforward comparison. Starting from this observation, one can formulate a new theoretical scheme where the tiny, irregular residuals observed so far, from Michelson-Morley to the present experiments with optical resonators, point consistently toward the long sought preferred reference frame tight to the CMB. The existence of this scheme, while challenging the traditional null interpretation, presented in all textbooks and specialized reviews as a self-evident scientific truth, further emphasizes the central role of these experiments for Relativity, Cosmology and Quantum Physics. Keywords: Relativity; preferred reference system; quantum nonlocality
1
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
2 Authors Names
1. Introduction From the very beginning there are two interpretations of Relativity: Einsteins Special Relativity1 and the Lorentzian formulation.2 Apart from all historical aspects, the difference could simply be phrased as follows. In a Lorentzian approach, the relativistic effects originate from the individual motion of each observer S, S”...with respect to some preferred reference frame Σ, a convenient redefinition of Lorentz ether. Instead, according to Einstein, eliminating the concept of the ether leads to interpret the same effects as consequences of the relative motion of each pair of observers S and S”. This is possible because the basic quantitative ingredients, namely Lorentz Transformations, have a crucial group structure and are the same in both formulations. In the case of one-dimensional motion a, an intuitive representation is given in Fig.1.
Fig. 1. An intuitive representation of the two interpretations of Relativity.
For this reason, it has been generally assumed that there is a substantial phenomenological equivalence of the two formulations. This point of view was, for instance, already clearly expressed by Ehrenfest in his lecture On the crisis of the light ether hypothesis (Leyden, December 1912) as follows: “So, we see that the ether-less theory of Einstein demands exactly the same here as the ether theory of Lorentz. It is, in fact, because of this circumstance, that according to Einsteins theory an observer must observe exactly the same contractions, changes of rate, etc.
aWe ignore here the subtleties related to the Thomas-Wigner spatial rotation which is introduced when considering two Lorentz transformations along different directions, see e.g.35
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 3
in the measuring rods, clocks, etc. moving with respect to him as in the Lorentzian theory. And let it be said here right away and in all generality. As a matter of principle, there is no experimentum crucis between the two theories”. Therefore, by assuming that, in a Lorentzian perspective, the motion with respect to Σ could not be detected, the usual attitude was to consider the difference between the two interpretations as a philosophical problem.
However, it was emphasized by Bell6 that adopting the Lorentzian point of view could be crucial to reconcile hypothetical faster-than-light signals with causality, as with the apparent non-local aspects of the Quantum Theory. Indeed, if all reference frames are placed on the same footing, as in Special Relativity, how to decide of the time ordering of two events A and B along the world line of a hypothetical effect propagating with speed > c? This ordering can be different in different frames, because in some frame S one could find tA > tB and in some other frame S the opposite tB > tA . This causal paradox, which is the main reason why superluminal signals are not believed to exist, disappears in a Lorentzian formulation where the different views of the two observers become a sort of optical illusion, like an aberrationb.
But the mere logical possibility of Σ is not enough. For a full resolution of the paradox, the Σ− frame should show up through a determination of the kinematic parameters β′, β′′... Thus, we arrive to the main point of this article: the prejudice that, even in a Lorentzian formulation of relativity, the individual β′, β′′... cannot be experimentally determined. This belief derives from the assumption that the Michelson-Morley type of experiments, from the original 1887 trial to the modern versions with lasers stabilized by optical cavities, give null results, namely that the small residuals found in these measurements are just typical instrumental artifacts. We recall that in these precise interferometric experiments, one attempts to detect in laboratory an ether-wind, i.e. a small angular dependence of the velocity of light that might indicate the Earth motion with respect to the hypothetical Σ, e.g. the system where the Cosmic Microwave Background (CMB) is isotropic. While in Special Relativity, no ether wind can be observed by definition, in a Lorentzian perspective it is only a conspiracy of relativistic effects which makes undetectable the individual velocity parameters β′, β′′... But the conspiracy works exactly only when the velocity of light cγ propagating in the various interferometers coincides with the basic parameter c entering Lorentz transformations. Therefore, one may
bIf S is connected to Σ by a Lorentz transformation with parameter β′ = v/c, by the inverse transformation we can find the time coordinates in Σ starting from xA, ctA, xB and ctB , namely cTA = γxA + ctA) and cTB = γxB + ctB ), with 1/γ = 1 (β′)2. Analogously, for S and parameter β′′ = v/c, we will find the same values, i.e. cTA = γxA + ctA ) and cTB = γxB + ctB ), with now 1/γ = 1 (β′′)2. Thus no ambiguity is possible, either TA > TB or viceversa so that the view in the preferred Σframe becomes the relevant one to decide on causal effects.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
4 Authors Names
ask, what happens if cγ ̸= c, for instance when light propagates in air or in gaseous helium as in the old experiments? Starting from this observation, we have formulated a new theoretical scheme710 where the small residuals observed so far, from Michelson-Morley to the present experiments with optical resonators, point consistently toward the long sought preferred reference frame tight to the CMB. In this sense, our scheme is seriously questioning the standard null interpretation of these experiments which is presented in all textbooks and specialized reviews as a self-evident scientific truth. In this article we will review the main results of our extensive work and also propose further experimental tests.
We emphasize that, besides Relativity, our reinterpretation of the data intertwines with and influences other areas of contemporary physics, such as the nonlocality of the Quantum Theory, the current vision of the Vacuum State and Cosmology. These implications are so important to deserve a preliminary discussion in this Introduction.
1.1. Relativity and Quantum Non-Locality
The existence of intrinsically non-local aspects in the Quantum Theory and the relationship with relativity has been the subject of a countless number of books and articles, growing more and more rapidly in recent times, see e.g.1113 for a list of references. The problem dates back to the very early days of Quantum Mechanics, even before the seminal work of Einstein-Podolski-Rosen (EPR).14 Indeed, the basic issue is already found in Heisenbergs 1929 Chicago Lectures: “ We imagine a photon represented by a wave packet... By reflection at a semi-transparent mirror, it is possible to decompose into a reflected and a transmitted packet...After a sufficient time the two parts will be separated by any distance desired; now if by experiment the photon is found, say, in the reflected part of the packet, then the probability of finding the photon in the other part of the packet immediately becomes zero. The experiment at the position of the reflected packet thus exerts a kind of action (reduction of the wave packet) at the distant point and one sees that this action is propagated with a velocity greater than that of light”. After that, Heisenberg, almost frightened by his same words, feels the need to add the following remark: “However, it is also obvious that this kind of action can never be utilized for the transmission of signals so that it is not in conflict with the postulates of relativity”.
Heisenbergs final observation is one of the first formulations of the so called peaceful coexistence. Actually, presenting as an obvious fact that this type of effects can never be used to communicate between observers at a space-like separation sounds more as a way to avoid the causal paradox, which is present in Special Relativity, when dealing with faster than light signals. But, independently of that, this observation expresses a position that can hardly be considered satisfactory. In fact, if there were really some Quantum Information which propagates with a speed vQI ≫ c, could such extraordinary thing be so easily dismissed? Namely, could we ignore this something just because, apparently, it cannot be efficiently controlled
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 5
to send messagesc? After all, this explains why Dirac, more than forty years later, was still concluding that “The only theory which we can formulate at the present is a non-local one, and of course one is not satisfied with such a theory. I think one ought to say that the problem of reconciling quantum theory and relativity is not solved”.17
But only with Bells contribution6 the real terms of the problem were fully understood. He clearly spelled out the local, realistic point of view. If physical influences must propagate continuously through space, it becomes unavoidable to complete the quantum formalism by introducing additional hidden variables associated with the space-time regions in question d. But, then, it is possible to derive a bound on the degree of correlation of physical systems that are no longer interacting but have interacted in their past. This bound has been used to rule out experimentally1820 the class of local, hidden-variable theories which are based on causal influences propagating at subluminal speed. Experimentally excluding this class of theories means rejecting a familiar vision of reality. Thanks to Bell, “A seemingly philosophical debate about the nature of physical reality could be settled by an experiment! ...The conclusion is now clear: Einsteins view of physical reality cannot be upheld”.21
Thus, the importance of Bells work cannot be underestimated: “Bells result combined with the EPR argument was not a no hidden variables theorem but a non-locality theorem, the impossibility of hidden variables being only one step in a two-step argument...It means that some action at a distance exists in Nature, even though it does not say what this action consists of”.13 It was this awareness to give him the perception that “... we have an apparent incompatibility, at the deepest level, between the two fundamental pillars of contemporary theory”,6 namely Quantum Theory and Special Relativity. This inspired his view where the existence of the preferred Σframe would free ourselves from the no-signalling argument to dispose of the causality paradox.
cExperimental correlations between spacelike separated measurements can in principle be explained through hidden influences propagating at a finite speed vQI ≫ c provided vQI is large enough.15 But in ref.16 it is also shown that for any finite vQI, with c < vQI < ∞, one can construct combined correlations to be used for faster-than-light communication. d“In particular, Jordan had been wrong in supposing that nothing was real or fixed in the microscopic world before observation. For after observing only one of the two particles the result of subsequently observing the other (possibly at very remote place) is immediately predictable. Could it be that the first measurement somehow fixes what was unfixed or makes real what was unreal, not only for the near particle but also for the remote one? For EPR that would be an unthinkable spooky action at distance. To avoid such action at distance one has to attribute, to the space-time regions in question, real correlated properties in advance of the observation which predetermine the outcome of these particular observations. Since these real properties, fixed in advance of the observation, are not contained in the quantum formalism, that formalism for EPR is incomplete”.6
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
6 Authors Names
1.2. Relativity and the Vacuum State
A frequent objection to the idea of relativity with a preferred frame is that, after all, Quantum Mechanics is not a fundamental description of the world. What about, if we started from a more fundamental Quantum Field Theory (QFT)? In this perspective, the issue of the preferred frame can be reduced to find a particular, logical step that prevents to deduce that Einstein Special Relativity, with no preferred frame, is the physically realized version of relativity. This is the version which is always assumed when computing S-matrix elements for microscopic processes. However, what one is actually using is the machinery of Lorentz transformations whose first, complete derivation dates back, ironically, to Larmor and Lorentz who were assuming the existence of a fundamental state of rest (the ether).
Our point, discussed in,2225 is that there is indeed a particular element which has been missed so far and which concerns the assumed Lorentz invariance of the vacuum state. Even though one is still using the Latin word vacuum, which means empty, here we are dealing with the lowest energy state. According to the present view, this is not trivially empty but is determined by the condensation process of some elementary quanta e. Namely the energy is minimized when these quanta, such as Higgs particles, quark-antiquark pairs, gluons... macroscopically occupy the same quantum state, i.e. the zero-3-momentum state f . Thus, if the condensation process singles out a certain reference frame Σ, the fundamental question is how to reconcile this picture with the basic postulate of axiomatic QFT: the exact Lorentz invariance of the vacuum.28 This postulate, meaning that the vacuum state must remain unchanged under Lorentz boost, should not be confused with the condition that only local scalars (as the Higgs field, or the gluon condensate, or the chiral condensate...) acquire a non-zero vacuum expectation value.
To make this evident, let us introduce the reference vacuum state |Ψ(0)⟩, appropriate to the observer at rest in the Σframe singled out by the condensation process, and the corresponding vacuum states |Ψ′⟩, |Ψ′′⟩,.. appropriate to moving observers S, S,... By assuming that these different vacua are generated by
eBefore our work, the idea that the phenomenon of vacuum condensation could produce conceptual tensions with both Special and General Relativity, was discussed by Chiao:26 “The physical vacuum, an intrinsically nonlocal ground state of a relativistic quantum field theory, which possesses certain similarities to the ground state of a superconductor... This would produce an unusual quantum rigidity of the system, associated with what London called the rigidity of the macroscopic wave function... The Meissner effect is closely analog to the Higgs mechanism in which the physical vacuum also spontaneously breaks local gauge invariance ”.26 f In the physically relevant case of the Standard Model, the phenomenon of vacuum condensation can be summarized by saying that “What we experience as empty space is nothing but the configuration of the Higgs field that has the lowest possible energy. If we move from field jargon to particle jargon, this means that empty space is actually filled with Higgs particles. They have Bose condensed”.27
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 7
Lorentz boost unitary operators U , U ...acting on |Ψ(0)⟩, say |Ψ′⟩ = U |Ψ(0)⟩, |Ψ′′⟩ = U |Ψ(0)⟩... For any Lorentz-scalar operator G, such that G = (U )†GU = (U )†GU ..., it follows trivially
⟨G⟩Ψ(0) = ⟨G⟩Ψ = ⟨G⟩Ψ = ..
(1)
However, this by no means implies |Ψ(0)⟩ = |Ψ′⟩ = |Ψ′′⟩... To this end, one should
construct explicitly the three boost generators L0i (with i=1,2,3) and show that L0i|Ψ(0)⟩ = 0. But, in four space-time dimensions, the explicit construction of these operators, and of the corresponding Poincar´e algebra g is only known for the free-
field case through the simple Wick-ordering prescription relatively to the free-field
vacuum |0⟩. In an interacting theory, the construction is implemented order by
order in perturbation theory. Therefore, in the presence of non-perturbative phe-
nomena (such as Spontaneous Symmetry Breaking, chiral symmetry breaking, gluon condensation...) where the physical vacuum |Ψ(0)⟩ cannot be constructed from the
free-field vacuum |0⟩ order by order in perturbation theory, proving the Lorentz
invariance of the vacuum represents an insurmountable problem. In this situation,
with a Lorentz-invariant interaction, the resulting theory will still be Lorentz covari-
ant but, with a non-Lorentz-invariant vacuum, there would be a preferred reference frame h.
1.3. Relativity and the CMB
Finally, some remarks about the physical nature of the hypothetical Σframe. A natural candidate is the reference system where the temperature of the CMB looks exactly isotropic or, more precisely, where the CMB kinematic dipole30 vanishes. This dipole is in fact a direct consequence of the motion of the Earth (β = V /c)
T (θ) = To 1 β2
(2)
1 β cos θ
gThis means an operatorial representation of the 10 generators Pµ and Lµν (µ, ν= 0, 1, 2, 3), where Pµ describe the space-time translations and Lµν = Lνµ the space rotations and Lorentz boosts, with commutation relations [Pµ, Pν ] = 0, [Lµν , Pρ] = iηνρ iηµρPν and [Lµν , Lρσ] = iηµρLνσ + iηµσLνρ νσρ + iηνρσ where ηµν = diag(1, 1, 1, 1) is the Minkowski tensor. A Lorentz-invariant vacuum has to be annihilated by all 10 generators. hTo our knowledge,in four space-time dimensions, a non-perturbative analysis of a Lorentzinvariant vacuum has been attempted by very few authors. In the case of non-linear field theories with P (Φ(x)) interactions, such as Φ4(x), this was discussed by Segal.29 He considered a suitable generalization of the standard Wick ordering : P (Φ) : relative to |0⟩, say :: P (Φ) ::, such that ⟨Ψ(0)| :: P (Φ) :: |Ψ(0)⟩ = 0 in the true vacuum state. His conclusion was that :: P (Φ) :: is not well-defined until the physical vacuum is known, but, at the same time, the physical vacuum also depends on the definition given for :: P (Φ) ::. From this type of circularity Segal concluded that, in general, in such a nonlinear QFT, the physical vacuum will not be invariant under the full Lorentz symmetry of the underlying Lagrangian density.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
8 Authors Names
Accurate observations with satellites in space31 have shown that the measured temperature variations correspond to a motion of the solar system described by an average velocity V 370 km/s, a right ascension α 168o and a declination γ 7o, pointing approximately in the direction of the Leo constellation. This means that, if one sets To 2.7 K and β 0.00123, there are angular variations of a few millikelvin
∆T CMB(θ) Toβ cos θ = ±3.3 mK
(3)
which represent by far the dominant contribution to the CMB anisotropy. Could the reference system with vanishing CMB dipole represent the funda-
mental preferred frame for relativity as in the original Lorentzian formulation? The standard answer is that one should not confuse these two concepts. The CMB is a definite medium and, as such, sets a rest frame where the dipole anisotropy is zero. Our motion with respect to this system has been detected but, by itself, this is not in contradiction with Special Relativity. Though, to good approximation, this kinematic dipole arises by combining the various forms of peculiar motion which are involved (rotation of the solar system around the center of the Milky Way, motion of the Milky Way toward the center of the Local Group, motion of the Local Group of galaxies in the direction of the Great Attractor...)31 . Thus, if one could switch-off the local inhomogeneities that produce these peculiar forms of motion, it is natural to imagine a global frame of rest associated with the Universe as a whole. A vanishing CMB dipole could then just indicate the existence of this fundamental system Σ that we may conventionally decide to call ether but the cosmic radiation itself would not coincide with this form of ether.
This is why, to discriminate between the two concepts, Michelson-Morley type of experiments become crucial. Detecting a small angular dependence of the velocity of light in the Earth laboratory, and correlating this angular dependence with the Earth cosmic motion, would provide the missing link with the logical arguments from Quantum Non-Locality i and with the idea of a condensed vacuum which selects a particular reference frame through the macroscopic occupation of the same zero 3-momentum state. More generally a non-null interpretation of the MichelsonMorley experiments would resolve the puzzle of a world endowed with a fundamental space and a fundamental time whose existence, otherwise, would remain forever hidden to us.
After this general Introduction, we will start by reviewing in Sect.2 the basic ingredients for a modern analysis of the Michelson-Morley experiments. Then we will summarize in Sects.3 and 4 our re-analysis710 of the classical experiments
i“Non-Locality is most naturally incorporated into a theory in which there is a special frame of reference. One possible candidate for this special frame of reference is the one in which the CMB is isotropic. However, other than the fact that a realistic interpretation of quantum mechanics requires a preferred frame and the CMB provides us with one, there is no readily apparent reason why the two should be linked”.32
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 9
and in Sect.5 the corresponding treatment of the present experiments with optical resonators. As a matter of fact, once the small residuals are analyzed in our scheme, the long sought Σframe tight to the CMB is naturally emerging. Sect.6 will finally contain a summary and our conclusions.
2. A modern view of the ether-drift experiments The Michelson-Morley experiments are also called ether-drift experiments because they were designed to detect the drift of the Earth in the ether by observing a dragging of light associated with the Earth cosmic motion. Today, experimental evidence for both the undulatory and corpuscular aspects of radiation has substantially modified the consideration of an underlying ethereal medium, as support of the electromagnetic waves, and its logical need for the physical theory. Besides, Lorentz Transformations forbid dragging and the irregular behavior of the small observed residuals is inconsistent with the smooth time modulations that one would expect from the Earth rotation. Therefore, at first sight, the idea of detecting a non-null effect seems hopeless.
However, as anticipated, dragging is only forbidden if the velocity of light in the interferometers is the same parameter c of Lorentz transformations. For instance, when light propagates in a gas, the sought effect of a preferred system Σ could be due to the small fraction of refracted light. Obviously, this would be much smaller than classically expected but, in view of the extraordinary precision of the interferometers, it could still be measurable. In addition, the idea of smooth time modulations of the signal reflects the traditional identification of the local velocity field, which describes the drift, with the projection of the global Earth motion at the experimental site. This identification is equivalent to a form of regular, laminar flow where global and local velocity fields coincide. Instead, depending on the nature of the physical vacuum, the two velocity fields could only be related indirectly, as it happens in turbulent flows, so that numerical simulations would be needed for a consistent statistical description of the data.
In the following, we will summarize the scheme of refs.710 starting with the case of light propagating in gaseous media, as for the classical experiments.
2.1. Basics of the ether-drift experiments In the classical measurements in gases (Michelson-Morley, Miller, Tomaschek, Kennedy, Illingworth, Piccard-Stahel, Michelson-Pease-Pearson, Joos)33-,43 one was measuring the fringe shifts produced by a rotation of the interferometer. Instead, in modern experiments, with lasers stabilized by optical cavities, see e.g.44 for a review, one measures frequency shifts. The modern experiments adopt a different technology but, in the end, have exactly the same scope: searching for an anisotropy of the two-way velocity of light c¯γ(θ) which is the only one that can be measured
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
10 Authors Names
Fig. 2. A schematic illustration of the Michelson interferometer. Note that, by computing the transit times as in Eq.(6), we are assuming the validity of Lorentz transformations so that the length of a rod does not depend on its orientation in the frame S where it is at rest.
unambiguously
γ (θ)
=
2cγ(θ)cγ(π + θ) cγ(θ) + cγ(π + θ)
(4)
By introducing its anisotropy
∆c¯θ = c¯γ(π/2 + θ) γ(θ)
(5)
there is a simple relation with the time difference ∆t(θ) for light propagation back and forth along perpendicular rods of length D. In fact, by assuming the validity of Lorentz transformations, the length of a rod does not depend on its orientation, in the S frame where it is at rest, see Fig.2, and one finds,
2D
∆t(θ) =
2D
2D
∆c¯θ
(6)
γ(θ) c¯γ(π/2 + θ) c c
(where, in the last relation, we have assumed that light propagates in a medium of refractive index N = 1 + ϵ, with ϵ ≪ 1). This gives directly the fringe patterns (λ is the light wavelength)
∆λ(θ)
2D
∆c¯θ
(7)
λ
λc
which were measured in the classical experiments.
In modern experiments, on the other hand, a possible anisotropy of c¯γ(θ) would show up through the relative frequency shift, i.e. the beat signal, ∆ν(θ) of two
orthogonal optical resonators, see Fig.3. Their frequency
ν(θ) = c¯γ(θ)m
(8)
2L
is proportional to the two-way velocity of light within the resonator through an
integer number m, which fixes the cavity mode, and the length of the cavity L as
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 11
Fig. 3. The scheme of a modern ether-drift experiment. The light frequencies are first stabilized by coupling the lasers to Fabry-Perot optical resonators. The frequencies ν1 and ν2 of the resonators are then compared in the beat note detector which provides the frequency shift ∆ν(θ) = ν1(π/2 + θ) ν2(θ). For a review, see e.g.44
measured in the laboratory S frame. Therefore, once the length of a cavity in its rest frame does not depend on its orientation, one finds
∆ν(θ) ∆c¯θ
(9)
ν0
c
where ν0 is the reference frequency of the two resonators.
2.2. The limit of refractive index N = 1 + ϵ
Let us consider light propagating in a medium which is close to the ideal vacuum limit, i.e. whose refractive index is N = 1 + ϵ with ϵ ≪ 1. The medium (e.g. a gas) fills an optical cavity at rest in a frame S which moves with velocity v with respect to the hypothetical Σ. Now, by assuming i) that the velocity of light is exactly isotropic when S ≡ Σ and ii) the validity of Lorentz transformations, then any anisotropy in S should vanish identically either for v = 0 or for the ideal vacuum case N = 1 when the velocity of light cγ coincides with the basic parameter c of Lorentz transformations. Thus, one can expand in powers of the two small parameters ϵ and β = v/c. By taking into account that, by its very definition, the two-way velocity c¯γ(θ) is invariant under the replacement β → −β and that, for any fixed β, is also invariant under the replacement θ → π + θ, to lowest non-trivial level O(ϵβ2), one finds the general expression7, 24
c c¯γ(θ) N
1 ϵ β2 ζ2nP2n(cos θ)
(10)
n=0
Here, to take into account invariance under θ → π + θ, the angular dependence
has been given as an infinite expansion of even-order Legendre polynomials with
arbitrary coefficients ζ2n = O(1). In Einsteins Special Relativity, where there is
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
12 Authors Names
no preferred reference frame, these ζ2n coefficients should vanish identically. In a
Lorentzian approach, on the other hand, there is no reason why they should vanish
a priori j.
By leaving out the first few ζs as free parameters in the fits, Eq.(10) could
already represent a viable form to compare with experiments. Still, one can further
sharpen the predictions by exploiting one more derivation of the ϵ → 0 limit with
a preferred frame. This other argument is based on the effective space-time metric
ν = gµν (N ) which, through the relation gµν pµpν = 0, describes light propagation
in a medium of refractive index N . For the quantum theory, a derivation of this
metric from first principles was given by Jauch and Watson46 who worked out
the quantization of the electromagnetic field in a dielectric. They noticed that the
procedure introduces unavoidably a preferred reference frame, the one where the
photon energy spectrum does not depend on the direction of propagation, and which
is “usually taken as the system for which the medium is at rest”. However, such an
identification reflects the point of view of Special Relativity with no preferred frame.
Instead, one can adapt their results to the case where the angle-independence of
the photon energy is only valid when both medium and observer are at rest in some
particular frame Σ.
In this perspective, let us consider two identical optical cavities, namely cavity
1,
at
rest
in
Σ,
and
cavity
2,
at
rest
in
S,
and
denote
by
πµ
(
Eπ c
,
π
)
the
light
4-momentum
for
Σ
in
his
cavity
1
and
by
(
Ep c
,
p)
the
corresponding
light
4-momentum for S in his cavity 2. Let us also denote by gµν the space-time metric
that S uses in the relation gµνpµpν = 0 and by
γµν = diag(N 2, 1, 1, 1)
(11)
the metric used by Σ in the relation γµνπµπν = 0 and which gives an isotropic
velocity
cγ
=
Eπ /|π|
=
c N
.
Now, Special Relativity was formulated, more than a century ago, by assuming
the perfect equivalence of two reference systems in uniform translational motion.
Instead, with a preferred frame Σ, as far as light propagation is concerned, this
physical equivalence is only assumed in the ideal N = 1 limit. As anticipated, for
N ̸= 1, where light gets absorbed and then re-emitted, the fraction of refracted light
could keep track of the particular motion of matter with respect to Σ and produce,
in the frame S where matter is at rest, an angular dependence of the velocity of
light. Equivalently, assuming that the solid parts of cavity 2 are at rest in a frame
S, which is in uniform translational motion with respect to Σ, no longer implies
that the medium which stays inside, e.g. the gas, is in a state of thermodynamic
jAs anticipated, for Lorentz, only a conspiracy of effects prevents to detect the motion with respect to the ether, which, however different might be from ordinary matter, is nevertheless endowed with a certain degree of substantiality. For this reason, in his view, “it seems natural not to assume at starting that it can never make any difference whether a body moves through the ether or not”.45
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 13
equilibriumk. Thus, one should keep an open mind and exploit the implications of the basic condition
ν (N = 1) = γµν (N = 1) = ηµν
(12)
where ηµν is the Minkowski tensor. This standard equality amounts to introduce a transformation matrix, say Aµν , such that
ν = Aµρ Aνσηρσ = ηµν
(13)
This relation is strictly valid for N = 1. However, by continuity, one is driven to
conclude that an analogous relation between gµν and γµν should also hold in the
ϵ → 0 limit. The only subtlety is that relation (13) does not fix uniquely Aµν . In fact, one can either choose the identity matrix, i.e. Aµν = δνµ, or a Lorentz transformation, i.e. Aµν = Λµν . Since for any finite v these two matrices cannot be related by an infinitesimal transformation, it follows that Aµν is a two-valued function in the ϵ → 0 limit. Therefore, in principle, there are two solutions. If Aµν is the identity matrix,
we find a first solution
[gµν (N )]1 = γµν ηµν + 2ϵδ0µδ0ν
(14)
while, if Aµν is a Lorentz transformation, we find the other solution
[gµν (N )]2 = Λµρ Λνσγρσ ηµν + 2ϵvµvν
(15)
vµ being the dimensionless S 4-velocity, vµ ≡ (v0, v/c) with vµvµ = 1. Notice that with the former choice, implicitly adopted in Special Relativity
to preserve isotropy in all reference systems also for N ̸= 1, one is introducing
a discontinuity in the transformation matrix for any ϵ ̸= 0. Indeed, the whole emphasis on Lorentz transformations depends on enforcing Eq.(13) for Aµν = Λµν so that ΛµσΛνσ = ηµν and the Minkowski metric applies to all equivalent frames.
On the other hand, with the latter solution, by replacing in the relation pµpνν = 0, the photon energy now depends on the direction of propagation. Then, by defining the light velocity cγ(θ) from the ratio Ep/|p|, one finds7, 24
c cγ(θ) N
1 2ϵβ cos θ ϵβ2(1 + cos2 θ)
(16)
and a two-way velocity
γ (θ)
=
2cγ(θ)cγ(π + θ) cγ(θ) + cγ(π + θ)
c N
1 ϵβ2 1 + cos2 θ
c ≡ N¯ (θ)
(17)
kThink for instance of the collective interaction of a gaseous medium with the CMB radiation or with hypothetical dark matter in the Galaxy. However weak this interaction may be, it would mimic a non-local thermal gradient that could bring the gas out of equilibrium. The advantage of the following analysis is that it only uses symmetry properties without requiring a knowledge of the underlying dynamical processes.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
14 Authors Names
here θ is the angle between v and p (as defined in the S frame) and
N¯ (θ) N [1 + ϵβ2(1 + cos2 θ)]
(18)
Eq.(17) corresponds to setting in Eq.(10) ζ0 = 4/3, ζ2 = 2/3 and all ζ2n = 0 for n > 1 and can be considered a modern version of Maxwells original calculation.48
It represents a definite, alternative model for the interpretation of experiments per-
formed close to the ideal vacuum limit ϵ = 0, such as in gaseous systems, and will be adopted in the following l.
3. A first look at the classical experiments
From Eq.(17) we find a fractional anisotropy
∆c¯θ = c¯γ(π/2 + θ) γ(θ) ϵ (v2/c2) cos 2θ
(19)
c
c
which produces a fringe pattern
∆λ(θ) λ
2D =
λ
∆c¯θ c
D
λ
v2 c2 cos 2θ
(20)
The dragging of light in the Earth frame is then described as a pure 2nd-harmonic effect, which is periodic in the range [0, π], as in the classical theory (see e.g.47). However, its amplitude
D v2
A2 = λ 2ϵ c2
(21)
is now much smaller being suppressed by the factor 2ϵ relatively to the classical value. This was traditionally reported for the orbital velocity of 30 km/s as
Ac2lass
=
D 30 (
λ
km/s )2 c
(22)
This difference could then be re-absorbed into an observable velocity which is related
to the kinematical velocity v through the gas refractive index
vo2bs 2ϵv2
(23)
so that
D A2 = λ
vo2bs c2
(24)
lA conceptual detail concerns the relation of the gas refractive index N , as defined in the Σframe
through Eq.(11), to the experimental quantity Nexp which is extracted from measurements of the
two-way velocity in the Earth laboratory. By assuming a θdependent refractive index as in Eq.(18)
one should thus define Nexp by an angular average, i.e. Nexp ≡ ⟨N¯ (θ)⟩θ = N
1+
3 2
(N
1)β2
.
One can then determine the unknown value N ≡ N (Σ) (as if the container of the gas were at rest
in Σ), in terms of the experimentally known quantity Nexp ≡ N (Earth) and of v. As discussed in refs.7-,10 for v 370 km/s, the resulting difference |Nexp N | ≲ 109 is well below the
experimental accuracy on Nexp and, for all practical purposes, can be ignored.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 15
Fig. 4. The observable velocity reported by Miller 35 for various experiments.
Thus, this vobs is the small velocity traditionally extracted from the measured amplitude AE2 XP in the classical analysis of the experiments
vobs 30 km/s
AE2 XP Ac2lass
(25)
see e.g. Fig.4.
We can now understand the pattern observed in the classical experiments. For
instance,
in the
original Michelson-Morley experiment, where
D λ
2 · 107,
the
classically expected amplitude was Ac2lass 0.2. But the experimental amplitude
measured in the six sessions was AE2 XP = 0.01 ÷ 0.02. This corresponds to an
average
anisotropy
|∆c¯θ |exp c
4 · 1010
and
was
originally
interpreted
in
terms
of
a velocity vobs 8 km/s. However, for an experiment in air at room temperature
and atmospheric pressure where ϵ 2.8 · 104, this observable velocity would now
correspond to a true kinematic value v 340 km/s which would fit well with the
cosmic motion indicated by the CMB dipole anistropy. Therefore, the importance
of the issue requires to sharpen the analysis of the old experiments, starting from
the early 1887 trials.
3.1. The 1887 Michelson-Morley experiment in Cleveland
The precision of the Michelson-Morley apparatus33 was extraordinary, about ±0.004 of a fringe. For all details, we address the reader to our book.9 Here, we just limit ourselves to quote from Born.49 When discussing the classically expected fringe shift upon rotation of the apparatus by 90 degrees, namely 2Ac2lass 0.4, he says explicitly: “Michelson was certain that the one-hundredth part of this displacement would still be observable” (i.e. 0.004). As a check, the Michelson-Morley fringe shifts were recomputed in refs.7, 9, 50 from the original article,33 see Table 1. These data
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
16 Authors Names
were then analyzed in a Fourier expansion, see e.g. Fig.5 (note that a 1st-harmonic has to be present in the data due to the arrangement of the mirrors needed to have fringes of finite width, see35, 51). One can thus extract amplitude and phase of the 2nd-harmonic component by fitting the even combination of fringe shifts
∆λ(θ) + ∆λ(π + θ)
B(θ) =
(26)
see Fig.6.
Table 1.
The
fringe
shifts
∆λ(i) λ
for
all
noon
(n.)
and
evening
(e.)
sessions
of
the
Michel-
son-Morley
experiment.
The
angle
of
rotation
is
defined
as
θ
=
i1 16
2π.
The
Table
is
taken
from ref.7
i
July 8 (n.) July 9 (n.) July 11 (n.) July 8 (e.) July 9 (e.) July 12 (e.)
1
0.001
+0.018
+0.016
0.016
+0.007
+0.036
2
+0.024
0.004
0.034
+0.008
0.015
+0.044
3 +0.053
0.004
0.038
0.010
+0.006
+0.047
4
+0.015
0.003
0.066
+0.070
+0.004
+0.027
5
0.036
0.031
0.042
+0.041
+0.027
0.002
6
0.007
0.020
0.014
+0.055
+0.015
0.012
7
+0.024
0.025
+0.000
+0.057
0.022
+0.007
8
+0.026
0.021
+0.028
+0.029
0.036
0.011
9
0.021
0.049
+0.002
0.005
0.033
0.028
10 0.022
0.032
0.010
+0.023
+0.001
0.064
11 0.031
+0.001
0.004
+0.005
0.008
0.091
12 0.005
+0.012
+0.012
0.030
0.014
0.057
13 0.024
+0.041
+0.048
0.034
0.007
0.038
14 0.017
+0.042
+0.054
0.052
+0.015
+0.040
15 0.002
+0.070
+0.038
0.084
+0.026
+0.059
16 +0.022
0.005
+0.006
0.062
+0.024
+0.043
The 2nd-harmonic amplitudes for the six experimental sessions are reported in
Table 2. Due to their statistical consistency, one can compute the mean and variance
and obtain AE2 XP 0.016±0.006. This value is consistent with an observable velocity
vobs
8.5
+1.7 2.2
km/s
(27)
in complete agreement with Miller, ses Fig.4. In this sense, our re-analysis supports the claims of Hicks and Miller. The fringe shifts were much smaller than expected but in two experimental sessions (11 July noon and 12 July evening), the secondharmonic amplitude is non-zero at the 5σ level and in other two sessions (July 9 noon and July 8 evening) is non-zero at the 3σ level. As such, the average measured amplitude AE2 XP 0.016, although much smaller than the classical expectation Ac2lass 0.2, was not completely negligible. Thus it is natural to ask: should this
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 17
Fig. 5. The fringe shifts for the session of July 9 evening. The fit is performed by including terms up to fourth harmonics. The figure is taken from ref.50
Fig. 6. A fit to the even combination B(θ) Eq.(26). The second harmonic amplitude is AE2 XP = 0.025 ± 0.005 and the fourth harmonic is AE4 XP = 0.004 ± 0.005. The figure is taken from ref.50
value be interpreted as a typical instrumental artifact (a “null result”) or could also indicate a genuine ether-drift effect? Of course, this question is not new and, in the past, greatest experts have raised objections to the standard null interpretation of the data. This point of view was well summarized by Miller in 193335 as follows: “The brief series of observations (by Michelson and Morley) was sufficient to show clearly that the effect did not have the anticipated magnitude. However, and this fact must be emphasized, the indicated effect was not zero”. The same conclusion had already been obtained by Hicks in 1902:51 “The data published by Michelson and Morley, instead of giving a null result, show distinct evidence for an effect of the kind to be expected”. There was a 2nd-harmonic effect whose amplitude, however, was substantially smaller than the classical expectation (see Fig.7).
Thus the real point about the Michelson-Morley data does not concern the small
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
18 Authors Names
Table 2. The 2nd-harmonic amplitudes for the six experimental sessions of the Michelson-Morley experiment. The table is taken from ref.7
SESSION July 8 (noon) July 9 (noon) July 11 (noon) July 8 (evening) July 9 (evening) July 12 (evening)
AE2 XP 0.010 ± 0.005 0.015 ± 0.005 0.025 ± 0.005 0.014 ± 0.005 0.011 ± 0.005 0.024 ± 0.005
Fig. 7. The even combination of fringe shifts B(θ) for the Michelson-Morley data as reported by Hicks.51 Solid and dashed lines refer respectively to noon and evening observations. Compare the solid curve of July 11th with the analogous curve in Fig.6.
magnitude of the amplitude but the sizeable changes in the azimuth, i.e in the phase θ2 of the 2nd-harmonic which gives the direction of the drift in the plane of the interferometer. By performing observations at the same hour on consecutive days (so that variations in the orbital motion are negligible) one expects that this angle should remain the same within the statistical errors. Now, by taking into
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 19
account that, in a 2nd-harmonic effect, the angle is always defined up to ±180o, one choice for the experimental θ2values is 357o ± 14o, 285o ± 10o and 317o ± 8o respectively for the noon sessions of July 8th, 9th and 11th. For this assignment, the individual velocity vectors vobs(cos θ2, sin θ2) and their mean are shown in Fig. 8. As a consequence, directly averaging the amplitudes of the individual sessions is considerably different from first performing the vector average of the data and then computing the resulting amplitude. In the latter case, the average amplitude is reduced from 0.016 to about 0.011, with a central value of the observable velocity which decreases from 8.5 km/s to 7 km/s.
This irregular character of the observations has always represented a strong argument to interpret the small observed residuals as typical instrumental effects.
Fig. 8. The observable velocities for the three noon sessions of the Michelson-Morley experiment and their mean. The x-axis corresponds to θ2 = 0o ≡ 360o and the y-axis to θ2 = 270o. Statistical uncertainties of the various determinations are not reported. All individual directions could also be reversed by 180o. The figure is taken from.7
3.2. Further insights: Miller vs. Piccard-Stahel To get further insights we have compared two other sets of measurements, namely Millers observations35 and those performed by Piccard and Stahel.39 Millers large interferometer had an optical path of about 32 metres and was installed on top of Mount Wilson. His most extensive observations were made in blocks of ten days around April 1, August 1, September 15, 1925, and later on around February 8, 1926, with a total number of 6402 turns of the interferometer.35 The result of his 1925 measurements, presented at the APS meeting in Kansas City on December 1925, was confirming his original claim of 1921, namely “there is a positive, systematic ether-drift effect, corresponding to a relative motion of the Earth and the ether, which at Mt. Wilson has an apparent velocity of 10 km/s”.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
20 Authors Names
Being aware that Millers previous 1921 announcement of a non-zero ether-drift of about 9 km/s, if taken seriously, could undermine the basis of Einsteins relativity (Millers results “carried a mortal blow to the theory of relativity”), Piccard and Stahel realized a precise apparatus with a small optical path of 280 cm that could be carried on board of a free atmospheric balloon (at heights of 2500 and 4500 m) to check the dependence on altitude. In this first series of measurements thermal disturbances were so strong that they could only set an upper limit of about 9 km/s to the magnitude of any ether-drift. However, after this first series of trials, precise observations were performed on dry land in Brussels and on top of Mt.Rigi in Switzerland (at an height of 1800 m).
Despite the optical path was much shorter than the size of the instruments used at that times in the United States, Piccard and Stahel were convinced that the precision of their measurements was higher because spurious disturbances were less important. In particular, with respect to the traditional direct observation, the fringe shifts were registered by photographic recording. Also, for thermal insulation, the interferometer was surrounded either by a thermostat filled with ice or by an iron enclosure where it could be possible to create a vacuum. This last solution was considered after having understood that the main instability in the fringe system was due to thermal disturbances in the air of the optical arms (rather than to temperature differences in the solid parts of the apparatus). However, very often the interference fringes were put out of order after few minutes by the presence of residual bubbles of air in the vacuum chamber. For this reason, they finally decided to run the experiment at atmospheric pressure with the ice thermostat which, by its great heat capacity, was found to stabilize the temperature in a satisfactory way.
We have thus considered the compatibility of these two experiments. Miller was always reporting his observations by quoting separately the amplitude and the phase of the individual sessions. In this way, as shown in Fig.4, the average observable velocity, obtained from a classical interpretation of his data, was vobs 8.4 ± 2.2 km/s. Piccard and Stahel were instead first performing a vector average of the data and, since the phase was found to vary in a completely arbitrary way, were quoting the much smaller value vobs (1.5 ÷ 1.7) km/s. For this reason, their measurements are traditionally considered a definite refutation of Miller.
But suppose that the ether-drift phenomenon has an intrinsic non-deterministic nature, which induces random fluctuations in the direction of the local drift. In this case, a vector average of the data from various sessions would completely obscure the physical information contained in the individual observations. For this reason, a meaningful comparison with Miller requires to apply his same procedure to the Piccard-Stahel data. Namely, first summarizing each measurement into a definite pair (AE2 XP, θ2EXP) for amplitude and azimuth, and then computing the magnitude of the observable velocity from the measured amplitudes. With this different procedure, the Piccard and Stahel observable velocity, at the 75% CL, becomes now
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 21
Fig. 9. We report in panel (a) the probability histogram W for the observable velocity obtained from the 24 individual amplitudes reported by Piccard and Stahel in.39 In panel (c) we report the analogous histogram obtained from Millers Figure 22d in.35 In both cases, the vertical normalization is to a unit area. Finally, in panel (b) we report the overlap of the two histograms. The area of the overlap is 0.645. This gives a consistency between the two experiments of about 64%. The figure is taken from.9
much larger, namely
vobs = 6.3+12..50 km/s
(Piccard Stahel)
(28)
and is now compatible with Millers results. For a more refined test, we constructed probability histograms by considering the large set of measurements reported by Miller in Figure 22d of35 and the 24 individual amplitudes reported by Piccard and Stahel in,39 see Fig.9. From the area of the overlap, the consistency of the two experiments can be estimated to be about 64% which is a quite high level. At the same time, since the agreement is restricted to the region vobs < 9 km/s, Millers higher values are likely affected by systematic disturbances. This would confirm Piccard and Stahels claim that their apparatus, although of a smaller size, was more precise.
Therefore, summarizing, there is a range of observable velocity, say vobs 6.0 ± 2.0 km/s, where the results of the three experiments we have considered, namely Michelson-Morley, Miller and Piccard and Stahel, overlap consistently. This common range is obtained from the 2nd-harmonic amplitudes measured in a plenty of experimental sessions performed at different sidereal times and in different laboratories. As such, to a large extent, it should also be independent of spurious systematic effects. On the basis of Eq.(23), this range corresponds to a true kinematic velocity v 250 ± 80 km/s which could reasonably fit with the projection of the Earth velocity within the CMB at the various laboratories. Truly enough, this is only a first, partial view which must be supplemented by a deeper understanding of the observed random variability of the phase.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
22 Authors Names
Fig. 10. The propagation of light in an optical cavity. It is emphasized that, independently of its particular name (physical vacuum, ether...) and differently from the solid parts of the apparatus, the underlying substrate is not completely entrained with the Earth motion. Thus, in general, its state of motion vµ(t) is different from v˜µ(t).
4. Going deeper into the ether-drift phenomenon The traditional expectation that an ether drift should precisely follow the smooth modulations induced by the Earth rotation, derives from the identification of the local velocity field which describes the drift in the plane of the interferometer, say vµ(t), with the corresponding projection of the global Earth motion, say v˜µ(t). By comparing with the motion of a macroscopic body in a fluid, this identification is equivalent to assume a form of regular, laminar flow, where global and local velocity fields coincide. Depending on the nature of the fluid, this assumption may be incorrect.
To formulate an alternative model of ether drift, in refs.,710 we started from Maxwells original view48 of light as a wave process which takes place in some substrate: “We are therefore obliged to suppose that the medium through which light is propagated is something distinct from the transparent media known to us”. He was calling the underlying substrate ether while, today, we prefer to call it physical vacuum. However, this is irrelevant. The essential point for the propagation of light, e.g. inside an optical cavity, is that, differently from the solid parts of the apparatus, this physical vacuum is not completely entrained with the Earth motion see Fig.10.
Thus, to explain the irregular character of the data, one could try to model the state of motion of the vacuum substrate as in a turbulent fluid52, 53 or, more precisely, as in a fluid in the limit of zero viscosity. Then, the simple picture of a laminar flow is no more obvious due to the subtlety of the infinite-Reynolds-number limit, see e.g. Sect. 41.5 in Vol.II of Feynmans lectures.54 In fact, beside the laminar regime where vµ(t) = v˜µ(t), there is also another solution where vµ(t) becomes a continuous but nowhere differentiable velocity field55, 56 m.
mThe idea of the physical vacuum as an underlying stochastic medium, similar to a turbulent fluid, is deeply rooted in basic foundational aspects of both quantum physics and relativity. For
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 23
Fig. 11. We compare two different signals. Panel a) reports the turbulent velocity field measured in a wind tunnel.67 Panel b) reports the instantaneous frequency shift measured with vacuum optical resonators in ref.68 For the adopted laser frequency ν0 = 2.8 · 1014 Hz a ∆ν = ±1 Hz corresponds to a fractional value ∆ν/ν0 of about ±3.5 · 1015.
Together with these theoretical arguments, the analogy with a turbulent flow finds support in modern ether drift experiments where one measures the frequency shifts of two optical resonators. To this end, consider Fig.11. Panel a) reports the turbulent velocity field measured in a wind tunnel.67 No doubt, this is a genuine signal, not noise. Panel b) reports instead the instantaneous frequency shift measured with vacuum optical cavities in ref.68 So far, this other signal is interpreted as spurious noise.
Consider now Fig.12. Panel a) shows the power spectrum S(ω) ω1.5 of the
instance, at the end of XIX century, the last model of the ether was a fluid full of very small whirlpools (a vortex-sponge).57 The hydrodynamics of this medium was accounting for Maxwell equations and thus providing a model of Lorentz symmetry as emerging from a system whose elementary constituents are governed by Newtonian dynamics. In a different perspective, the idea of a quantum ether, as a medium subject to the fluctuations of the uncertainty relations, was considered by Dirac.58 More recently, the model of turbulent ether has been re-formulated by Troshkin59 (see also60 and61) within the Navier-Stokes equation, by Saul62 within Boltzmanns transport equation and in63 within Landaus hydrodynamics. The same picture of the vacuum (or ether) as a turbulent fluid was Nelsons64 starting point. In particular, the zero-viscosity limit gave him the motivation to expect that “the Brownian motion in the ether will not be smooth” and, therefore, to conceive the particular form of kinematics at the base of his stochastic derivation of the Schr¨odinger equation. A qualitatively similar picture is also obtained by representing relativistic particle propagation from the superposition, at short time scales, of non-relativistic particle paths with different Newtonian mass.65 In this formulation, particles randomly propagate (as in a Brownian motion) in an underlying granular medium which replaces the trivial empty vacuum.66
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
24 Authors Names
Fig. 12. We compare two different signals. Panel a) reports the power spectrum S(ω) ω1.5 of the wind turbulence measured at the Florence Airport.69 Panel b) shows the spectral amplitude
S(ω) ω0.7 of the frequency shift measured by Nagel et al.70 Above some minimal frequency the two curves reach a flat plateau. This corresponds to the maximum integration time beyond which the signal ceases to behave as a pure white-noise.
wind turbulence measured at the Florence Airport.69 No doubt, this is a physical signal. Panel b) shows the spectral amplitude S(ω) ω0.7 of the frequency shift measured by Nagel et al.70 Again, this latter signal is interpreted as spurious noise.
Clearly these are just analogies but, very often, physical understanding proceeds
by analogies. We have thus exploited the idea that the irregular signal observed in
ether-drift experiments has a fundamental stochastic nature as when turbulence, at
small scales, becomes statistically homogeneous and isotropic. With such an irregu-
lar signal numerical simulations are needed for a consistent description of the data.
Therefore, for a check, one should first extract from the data the (2nd-harmonic)
phase and amplitude and concentrate on the latter which is positive definite and re-
mains non-zero under any averaging procedure. When measured at different times,
this amplitude will anyhow exhibit modulations that, though indirectly, can provide
information on the underlying cosmic motion.
To put things on a quantitative basis, let us assume the set of kinematic param-
eters (V, α, γ)CMB for the Earth motion in the CMB, a latitude ϕ of the laboratory
and
a
given
sidereal
time
τ
=
ωsidt
of
the
observations
(with
ωsid
2π 23h 56
).
Then,
for short-time observations of a few days, where the only time dependence is due
to the Earth rotation, a simple application of spherical trigonometry71 gives the
projections in the (x, y) plane of the interferometer
v˜x(t) = v˜(t) cos θ˜2(t) = V [sin γ cos ϕ cos γ sin ϕ cos(τ α)]
(29)
v˜y(t) = v˜(t) sin θ˜2(t) = V cos γ sin(τ α)
(30)
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 25
with a magnitude
v˜(t) = V | sin z(t)|
(31)
and
cos z(t) = sin γ sin ϕ + cos γ cos ϕ cos(τ α)
(32)
As for the signal, let us also re-write Eq.(19) as
∆c¯θ (t) c
v2(t) ϵ c2
cos 2(θ
θ2(t))
(33)
where v(t) and θ2(t) now indicate respectively the magnitude and direction of the local drift in the same (x, y) plane of the interferometer. This can also be re-written
as
∆c¯θ(t) 2S(t) sin 2θ + 2C(t) cos 2θ
(34)
c
with
2C (t)
=
ϵ
vx2(t) vy2(t) c2
2S(t)
=
ϵ
2vx (t)vy (t) c2
(35)
and vx(t) = v(t) cos θ2(t), vy(t) = v(t) sin θ2(t). In an analogy with a turbulent flow, the requirement of statistical isotropy means
that the local quantities vx(t) and vy(t), which determine the observable properties
of the drift, are very irregular functions that differ non trivially from their smooth,
global counterparts v˜x(t) and v˜y(t), and can only be simulated numerically. To this end, a representation in terms of random Fourier series55, 72, 73 was adopted in
refs.710 in a simplest uniform-probability model, where the kinematic parameters
of the global v˜µ(t) are just used to fix the boundaries for the local random vµ(t). The basic ingredients are summarized in the Appendix.
In this model, the functions S(t) and C(t) have the characteristic behaviour of a
white-noise signal with vanishing statistical averages ⟨C(t)⟩stat = 0 and ⟨S(t)⟩stat =
0 at any time t and whatever the global cosmic motion of the Earth. One can then
understand the observed irregular behaviour of the fringe shifts
∆λ(θ; t) 2D
= [2S(t) sin 2θ + 2C(t) cos 2θ]
(36)
λ
λ
In fact, their averages would be non vanishing just because the statistics is finite.
Otherwise with more and more observations one would find
∆λ(θ; t)
2D
⟨ λ ⟩stat = λ [2 sin 2θ ⟨S(t)⟩stat + 2 cos 2θ ⟨C(t)⟩stat] → 0
(37)
In particular, the direction θ2(t) of the local drift, defined by the relation
tan 2θ2(t) = S(t)/C(t), would vary randomly with no definite limit.
We have then checked the model by comparing with the amplitudes. Here we have first to consider the theoretical amplitude A˜2(t) associated with the global
motion
A˜2(t)
D λ
·
·
V
2
sin2 c2
z(t)
(38)
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
26 Authors Names
and then the amplitude A2(t) associated with the local non differentiable velocity components vx(t) and vy(t), Eqs.(83) and (84) of the Appendix
A2(t)
D λ
·
·
vx2(t) + c2
vy2(t)
(39)
Clearly, the latter will exhibit sizeable fluctuations and be very different from the smooth A˜2(t). However, as shown in the Appendix, the relation between A˜2(t) and
the statistical average ⟨A2(t)⟩stat is extremely simple
D ⟨A2(t)⟩stat = λ
· 2ϵ ·
⟨vx2(t) + vy2(t)⟩stat c2
π2 18
· A˜2(t)
(40)
so that, by averaging the amplitude at different sidereal times, one can obtain the crucial information on the angular parameters α and γ.
Altogether, the amplitudes of those old measurements can thus be interpreted in terms of three different velocities: a) as 6 ± 2 km/s in a classical picture b) as 250 ± 80 km/s, in a modern scheme and in a smooth picture of the drift c) as 340 ± 110 km/s, in a modern scheme but now allowing for irregular fluctuations of the signal. In fact, by replacing Eq.(38) with Eq.(40), from the same data, one would now obtain kinematical velocities which are larger by a factor 18/π2 1.35. In this third interpretation, the range of velocity agrees much better with the CMB value of 370 km/s.
To illustrate the agreement of our scheme with all classical measurements, we address to our book9 where a detailed description is given of the experiments by Morley-Miller,34 Miller,35 Kennedy,36 Illingworth,37 Tomaschek38 and PiccardStahel.39 Instead, here, we will only consider the two most precise experiments that, traditionally, have been considered as definitely ruling out Millers claims for a non-zero ether drift. Namely the Michelson-Pease-Pearson (MPP) observations at Mt. Wilson and the experiment performed in 1930 by Joos in Jena.43 In particular, the latter remains incomparable among the classical experiments. To have an idea, Sommerfeld, being aware that the residuals in the Michelson-Morley data were not entirely negligible, concluded that only “After its repetition at Jena by Joos, the negative result of Michelsons experiment can be considered as definitely established” (A. Sommerfeld, Optics). However, there is again a subtlety because, as we shall see, Joos experiment was not performed in the same conditions as the other experiments we have previously considered.
4.1. Reanalysis of the MPP experiment
To re-analyze the Michelson-Pease-Pearson (MPP) experiment, we first observe that no numerical results are reported in the original articles.40, 41 Instead, for more precise indications, one should look at Peases paper.42 There, one learns that they concentrated on a purely differential type of measurements. Namely, they were first statistically averaging the fringe shifts at those sidereal times that, according to
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 27
Fig. 13. The histogram W of a numerical simulation of 10,000 instantaneous amplitudes for the single session of January 13, 1928, reported by Pease42 . The vertical normalization is to a unit area. We show the median and the 70% CL. The limits on the random Fourier components Eqs.(83) and (84) of the Appendix were fixed by inserting the CMB kinematical parameters in Eq.(87).
Miller, were corresponding to maxima and minima of the ether-drift effect. Then, they were forming the difference
δ(θ)
=
∆λ(θ; tmax) λ
⟩stat
∆λ(θ; tmin) λ
⟩stat
(41)
which are the only numbers reported by Pease. These δvalues have a maximal
magnitude of ±0.004 and this is also the order of magnitude of the experimen-
tal amplitude AE2 XP 0.005 that is usually reported74 for the MPP experiment when comparing with the much larger expected classical amplitudes Ac2lass 0.45 or Ac2lass 0.29 for optical paths of eighty-five or fifty-five feet respectively. Now,
our stochastic, isotropic model predicts exactly zero statistical averages for vector
quantities such as the fringe shifts, see Eq(37). Therefore, it would be trivial to
reproduce the small δ-values in Eq.(41) in a numerical simulation with sufficiently
high statistics. We have thus decided to compare instead with the only basic experimental session reported by Pease42 (for optical path of fifty-five feet) which indicates a 2nd-harmonic amplitude AE2 XP 0.006. By comparing with the classical prediction for 30 km/s, namely Ac2lass 0.29, this amplitude corresponds to an observable velocity vobs 4.3 km/s but to a much larger value on the basis of
Eq.(39).
Since we are dealing with a single measurement, to obtain a better understand-
ing of its probability content, we have performed a direct numerical simulation by
generating 10,000 values of the amplitude at the same sidereal time 5:30 of the MPP
Mt. Wilson observation. The CMB kinematical parameters were used to bound the
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
28 Authors Names
random Fourier components of the stochastic velocity field Eqs.(83) and (84) of the
Appendix. The resulting histogram is reported in Fig.13. From this histogram one obtains a mean simulated amplitude ⟨As2imul⟩ 0.014. This corresponds to replace the value of the scalar velocity v˜(t) 370 km/s Eq.(77), at the sidereal time of
Peases observation, in the relation for the statistical average of the amplitude
2ϵD ⟨A2(t)⟩stat = λ
⟨vx2(t) + vy2(t)⟩stat c2
(1.6 · 104) ·
π2 18
·
v˜2(t) c2
v˜2(t) 0.009 · (300 km/s)2
(42)
In the above relation we have replaced D/λ 2.9 · 107 (for optical path of fifty-five
feet) and ϵ 2.8 · 104.
Note that the median of the amplitude distribution is about 0.007. As a conse-
quence, the value AE2 XP 0.006 lies well within the 70% Confidence Limit. Also, the probability content becomes large at very small amplitudes n and there is a long
tail extending up to about A2 0.030. The wide interval of amplitudes corresponding to the 70% C. L. indicates that,
in our stochastic model, one could account for single observations that differ by an
order of magnitude, say from 0.003 to 0.030. Thus, beside the statistical vanishing of
vector quantities, this is another crucial difference with a purely deterministic model
of the ether-drift. In this traditional view, in fact, within the errors, the amplitude
can vary at most by a factor r = (v˜max/v˜min)2 where v˜max and v˜min are respectively the maximum and minimum of the projection of the Earth velocity Eq.(31). Since,
for the known types of cosmic motion, one finds r 2, the observation of such large
fluctuations in the data would induce to conclude, in a deterministic model, that
there is some systematic effect which affects the measurements in an uncontrolled
way. With an ether drift of such irregular nature, it then becomes understandable
the MPP reluctance to quote the individual results and instead report those partic-
ularly small combinations in Eq.(41) obtained by averaging and further subtracting
large samples of data. This general picture of a highly irregular phenomenon is also
confirmed by our reanalysis of Joos experiment in the following subsection.
4.2. Joos experiment We will only give a brief description of Joos 1930 experiment43 and address to our book9 for more details. Its sensitivity was about 1/3000 of a fringe, the fringes were recorded photographically with an automatic procedure and the optical system was enclosed in a hermetic housing. As reported by Miller,35, 75 it has been traditionally
nStrictly speaking, for a more precise description of the data, one should fold the histogram with a smearing function which takes into account the finite resolution ∆ of the apparatus. This smearing would force the curve to bend for A2 → 0 and tend to some limit which depends on ∆. Nevertheless, this refinement should not modify substantially the probability content around the median which is very close to A2 = 0.007.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 29
Fig. 14. The fringe shifts reported by Joos.43 The yardstick corresponds to 1/1000 of a wavelength.
believed that the measurements were performed in vacuum. In his article, however, Joos is not clear on this particular aspect. Only when describing his device for electromagnetic fine movements of the mirrors, he refers to the condition of an evacuated apparatus.43 Instead, Swenson76, 77 declares that Joos fringe shifts were finally recorded with optical paths placed in a helium bath. Therefore, we have decided to follow Swensons explicit statements and assumed the presence of gaseous helium at atmospheric pressure.
From Eq.(40), by replacing D/λ = 3.75 · 107 and the refractive index Nhelium 1.000033 for gaseous helium, an average daily projection of the cosmic Earth velocity v˜(t) = V | sin z(t)| 330 km/s (appropriate for a Central-Europe laboratory) would provide the same amplitude as classically expected for the much smaller observable velocity of 2 km/s. We can thus understand the substantial reduction of the fringe shifts observed by Joos, with respect to the other experiments in air.
The data were taken at steps of one hour during the sidereal day and two observations (1 and 5) were finally deleted by Joos with the motivations that there were spurious disturbances, see Fig.14. From this picture, Joos adopted 1/1000 of a wavelength as upper limit and deduced the bound vobs ≲ 1.5 km/s. To this end, he was comparing with the classical expectation that, for his apparatus, a velocity of 30 km/s should have produced a 2nd-harmonic amplitude of 0.375 wavelengths. Though, since it is apparent that some fringe displacements were certainly larger than 1/1000 of a wavelength, we have performed 2nd-harmonic fits to Joos data, see Fig.15. The resulting amplitudes are reported in Fig.16.
We note that a 2nd-harmonic fit to the large fringe shifts in picture 11 has a
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
30 Authors Names
Fig. 15. Some 2nd-harmonic fits to Joos data. The figure is taken from ref.9
Fig. 16. Joos 2nd-harmonic amplitudes, in units 103. The vertical band between the two lines corresponds to the range (1.4 ± 0.8) · 103. The uncertainty of each value is about ±3 · 104. The figure is taken from ref.7
very good chi-square, comparable and often better than other observations with smaller values, see Fig.15. Therefore, there is no reason to delete the observation n.11. Its amplitude, however, (4.1 ± 0.3) · 103 is abot ten times larger than the average amplitude (0.4 ± 0.3) · 103 from the observations 20 and 21. This difference cannot be understood in a smooth model of the drift where, as anticipated, the projected velocity squared at the observation site can at most differ by a factor of two, as for the CMB motion at typical Central-Europe latitude where (v˜)min 250 km/s and (v˜)max 370 km/s. To understand these characteristic fluctuations, we have thus performed various numerical simulations of these amplitudes7, 9 in the
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 31
stochastic model described in the Appendix and using the kinematical parameters (V, α, γ)CMB to place the limits on the random velocity component Eqs.(83) and (84). Two simulations are shown in Figs.17 and 18 (the corresponding numerical values are reported in7, 9).
Fig. 17. Joos 2nd-harmonic amplitudes, in units 103 (black dots), are compared with a single simulation (red diamonds) at the same sidereal times of Joos observations. Two 5th-order polynomial fits to the two sets of values are also shown. The figure is taken from ref.7 .
We want to emphasize two aspects. First, Joos average amplitude ⟨AE2 XP⟩ =
(1.4 ± 0.8) · 103 when compared with the classical prediction for his interferometer
Ac2lass
=
D (30km/s)2
λ
c2
0.375 gives indeed an observable velocity vobs
(1.8 ±
0.5) km/s very close to the 1.5 km/s value quoted by Joos. But, when comparing
with our prediction in the stochastic model Eq.(40) one would now find a true
kinematical velocity v˜ = 305+81500 km/s. Second, when fitting with Eqs.(76) and (77) the smooth black curve of the Joos data in Fig.17 one finds7 a right ascension
α(fit Joos) = (168±30) degrees and an angular declination γ(fit Joos) = (13±
14) degrees which are consistent with the present values α(CMB) 168 degrees
and γ(CMB) 7 degrees. This confirms that, when studied at different sidereal
times, the measured amplitude can provide precious information on the angular
parameters.
4.3. Summary of all classical experiments A comparison with all classical experiments is finally shown in Table 3.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
32 Authors Names
Fig. 18. Joos 2nd-harmonic amplitudes in units 103 (black dots) are now compared with a simulation where one averages ten measurements, performed on 10 consecutive days, at the same sidereal times of Joos observations (red diamonds). The change of the averages observed by varying the parameters of the simulation was summarized into a central value and a symmetric error. The figure is taken from ref.7
Note the substantial difference with the analogous summary Table I of ref.74 where those authors were comparing with the classical amplitudes Eq.(22) and emphasizing the much smaller magnitude of the experimental fringes. Here, is just the opposite. In fact, our theoretical statistical averages are often smaller than the experimental results indicating, most likely, the presence of systematic effects in the measurements.
At the same time, by adopting Eq.(40), we find v˜exp 418 ± 62 km/s from all experiments in air and v˜exp 323 ± 70 km/s from the two experiments in gaseous helium, with a global average ⟨v˜exp⟩ 376±46 km/s which agrees well with the 370 km/s from the CMB observations. Even more, from the two most precise experiments of Piccard-Stahel (Brussels and Mt.Rigi) and Joos (Jena), we find two determinations, v˜exp = 360+81510 km/s and v˜exp = 305+81500 km/s respectively, whose average ⟨v˜ 332+6800 km/s reproduces to high accuracy the projection of the CMB velocity at a typical Central-Europe latitude.
4.4. The intriguing role of temperature
As anticipated in Sect.2 (see footnote k), symmetry arguments can successfully
describe a phenomenon regardless of the physical mechanisms behind it. The same
is
true
here
with
our
relation
|∆c¯θ | c
ϵ(v2/c2)
in
Eq.(19).
It
works
but
does
not
explain the ultimate origin of the small effects observed in the gaseous systems. For
instance, as a first mechanism, we considered the possibility of different polarizations
in different directions in the dielectric, depending on its state of motion. But, if this
works in weakly bound gaseous matter, the same mechanism should also work in
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 33
Table 3. The average 2nd-harmonic amplitudes of classical ether-drift experiments. These were extracted from the original papers by averaging the amplitudes of the individual observations and assuming the direction of the local drift to be completely random (i.e. no vector averaging of different sessions). These experimental values are then compared with the full statistical average Eq.(40) for a projection 250 km/s ≲ v˜ ≲ 370 km/s of the Earth motion in the CMB and refractivities ϵ = 2.8 · 104 for air and ϵ = 3.3 · 105 for gaseous helium. The experimental value for the Morley-Miller experiment is taken from the observed velocities reported in Millers Figure 4, here our Fig.4. The experimental value for the Michelson-Pease Pearson experiment refers to the only known session for which the fringe shifts are reported explicitly42 and where the optical path was still fifty-five feet. The symbol ±.... means that the experimental uncertainty cannot be determined from the available informations. The table is taken from ref.10
Experiment Michelson(1881) Michelson-Morley(1887) Morley-Miller(1902-1905) Miller(1921-1926) Tomaschek (1924) Kennedy(1926) Illingworth(1927) Piccard-Stahel(1928) Mich.-Pease-Pearson(1929)
Joos(1930)
gas air air air air air helium helium air air helium
AE2 XP (7.8 ± ....) · 103 (1.6 ± 0.6) · 102 (4.0 ± 2.0) · 102 (4.4 ± 2.2) · 102 (1.0 ± 0.6) · 102
< 0.002 (2.2 ± 1.7) · 104 (2.8 ± 1.5) · 103 (0.6 ± ...) · 102 (1.4 ± 0.8) · 103
2D λ
4 · 106 4 · 107 1.12 · 108 1.12 · 108 3 · 107 7 · 106 7 · 106 1.28 · 107 5.8 · 107 7.5 · 107
⟨A2 (t)⟩stat (0.7 ± 0.2) · 103 (0.7 ± 0.2) · 102 (2.0 ± 0.7) · 102 (2.0 ± 0.7) · 102 (0.5 ± 0.2) · 102 (1.4 ± 0.5) · 104 (1.4 ± 0.5) · 104 (2.2 ± 0.8) · 103 (1.0 ± 0.4) · 102 (1.5 ± 0.6) · 103
a strongly bound solid dielectric, where the refractivity is (Nsolid 1) = O(1), and
thus
produce
a
much
larger
|∆c¯θ | c
(Nsolid
1)(v2/c2)
106.
This
is
in
contrast
with the Shamir-Fox78 experiment in perspex where the observed value was smaller
by orders of magnitude.
We have thus re-considered8, 9, 79 the traditional thermal interpretation74, 80 of
the observed residuals. The idea was that, in a weakly bound system as a gas, a
small temperature difference ∆T gas(θ) in the air of the two optical arms produces
a difference in the refractive index and a (∆c¯θ/c) ϵgas∆T gas(θ)/T , where T 300 K is the absolute temperature of the laboratory o. Miller was aware35, 75 that
his results could be due to a ∆T gas(θ) ≲ 1 mK but objected that casual changes
of the ambiance temperature would largely cancel when averaging over many mea-
surements. Only temperature effects with a definite periodicity would survive. For
a quantitative estimate, by averaging over all experiments in Table 3 we found
⟨v˜exp⟩ 376 ± 46 km/s. Therefore, by comparing Eq.(40) with the corresponding
oThe starting point is the Lorentz-Lorenz equation for the molecular polarizability in the ideal-gas approximation (as for air or gaseous helium at atmospheric pressure), see8, 9 for the details.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
34 Authors Names
form for a thermal light anisotropy, we find
|∆c¯θ | c
ϵgas
π2 18
⟨v˜exp⟩2 c2
|∆T gas(θ)|
ϵgas
T
(43)
and a value8, 9 |∆T gas(θ)| (0.26 ± 0.07) mK p.
This motivates the following two observations. First, after a century from those
old measurements, in a typical room-temperature laboratory environment, a stability at the level of a fraction of millikelvin is still state of the art, see.8183 This
would support the idea that we are dealing with a non-local effect that places a
fundamental limit. Second, as for possible dynamical explanations, we mentioned in footnote k a
collective interaction of the gaseous system with hypothetical dark matter in the
Galaxy or with the CMB radiation. For the consistency with the velocity of 370
km/s, the latter hypothesis seems now more plausible. In this interpretation, these
interactions would be so weak that, on average, the induced temperature differences in the optical paths are only 1/10 of the whole ∆T CMB(θ) in Eq.(3).
Nevertheless, whatever its precise origin, this thermal explanation can help intuition. In fact, it can explain the quantitative reduction of the effect in the vacuum
limit where ϵgas → 0 and the qualitative difference with solid dielectric media where temperature differences of a fraction of millikelvin cannot produce any appreciable
deviation from isotropy in the rest frame of the medium.
Admittedly, the idea that small modifications of gaseous matter, produced by the
tiny CMB temperature variations, can be detected by precise optical measurements
in a laboratory, while certainly unconventional, has not the same implications of a genuine preferred-frame effect due to the vacuum structure. Still, this thermal explanation of the small residuals in gases, very recently reconsidered by Manley,84
has a crucial importance. In fact, it implies that if a tiny, but non-zero, fundamental
signal could definitely be detected in vacuum then, with very precise measurements,
the same universal signal should also show up in a solid dielectric where a disturbing
∆T of a fraction of millikelvin becomes irrelevant. Detecting such non-thermal light
anisotropy, for the same cosmic motion indicated by the CMB observations, is thus necessary to confirm the idea of a fundamental preferred frame.
5. The modern ether-drift experiments
Searching for a non-thermal light anisotropy, which could be detected with light
propagating in vacuum and/or in solid dielectrics, we will now compare with the
modern experiments44
where
∆c¯θ c
∆ν(θ) ν0
is now extracted from the frequency shift
of two optical resonators, see Fig.3. The particular type of laser-cavity coupling used
in the experiments is known in the literature as the Pound-Drever-Hall system,85, 86
pNote that in Eq.(43) the gas refractivity drops out. The old estimates74, 80 of about 1 mK, based on the relation ϵgas∆T gas(θ)/T (vM 2 iller/2c2), with vMiller 10 km/s, were slightly too large.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 35
see Blacks tutorial article87 for a beautiful introduction. A laser beam is sent into
a Fabry-Perot cavity which acts as a filter. Then, a part of the output of the cavity
is fed back to the laser to suppress its frequency fluctuations. This method provides
a very narrow bandwidth and has been crucial for the precision measurements we
are going to describe.
The first application to the ether-drift experiments was realized by Brillet and
Hall in 1979.88 They were comparing the frequency of a CH4 reference laser (fixed in the laboratory) with the frequency of a cavity-stabilized He-Ne laser (ν0 8.8 · 1013
Hz) placed on a rotating table. Since the stabilizing optical cavity was placed inside
a vacuum envelope, the measured shift ∆ν(θ) was giving a measure of the anisotropy
of the velocity of light in vacuum. The short-term stability of the cavity-laser system
was found to be about ± 20 Hz for a 1-second measurement, and comparable to the
stability of the reference CH4 laser. It was also necessary to correct the data for a substantial linear drift of about 50 Hz/s.
By grouping the data in blocks of 10-20 rotations they found a signal with a
typical amplitude |∆ν| 17 Hz (or a relative level 1013) and with a phase θ2(t) which was randomly varying. Therefore, by increasing the statistics and projecting
along the axis corresponding to the Earth cosmic velocity obtained from the first
CMB observations,89 the surviving average effect was substantially reduced down
to about ±1 Hz. Finally, by further averaging over a period of about 200 days, the
residual ether-drift effect was an average frequency shift ⟨∆ν⟩ =0.13 ± 0.22 Hz, i.e.
about 100 times smaller than the instantaneous |∆ν|.
Since the 1979 Brillet-Hall article, substantial improvements have been intro-
duced in the experiments. Just to have an idea, in present-day measurements68, 90
with vacuum cavities the typical magnitude of the instantaneous fractional signal
|∆ν|/ν0 has been lowered from 1013 to 1015, the linear drift from 50 Hz/s to about 0.05 Hz/s and, after averaging over many observations, the best limit which
is
reported
is
a
residual
∆ν ν0
1018,68
i.e.
about
1000
times
smaller
than
the
instantaneous 1015 signal.
The assumptions behind the analysis of the data, however, are basically un-
changed. In fact, a genuine ether drift is always assumed to be a regular phenomenon
depending deterministically on the Earth cosmic motion and averaging more and
more observations is considered a way of improving the accuracy. But, as empha-
sized in Sect.4, the classical experiments indicate genuine physical fluctuations that
are not spurious noise but, instead, express how the cosmic motion of the Earth is
actually seen in a detector. For this reason, we will first consider the instantaneous
signal and try to understand if it can admit a physical interpretation.
5.1. A 109 refractivity for the vacuum (on the Earth surface)
As anticipated, after averaging many observations, the best limit which is reported for measurements with vacuum resonators is a residual ⟨∆ν/ν0⟩ ≲ 1018.68 This just reflects the very irregular nature of the signal because its typical magnitude
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
36 Authors Names
|∆ν|/ν0 1015 is about 1000 times larger, see Fig.19 or panel b) of Fig.11.
Fig. 19. The experimental frequency shift reported in Fig.9(a) of ref.90 (courtesy Optics Com-
munications). The black dots give the instantaneous signal, the red dots give the signal averaged over 1640 sequences. For a laser frequency ν0 = 2.8 · 1014 Hz a ∆ν = ±1 Hz corresponds to a fractional value ∆ν/ν0 of about ±3.5 · 1015.
The most interesting aspect however is that this 1015 instantaneous signal,
found in the room-temperature experiments of refs.90 and,68 was also found in ref.91
where the solid parts of the vacuum resonators were made of different material and
even in ref.92 were the apparatus was operating in the cryogenic regime. Since
it is very unlike that spurious effects (e.g. thermal noise93) remain the same for
experiments operating in so different conditions, one can meaningfully explore the
possibility that such an irregular 1015 signal admits a physical interpretation.
In the same model discussed for the classical experiments, we are then lead to
the concept of a refractive index Nv = 1 + ϵv for the vacuum or, more precisely, for the physical vacuum established in an optical cavity, as in Fig.10, when this
is placed on the Earth surface. The refractivity ϵv should be at the 109 level, in
order
to
give
|∆c¯θ | c
ϵv
(v2/c2)
1015 and thus would fit with the original idea
of94 where, for an apparatus placed on the Earths surface, a vacuum refractivity
ϵv (2GN M/c2R) 1.39·109 was considered, GN being the Newton constant and M and R the mass and radius of the Earth. Since this idea will sound unconventional
to many readers, we have first to recall the main motivations.
An effective refractivity for the physical vacuum becomes a natural idea when
adopting a different view of the curvature effects observed in a gravitational field. In
General Relativity these curvature effects are viewed as a fundamental modification
of Minkowski space-time. However, it is an experimental fact that many physical
systems for which, at a fundamental level, space-time is exactly flat are nevertheless
described by an effective curved metric in their hydrodynamic limit, i.e. at length
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 37
scales much larger than the size of their elementary constituents. For this reason, several authors, see e.g.,9597 have explored the idea that Einstein gravity might represent an emergent phenomenon and started to considered those gravity-analogs (moving fluids, condensed matter systems with a refractive index, Bose-Einstein condensates,...) which are known in flat space.
The main ingredient of this approach consists in the introduction of some background fields sk(x) in flat space expressing the deviations of the effective metric gµν(x) from the Minkowski tensor ηµν, i.e.
ν (x) ηµν = δgµν [sk(x)]
(44)
with δgµν[sk = 0] = 0. In this type of approach, to (partially) fill the conceptual gap with classical General Relativity, as in the original Yilmaz derivation,98 one could impose that Einsteins equations for the metric become algebraic identities which follow directly from the equations of motion for the sks in flat space, after introducing a suitable stress tensor tµν (sk) q.
As an immediate consequence, suppose that the sks represent excitations of the physical vacuum which therefore vanish identically in the equilibrium state. Then, if curvature effects are only due to departures from the lowest-energy state, one could immediately understand97 why the huge condensation energy of the unperturbed vacuum plays no role and thus obtain an intuitive solution of the cosmologicalconstant problem found in connection with the vacuum energy r.
Here, in our context of the ether-drift experiments, we will limit ourselves to explore some phenomenological consequence of this picture. To this end, let us assume a zeroth-order model of gravity with a scalar field s0(x) which behaves as the Newtonian potential (at least on some coarse-grained scale and consistently with the experimental verifications of the 1/r law at the sub-millimeter level100). How could the effects of s0(x) be effectively re-absorbed into a curved metric structure? At a pure geometrical level and regardless of the detailed dynamical mechanisms, the standard basic ingredients would be: 1) space-time dependent modifications of the physical clocks and rods and 2) space-time dependent modifications of the velocity of light s.
qIn the simplest, original Yilmaz approach98 there is only one inducing-gravity field s0(x) which plays the role of the Newtonian potential. Introducing its stress tensor tµν (s0) = ∂µs0∂ν s0 + 1/2δνµ ∂αs0∂αs0, to match the Einstein tensor, produces differences from the Schwarzschild metric which are beyond the present experimental accuracy. rThis is probably the simplest way to follow Feynmans indication: “The first thing we should understand is how to formulate gravity so that it doesnt interact with the vacuum energy”.99 sThis point of view can be well represented by some citations. For instance, “It is possible, on
the one hand, to postulate that the velocity of light is a universal constant, to define natural
clocks and measuring rods as the standards by which space and time are to be judged and then
to discover from measurement that space-time is really non-Euclidean. Alternatively, one can
define space as Euclidean and time as the same everywhere, and discover (from exactly the same
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
38 Authors Names
Within this interpretation, one could thus try to check the fundamental assumption of General Relativity that, even in the presence of gravity, the velocity of light in vacuum cγ is a universal constant, namely it remains the same, basic parameter c entering Lorentz transformations. Notice that, here, we are not considering the so called coordinate-dependent speed of light. Rather, our interest is focused on the value of the true, physical cγ as, for instance, obtained from experimental measurements in vacuum optical cavities placed on the Earth surface.
To spell out the various aspects, a good reference is Cooks article “Physical time and physical space in general relativity”.103 This article makes extremely clear which definitions of time and length, respectively dT and dL, are needed if all observers have to measure the same, universal speed of light (“Einstein postulate”). For a static metric, these definitions are dT 2 = g00dt2 and dL2 = gijdxidxj. Thus, in General Relativity, the condition ds2 = 0, which governs the propagation of light, can be expressed formally as
ds2 = c2dT 2 dL2 = 0
(45)
and, by construction, yields the same universal speed c = dL/dT .
For the same reason, however, if the physical units of time and space were instead defined to be dTˆ and dLˆ with, say, dT = q dTˆ and dL = p dLˆ, the same condition
ds2 = c2q2dTˆ2 p2dLˆ2 = 0
(46)
would now be interpreted in terms of the different speed
dLˆ q c
cγ = dTˆ = c p ≡ Nv
(47)
The possibility of different standards for space-time measurements is thus a simple
motivation for an effective vacuum refractive index Nv ̸= 1. With these premises, the unambiguous point of view of Special Relativity is that
the right space-time units are those for which the speed of light in the vacuum cγ, when measured in an inertial frame, coincides with the basic parameter c entering
Lorentz transformations. However, inertial frames are just an idealization. Therefore
the appropriate realization is to assume local standards of distance and time such
that the identification cγ = c holds as an asymptotic relation in the physical conditions which are as close as possible to an inertial frame, i.e. in a freely falling frame
(at least by restricting light propagation to a space-time region small enough that
measurements) how the velocity of light and natural clocks, rods and particle inertias really behave in the neighborhood of large masses”.101 Or “Is space-time really curved? Isnt it conceivable that space-time is actually flat, but clocks and rulers with which we measure it, and which we regard as perfect, are actually rubbery? Might not even the most perfect of clocks slow down or speed up and the most perfect of rulers shrink or expand, as we move them from point to point and change their orientations? Would not such distortions of our clocks and rulers make a truly flat space-time appear to be curved? Yes.”102
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 39
Fig. 20. A pictorial representation of the effect of a heavy mass M carried on board of a freelyfalling system, case (b). With respect to the ideal case (a), the mass M modifies the local space-time units and could introduce a vacuum refractivity so that now cγ ̸= c.
tidal effects of the external gravitational potential Uext(x) can be ignored). Note that this is essential to obtain an operational definition of the otherwise unknown parameter c.
As already discussed in ref.,94 light propagation for an observer S sitting on the Earths surface can then be described with increasing degrees of accuracy starting from step i), through ii) and finally arriving to iii):
i) S is considered a freely falling frame. This amounts to assume cγ = c so that, given two events which, in terms of the local space-time units of S, differ by (dx, dy, dz, dt), light propagation is described by the condition (ff=free-fall)
(ds2)ff = c2dt2 (dx2 + dy2 + dz2) = 0
(48)
ii) To a closer look, however, an observer S placed on the Earth surface can only be considered a freely-falling frame up to the presence of the Earth gravitational field. Its inclusion can be estimated by considering S as a freely-falling frame, in the same external gravitational field described by Uext(x), that however is also carrying on board a heavy object of mass M (the Earth mass itself) which affects the local space-time structure, see Fig.20. To derive the required correction, let us denote by δU the extra Newtonian potential produced by the heavy mass M at the experimental set up where one wants to describe light propagation. According to General Relativity, and to first order in δU , light propagation for the S observer is
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
40 Authors Names
now described by
ds2
=
c2dt2(1
2
|δU | c2 )
(dx2
+
dy2
+
dz2)(1
+
2
|δU | c2 )
c2dT
2
dL2
=
0
(49)
where
dT 2
=
(1
2
|δU c2
|
)dt2
and
dL2
=
(1
+
2
|δU c2
|
)(dx2
+
dy2
+
dz2)
are
the
phys-
ical units of General Relativity in terms of which one obtains the universal value
dL/dT = cγ = c.
Though, to check experimentally the assumed identity cγ = c one should compare
with a theoretical prediction for (c cγ) and thus necessarily modify some formal
ingredient of General Relativity. As a definite possibility, let us maintain the same definition of the unit of length dLˆ = dL but change the unit of time from dT to dTˆ.
The reason derives from the observation that physical units of time scale as inverse
frequencies and that the measured frequencies ωˆ for δU ̸= 0, when compared to the
corresponding value ω for δU = 0, are red-shifted according to
|δU |
ωˆ = (1 c2 ) ω
(50)
Therefore,
rather
than
the
natural
unit
of
time
dT
=
(1
|δU c2
|
)dt
of
General
Rela-
tivity, one could consider the alternative, but natural (see our footnote s), unit of
time
dTˆ
=
(1
+
|δU | c2 )
dt
(51)
Then, to reproduce ds2 = 0, we can introduce a vacuum refractive index
|δU |
Nv 1 + 2 c2
(52)
so that the same Eq.(49) takes now the form
ds2
=
c2dTˆ2 Nv2
dLˆ2
=
0
(53)
This
gives dLˆ/dTˆ = cγ
=
c Nv
and,
for
an
observer
placed
on
the
Earths surface,
a
refractivity
ϵv
= Nv 1
2GN M c2R
1.39 · 109
(54)
M and R being respectively the Earth mass and radius. Notice that, with this natural definition dTˆ, the vacuum refractive index asso-
ciated with a Newtonian potential is the same usually reported in the literature, at least since Eddingtons 1920 book,104 to explain in flat space the observed deflec-
tion of light in a gravitational field. The same expression is also suggested by the
formal analogy of Maxwell equations in General Relativity with the electrodynamics of a macroscopic medium with dielectric function and magnetic permeability105
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 41
ϵik
=
µik
=
g
. (gik )
g00
Indeed,
in
our
case,
from
the
relations
gilglk
=
δki
,
(gik) δki g00 , ϵik = µik = δki Nv , we obtain
√ Nv g
(1
2
|δU | c2 )(1
+
2
|δU c2
|
)3
1
+
|δU | 2 c2
(55)
A difference is found with Landaus and Lifshitz textbook107 where the vacuum
refractive index entering the constitutive relations is instead defined as Nv √g100
1
+
|δU c2
|
.
Concerning,
these
two
possible
definitions
of
Nv ,
we
address
the
reader
to Broekaerts article,106 see his footnote 3, where a very complete set of references
for the vacuum refractive index in gravitational field is reported. However, this
difference of a factor of 2 is not really essential and can be taken into account as a
theoretical uncertainty. The main point is that cγ, as measured in a vacuum cavity on the Earths surface (panel (b) in our Fig.20), could differ at a fractional level
109 from the ideal value c, as operationally defined with the same apparatus in a
true freely-falling frame (panel (a) in our Fig.20). In conclusion, this cγ c difference can be conveniently expressed through a vacuum refractivity of the form
ϵv
=
Nv
1
χ 2
1.39 · 109
(56)
where the factor χ/2 (with χ= 1 or 2) takes into account the mentioned theoretical uncertainty.
iii) Could one check experimentally if Nv ̸= 1? Today, the speed of light in
vacuum is assumed to be a fixed number with no error, namely 299 792 458 m/s.
Thus if, for instance, this estimate were taken to represent the value measured on the
Earth surface, in an ideal freely-falling frame there could be a slight increase, namely
+
χ 2
(0.42)
m/s
with
χ
=
1
or
2.
It
seems
hopeless
to
measure
unambiguously
such
a difference because the uncertainty of the last precision measurements performed
before the exactness assumption had precisely this order of magnitude, namely
±4 · 109 at the 3-sigma level or, equivalently, ±1.2 m/s.108
Therefore, as pointed out in ref.,94 an experimental test cannot be obtained from
the value of the isotropic speed in vacuum but, rather, from its possible anisotropy.
In fact, with a preferred frame and for Nv ̸= 1, an isotropic propagation as in
Eq.(53) would only be valid for a special state of motion of the Earth laboratory.
This provides the definition of Σ while for a non-zero relative velocity there would
be off diagonal elements g0i ̸= 0 in the effective metric.105 If Σ exists, we would then
expect
a
light
anisotropy
|∆c¯θ | c
ϵv (v/c)2
1015,
consistently
with
the
presently
measured value.
5.2. Some important technical aspects
Before considering the experiments, however, a rather technical discussion is necessary for an in-depth comparison with the data. In the mentioned cryogenic experi-
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
42 Authors Names
Fig. 21. The stability of the fractional frequency shift of ref.92 at different integration times. The upper solid curve, denoted as newCORE, reports the actual measurements with the cryogenic apparatus in 2013. The lower solid, dashed and dot-dashed curves, denoted as predicted newCORE, indicate future stability limits (2 ÷ 4) · 1017 that could be foreseen at that time.
ment of ref.,92 the instantaneous signal is not shown explicitly. However, its magnitude can be deduced from its typical variation observed over a characteristic time of 1÷2 seconds, see Fig.21. For a very irregular signal, in fact, this typical variation, of about 1015, gives the magnitude of the instantaneous signal itself and, indeed, it is in good agreement with the mentioned room-temperature measurements.
Fig. 22. The typical trend of the RAV for a signal in various regimes. The minimum of the white-noise trend τ 0.5 defines the value τ = τ¯ indicated by the arrow.
The quantity which is reported in Fig.21 is the Root Square of the Allan Variance (RAV) of the fractional frequency shift. In general, the RAV describes the variation
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 43
obtained by sampling a function f = f (t) at steps of time τ . By defining
1 ti+τ
f (ti; τ ) = τ ti dt f (t) ≡ f i
(57)
one generates a τ dependent distribution of f i values. In a large time interval Λ = M τ , the RAV is then defined as
RAV (f, τ ) = RAV 2(f, τ )
(58)
where
RAV 2(f, τ ) =
1
M 1
2(M 1)
f i f i+1
2
i=1
(59)
and the factor of 2 is introduced to obtain the standard variance σ(f ) for uncorrelated data with zero mean, as for a pure white-noise signal.
Note that the actual measurements in Fig.21 are indicated by the upper solid curve denoted as newCORE. These were obtained with the cryogenic apparatus in 2013 (CORE=Cryogenic Optical REsonators) and were giving a stability at the level of about 1.2 · 1015. The lower solid, dashed and dot-dashed curves, denoted as predicted newCORE, indicate instead possible improved limits (2 ÷ 4) · 1017 that could be foreseen at that time. As a matter of fact, these limits have not yet been achieved because the highest stability limits are still larger by an order of magnitude. This persistent signal, which is crucial for our work, does not depend on the absolute temperature and/or the characteristics of the optical cavities.109
After this preliminaries, we then arrive at our main point. As anticipated, numerical simulations in our stochastic model indicate that our basic signal has the same characteristics as a universal white noise. This means that it should be compared with the frequency shift of two optical resonators at the largest integration time τ¯ where the pure white-noise component is as small as possible but other disturbances, that can affect the measurements, are not yet important, see Fig.22. In the experiments we are presently considering this τ¯ is typically 1 ÷ 2 seconds so that one gets the relation with the average magnitude of the instantaneous signal
RAV (∆ν, τ¯) σ(∆ν) ⟨|∆ν|⟩stat
(60)
5.3. Comparing our model with experiments in vacuum
We will now compare with the type of signal observed in68, 90 in vacuum at room temperature. To this end, we will use the relation which connects the frequency shift between two orthogonal resonators ∆ν(θ; t) = ν1(θ; t) ν2(θ + π/2; t) to the angular dependence of the velocity of light, namely see (33)
∆ν(θ; t) = ∆c¯θ(t) = 2S(t) sin 2θ + 2C(t) cos 2θ
(61)
ν0
c
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
44 Authors Names
where S(t) and C(t) are given in Eqs.(34). As in the case of the classical experiments, the velocity components vx(t) and vy(t) will be expressed as random Fourier series through the Eqs.(83) and (84) of the Appendix. A simulation of two short-time sequences of 2C(t) and 2S(t) is shown in Fig.23.
Fig. 23. For ϵv as in Eq.(56) and χ = 2, we report a simulation of two sequence of 45 seconds for the functions 2C(t) and 2S(t) Eqs.(34). Units are 1015 and the two sets belong to the same random sequence for two sidereal times that differ by 6 hours. The boundaries of the stochastic velocity components, Eqs.(83) and (84) of the Appendix, are controlled by (V, α, γ)CMB through Eqs.(77) and (87). For a laser frequency of 2.8 · 1014 Hz, the range ±3.5 · 1015 corresponds to a typical frequency shift ∆ν in the range ±1 Hz, as in our Fig.19.
For a quantitative test, we concentrated on the observed value of the RAV of the frequency shift at the end point τ¯ = 1 ÷ 2 seconds of the white-noise branch of the spectrum, see Fig.3, bottom part of68 . This has a value
[RAV (∆ν, τ¯)]exp = (0.20 ÷ 0.24) Hz
(62)
or, in units of the reference frequency ν0 = 2.8 · 1014 Hz 68
∆ν RAV ( , τ¯)
= (7.8 ± 0.7) · 1016 Vacuum room temperature (63)
ν0
exp
As anticipated, our instantaneous, stochastic signal for ∆ν(t) is, to very good
approximation, a pure white noise for which the RAV coincides with the standard
variance. At the same time, for a very irregular signal with zero mean of the type
shown in Fig. 23, but whose magnitude can have a long-term time dependence, one
should replace in Eq. (60) ⟨|∆ν|⟩stat → ⟨|∆ν(t)|⟩stat and evaluate the RAV in the
corresponding temporal range. Therefore, from
∆ν ν0
=
∆c¯θ c
ϵv
·
v2 c2
,
we
arrive
at
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 45
our prediction
∆ν(t)
RAV (
, τ¯)
ν0
theor
ϵv
⟨vx2(t) + vy2(t)⟩stat c2
π2 18
·
ϵv
·
V2 c2
sin2
z(t)
(64)
Then, by using Eq.(56), for the projection v˜(t) = V | sin z(t)| = 250 ÷ 370 Km/s
used for the classical experiments, our prediction for the RAV can be expressed as
∆ν RAV ( , τ¯)
χ · (8.5 ± 3.5) · 1016
(65)
ν0
theor
2
with χ = 1 or 2. By comparing with the experimental Eq.(63), the data favour
χ = 2, which is the only free parameter of our scheme. Also, the good agreement
with our theoretical value indicates that, at the end point of the white-noise part
of the signal, the corrections to our simplest model should be small.
Notice, however, that the range in Eq.(65) is not a theoretical uncertainty but reflects the daily variations of V 2 sin2 z(t) in Eq.(64). This means that, depending
on the sidereal time, the measurements of the RAV at the white-noise end point
τ = τ¯ should exhibit definite daily variations in the range (for χ = 2)
5 · 1016 ≲
∆ν RAV ( , τ¯)
ν0
theor
12 · 1016
(66)
Thus it becomes crucial to understand whether these variations can be observed.
5.4. Comparing our model with experiments in solids
To consider modern experiments in solid dielectrics, we will compare with the very precise work of ref.70 This is a cryogenic experiment, with microwaves of 12.97 GHz,
where almost all electromagnetic energy propagates in a medium, sapphire, with re-
fractive index of about 3 (at microwave frequencies). As anticipated, with a thermal
interpretation of the residuals in gaseous media, we expect that the fundamental 1015 vacuum signal considered above, with very precise measurements, should also
become visible here. In particular, the large refractivity of the solid Nsolid 1 = O(1) should play no role.
Following refs.,810 we first observe that for Nv = 1 + ϵv there is a very tiny difference between the refractive index defined relatively to the ideal vacuum value
c and the refractive index relatively to the physical isotropic vacuum value c/Nv measured on the Earth surface. The relative difference between these two definitions is proportional to ϵv ≲ 109 and, for all practical purposes, can be ignored. All materials would now exhibit, however, the same background vacuum anisotropy. To
this end, let us replace the average isotropic value
c
c
Nsolid
Nv Nsolid
(67)
and then use Eq.(18) to replace Nv in the denominator with its θdependent value
N¯v(θ) 1 + ϵvβ2(1 + cos2 θ)
(68)
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
46 Authors Names
This is equivalent to define a θdependent refractive index for the solid dielectric
N¯solid(θ) Nsolid
1
+
ϵv β 2 (1
+
cos2
θ)
(69)
so that
c
c
[c¯γ (θ)]solid = N¯solid(θ) Nsolid
1 ϵvβ2(1 + cos2 θ)
(70)
with an anisotropy
[∆c¯θ ]solid [c/Nsolid]
ϵvβ2 cos 2θ
1015
(71)
In this way, a genuine 1015 vacuum effect, if there, could also be detected in a solid dielectric thus implying the same prediction Eq.(65).
Fig. 24. We report two typical sets of 2000 seconds for our basic white-noise (WN) signal and its colored version obtained by Fourier transforming the spectral amplitude of ref.70 The boundaries of the random velocity components Eqs.(83) and (84) were defined by Eq.(87) by plugging in Eq.(77) the CMB kinematical parameters, for a sidereal time t = 4000 6000 seconds and for the latitude of Berlin-Duesseldorf, see the Appendix. The figure is taken from ref.10
In ref.,10 a detailed comparison with70 was performed. First, from Figure 3(c) of,70 see also panel b) of our Fig.12, it is seen that the spectral amplitude of this particular apparatus becomes flat at frequencies ω ≥ 0.5 Hz indicating that the end-point of the white-noise branch of the signal is at an integration time τ¯ 1 ÷ 2 seconds. The data for the spectral amplitude were then fitted to an analytic, powerlaw form to describe the lower-frequency part 0.001 Hz ≤ ω ≤ 0.5 Hz which reflects apparatus-dependent disturbances. This fitted spectrum was then used to generate a signal by Fourier transform. Finally, very long sequences of this signal were stored to produce “colored” version of our basic white-noise signal.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 47
To get a qualitative impression of the effect, we report in Fig.24 a sequence of our basic simulated white-noise signal and a sequence of its colored version. By averaging over many 2000-second sequences of this type, the corresponding RAVs for the two simulated signals are then reported in Fig.25. The experimental RAV extracted from Figure 3(b) of ref.70 is also reported (for the non-rotating setup). At this stage, the agreement of our simulated, colored signal with the experimental data remains satisfactory only up τ = 50 seconds. Reproducing the signal at larger τ s would have required further efforts but this is not relevant, our scope being just to understand the modifications of our stochastic signal near the endpoint of the white-noise spectrum.
Fig. 25. We report the RAV for the fractional frequency shift obtained from many simulations of sequences of 2000 seconds for our basic white-noise (WN) signal (decreasing as τ 0.5) and for its colored version, see Fig.24. The direct experimental results of ref.,70 for the non-rotating setup, are also shown as red dots. The figure is taken from ref.10
As one can check from Fig.3(b) of ref.,70 see also the red dots in our Fig.25, the experimental RAV for the fractional frequency shift, at the white-noise end point τ¯ 1 ÷ 2 second, is in the range (6.8 ÷ 8.6) · 1016, say 70
∆ν RAV ( , τ¯)
= (7.7 ± 0.9) · 1016 Solid cryogenic
(72)
ν0
exp
As such, it coincides with Eq.(63) that we extracted from ref.68 after normalizing
their experimental result RAV (∆ν, τ¯)exp = 0.20 ÷ 0.24 Hz to their laser frequency ν0 = 2.8 · 1014 Hz. At the same time, it is well consistent with our theoretical prediction Eq.(65) for χ = 2. Therefore this beautiful agreement, between ref.68
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
48 Authors Names
(a vacuum experiment at room temperature) and ref.70 (a cryogenic experiment in a solid dielectric), on the one hand, and with our theoretical prediction Eq.(65), on the other hand, confirms our interpretation of the data in terms of a stochastic signal associated with the Earth cosmic motion within the CMB and determined by the vacuum refractivity ϵv Eq.(56), for χ = 2.
Two ultimate experimental checks still remain. First, as anticipated, one should try to detect our predicted, daily variations Eq.(66). Due to the excellent systematics, these variations should remain visible with both experimental setups. Second, one more complementary test should be performed by placing the vacuum (or solid dielectric) optical cavities on board of a satellite, as in the OPTIS proposal.110 In this ideal free-fall environment, as in panel (a) of our Fig.20, the typical instantaneous frequency shift should be much smaller (by orders of magnitude) than the corresponding 1015 value measured with the same interferometers on the Earth surface.
6. Summary and outlook
In this paper, we started from the present, basic ambiguity concerning the version of relativity which is physically realized in nature, namely Einstein Special Relativity vs. a Lorentzian formulation with a preferred reference frame Σ. This ambiguity is usually presented by a two-step argument. First, the basic quantitative ingredients, namely Lorentz transformations, are the same in both formulations. Second, even in a Lorentzian formulation, Michelson-Morley experiments can only produce null results. Therefore, rather than introducing an experimentally unobservable and logically superfluous entity, it seemed more satisfactory to adopt the point of view of Special Relativity where those effects (length contraction and time dilation), that were at the base of the original Lorentzian formulation, so to speak, become part of the kinematics. In this way, relativity becomes axiomatic and extendable beyond the original domain of the electromagnetic phenomena. This wider perspective has been the main reason for the traditional supremacy given to Einsteins view.
However, discarding all historical aspects, it was emphasized by Bell that a change of perspective, from Special Relativity to a Lorentzian formulation, could be crucial to reconcile hypothetical faster-than-light signals with causality, as with the apparent non-local aspects of the Quantum Theory. In addition, the present view of the lowest-energy state as a Bose condensate of elementary quanta (Higgs particles, quark-antiquark pairs, gluons...), indicates a vacuum structure with some degree of substantiality which could characterize non trivially the form of relativity which is physically realized in nature. So, there may be good reasons for a preferred reference frame but, without the possibility of detecting experimentally an ether wind in laboratory, the difference between the two formulations remains a philosophical problem.
This impossibility-in-principle, however, is somewhat mysterious. While it is certainly true that evidence for both the undulatory and corpuscular aspects of radia-
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 49
tion has substantially modified the consideration of an underlying ethereal medium,
yet, if Σ exists, only a conspiracy of relativistic effects would make undetectable
our motion with respect to it. But this conspiracy works exactly if the velocity of
light cγ propagating in the various interferometers, or more precisely its two-way
combination c¯γ, coincides with the basic parameter c entering Lorentz transforma-
tions. Therefore if c¯γ ̸= c, as for instance in the presence of matter, where light gets
absorbed and then re-emitted, nothing would really prevent an angular dependence
∆c¯θ = c¯γ(π/2 + θ) γ(θ) ̸= 0. If an angular dependence can be detected, and cor-
related with the cosmic motion of the Earth, the long sought Σ tight to the CMB
could finally emerge.
We have thus recalled the two key points of our extensive work. First, one
should impose that all measurable effects vanish exactly in the c¯γ → c limit, i.e.
in the ideal vacuum limit of a refractive index N = 1. Instead, in the infinitesimal
region
N
=
1+ϵ
simple
symmetry
arguments
lead
to
the
relation
|∆c¯θ | c
ϵ(v2/c2).
For a typical cosmic v 300 km/s and ϵ = 2.8 · 104 , for air, or ϵ = 3.3 · 105, for
gaseous helium, this reproduces the order of magnitude of the effects observed in
the classical experiments.
The other peculiar aspect of our analysis concerns the observed, irregular charac-
ter of the data that, giving often substantially different directions of the drift at the
same hour on consecutive days, were contradicting the traditional expectation of a
regular phenomenon completely determined by the cosmic motion of the Earth. As
we have emphasized, here again, there may be a logical gap. The relation between
the macroscopic motion of the Earth and the microscopic propagation of light in a
laboratory depends on a complicated chain of effects and, ultimately, on the physi-
cal nature of the vacuum. By comparing with the motion of a body in a fluid, the
standard view corresponds to a form of regular, laminar flow where the projection
v˜µ(t) of the global, cosmic velocity, at the site of the experiment, coincides with the
local vµ(t) that determines the signal in the plane of the interferometer. Instead,
some general arguments and some experimental analogies suggest that the physical
vacuum might rather resemble a turbulent fluid where large-scale and small-scale
flows are only related indirectly. In this different perspective, with forms of tur-
bulence that, as in most models, become statistically isotropic at small scales, the
local vµ(t) would fluctuate randomly within boundaries fixed by the global v˜µ(t)
(see the Appendix). Therefore, one should analyze the data in phase and amplitude
(giving respectively the instantaneous direction and magnitude of the drift) and
concentrate on the latter which is positive definite and remains non-zero under any
averaging procedure. In this way, by restricting to the amplitudes, experiments al-
ways believed in contradiction with each other, as Miller vs. Piccard-Stahel, become
consistent, see Fig.9. Most notably, by adopting the parameters (V, α, γ)CMB to fix
the boundaries of the local random vµ(t) in our stochastic model, one finds a good
description of the irregular behaviour of the amplitudes extracted from Joos very
precise observations (see Figs.17 and 18). Viceversa, by fitting Joos amplitudes
with Eqs.(76) and (77), one finds a right ascension α(fit Joos) = (168 ± 30) de-
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
50 Authors Names
grees and an angular declination γ(fit Joos) = (13 ± 14) degrees which are well
consistent with the present values α(CMB) 168 degrees and γ(CMB) 7 de-
grees. The summary of all classical experiments given in Table 3 shows the complete
consistency with our theoretical predictions.
To conclude our analysis of the classical experiments in gaseous systems, we
have
emphasized
that
our
basic
relation
|∆c¯θ | c
ϵgas(v2/c2)
derives
from
general,
symmetry arguments and does not explain the ultimate origin of the tiny observed
residuals. Due to the consistency with the velocity of 370 km/s, a plausible explana-
tion consists in a collective interaction of gaseous matter with the CMB radiation.
This could bring the gas out of equilibrium as if there were an effective temperature
difference, |∆T gas(θ)| = 0.2 ÷ 0.3 mK, in the gas along the two optical paths. This
magnitude is slightly smaller than the value of about 1 mK considered by Joos
and Shankland and, being just a small fraction of the whole ∆T CMB(θ) = ±3.3
mK in Eq.(3), indicates the weakness of the collective gas-CMB interactions. Most
notably, the thermal interpretation leads to an important prediction. In fact, it
implies that if a physical signal could definitely be detected in vacuum then, with
very precise measurements, the same signal should also show up in a solid dielectric
where disturbing temperature differences of a fraction of millikelvin become irrele-
vant. Detecting such non-thermal light anisotropy, through the combined analysis
of the modern experiments in vacuum and in solid dielectrics, for the same cosmic
motion indicated by the classical experiments, is thus necessary to confirm the idea
of a fundamental preferred frame.
Despite the much higher precision of modern experiments, the assumptions be-
hind the analysis of the data are basically the same as in the classical experiments.
A genuine signal is assumed to be a regular phenomenon, depending deterministi-
cally on the Earth cosmic motion, so that averaging more and more observations is
considered a way of improving the accuracy. But the classical experiments indicate
genuine physical fluctuations which are not spurious noise and, instead, express how
the cosmic motion of the Earth is actually seen in a detector. Therefore, the present
quoted
average,
namely
⟨∆c¯θ ⟩ c
1018,
could
just
reflect
the
very
irregular
nature
of
the
signal.
Indeed,
its
typical
instantaneous
magnitude
in
vacuum
|∆c¯θ | c
1015
is about 1000 times larger, see Fig.19 or panel b) of Fig.11.
To understand if this vacuum signal can admit a physical interpretation, a crucial
observation is that the same 1015 magnitude is found in measurements where the
resonators are made of different materials, in measurements at room-temperature
and also in the cryogenic regime. Since it is very unlike that spurious effects remain
the same in so different conditions, in the same model used for the classical experi-
ments we are driven to the idea of a refractive index Nv = 1 + ϵv for the vacuum or,
more precisely, for the physical vacuum established in an optical cavity placed on
the Earth surface. The refractivity ϵv should be at the 109 level, in order to give
|∆c¯θ | c
ϵv
(v2/c2)
1015 and thus would fit with the original idea of ref.94 The
motivation was that, if Einsteins gravity is a phenomenon which emerges, at some
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 51
small length scale, from a fundamentally flat space, for an apparatus placed on the
Earth surface (which is in free fall with respect to all masses in the Universe but not
with respect to the Earth, see Fig.20) there should be a tiny vacuum refractivity
ϵv (2GN M/c2R) 1.39 · 109, where GN is the Newton constant and M and R
are the mass and radius of the Earth. This is the same type of refractivity consid-
ered by Eddington, or much more recently by Broekaert, to explain in flat space the
deflection of light in a gravitational field. Therefore Michelson-Morley experiments,
by
detecting
a
light
anisotropy
|∆c¯θ | c
ϵv
(v2/c2)
1015, can also resolve this
other ambiguity.
With this identification of ϵv, we first compared qualitatively the observed signal,
in Fig.19 or in panel b) of Fig.11, with simulations in our stochastic model, see
Figs.23 and 24. For a more quantitative analysis, we then considered the value of a
particular statistical indicator which is used nowadays, namely the Allan Variance
of
the
fractional
frequency
shift
RAV
(
∆ν ν0
,
τ
)
as
function
of
the
integration
time
τ . Since the irregular signal of our stochastic model has the characteristics of a
universal white noise and should represent an irreducible component, we have thus
compared with the RAV measured at the end point of the white-noise branch of
the spectrum. This is defined as the largest integration time τ¯ where the white-
noise component is as small as possible but other spurious disturbances, that can
affect the measurements, are not yet important, see Fig.22. In this way, for the same
velocity range v˜ = 250÷370 km/s used for the classical experiments, our theoretical
prediction Eq.(65) (for χ = 2) is in very good agreement with the results of the most
precise experiment in vacuum Eq.(63).
But, then, the second crucial test. As anticipated, if this 1015 signal observed
in vacuum has a real physical meaning, the same effect should also be detected with
a very precise experiment in a solid dielectric, see Eq.(71). This expectation is con-
firmed by the extraordinary agreement between Eq.(72) and Eq.(63). Note that the
two experiments are completely different because in ref.70 light propagates in a solid
in the cryogenic regime and in ref.68 light propagates in vacuum at room tempera-
ture. As such, there is a plenty of systematic differences. Yet, the two experiments
give exactly the same signal at the white-noise end point. Therefore, there must be
an ubiquitous form of white noise that admits a definite physical interpretation. Our
theoretical prediction Eq.(65) is, at present, the only existing explanation. Together
with the classical experiments, we thus conclude that there is now an alternative
scheme challenging the traditional null interpretation of Michelson-Morley exper-
iments, always presented as a self-evident scientific truth.
We have also discussed two further experimental tests. First, one should try to
detect our predicted, daily variations Eq.(66). Second, one should also try to place
the optical cavities on a satellite, as in the OPTIS proposal.110 In this ideal free-fall
environment, as in panel (a) of our Fig.20, the typical instantaneous frequency shift
should be much smaller (by orders of magnitude) than the corresponding 1015
value measured with the same interferometers on the Earths surface.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
52 Authors Names
Appendix
In this appendix, we will summarize the stochastic model used in refs.710 to compare with experiments. To make explicit the time dependence of the signal let us first re-write Eq.(19) as
∆c¯θ (t) c
v2(t) ϵ c2
cos 2(θ
θ2(t))
(73)
where v(t) and θ2(t) indicate respectively the instantaneous magnitude and direction of the drift in the (x, y) plane of the interferometer. This can also be re-written as
∆c¯θ(t) 2S(t) sin 2θ + 2C(t) cos 2θ
(74)
c
with
2C (t)
=
ϵ
vx2(t) vy2(t) c2
2S(t)
=
ϵ
2vx (t)vy (t) c2
(75)
and vx(t) = v(t) cos θ2(t), vy(t) = v(t) sin θ2(t) As anticipated in Sect.3, the standard assumption to analyze the data has always
been based on the idea of regular modulations of the signal associated with a cosmic
Earth velocity. In general, this is characterized by a magnitude V , a right ascension
α and an angular declination γ. These parameters can be considered constant for
short-time observations of a few days where there are no appreciable changes due
to the Earth orbital velocity around the sun. In this framework, where the only
time dependence is due to the Earth rotation, the traditional identifications are v(t) ≡ v˜(t) and θ2(t) ≡ θ˜2(t) where v˜(t) and θ˜2(t) derive from the simple application of spherical trigonometry71
cos z(t) = sin γ sin ϕ + cos γ cos ϕ cos(τ α)
(76)
v˜(t) = V sin z(t)
(77)
v˜x(t) = v˜(t) cos θ˜2(t) = V [sin γ cos ϕ cos γ sin ϕ cos(τ α)]
(78)
v˜y(t) = v˜(t) sin θ˜2(t) = V cos γ sin(τ α)
(79)
Here z = z(t) is the zenithal distance of V, ϕ is the latitude of the laboratory,
τ
=
ωsidt
is
the
sidereal
time
of
the
observation
in
degrees
(ωsid
2π 23h 56
)
and
the
angle θ2 is counted conventionally from North through East so that North is θ2 = 0
and East is θ2 = 90o. With the identifications v(t) ≡ v˜(t) and θ2(t) ≡ θ˜2(t), one
thus arrives to the simple Fourier decomposition
S(t) ≡ S˜(t) = S0 + Ss1 sin τ + Sc1 cos τ + Ss2 sin(2τ ) + Sc2 cos(2τ ) (80)
C(t) ≡ C˜(t) = C0 + Cs1 sin τ + Cc1 cos τ + Cs2 sin(2τ ) + Cc2 cos(2τ ) (81)
where the Ck and Sk Fourier coefficients depend on the three parameters (V, α, γ) and are given explicitly in refs.7, 9
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 53
Though, the identification of the instantaneous quantities vx(t) and vy(t) with their counterparts v˜x(t) and v˜y(t) is not necessarily true. As anticipated in Sect.3, one could consider the alternative situation where the velocity field is a nondifferentiable function and adopt some other description, for instance a formulation in terms of random Fourier series.55, 72, 73 In this other approach, the parameters of the macroscopic motion are used to fix the typical boundaries for a microscopic velocity field which has an intrinsic non-deterministic nature.
The model adopted in refs.710 corresponds to the simplest case of a turbulence which, at small scales, appears homogeneous and isotropic. The analysis can then be embodied in an effective space-time metric for light propagation
ν (t) ηµν + 2ϵvµ(t)vν (t)
(82)
where vµ(t) is a random 4-velocity field which describes the drift and whose boundaries depend on a smooth field v˜µ(t) determined by the average Earth motion.
For homogeneous turbulence a series representation, suitable for numerical simulations of a discrete signal, can be expressed in the form
vx(tk) = [xn(1) cos ωntk + xn(2) sin ωntk]
(83)
n=1
vy(tk) = [yn(1) cos ωntk + yn(2) sin ωntk]
(84)
n=1
Here ωn = 2nπ/T and T is the common period of all Fourier components. Furthermore, tk = (k 1)∆t, with k = 1, 2..., and ∆t is the sampling time. Finally, xn(i = 1, 2) and yn(i = 1, 2) are random variables with the dimension of a velocity and vanishing mean. In our simulations, the value T = Tday= 24 hours and a sampling step ∆t = 1 second were adopted. However, the results would remain unchanged by any rescaling T → sT and ∆t → s∆t.
In general, we can denote by [dx(t), dx(t)] the range for xn(i = 1, 2) and by [dy(t), dy(t)] the corresponding range for yn(i = 1, 2). Statistical isotropy would require to impose dx(t) = dy(t). However, to illustrate the more general case, we will first consider dx(t) ̸= dy(t).
If we assume that the random values of xn(i = 1, 2) and yn(i = 1, 2) are chosen with uniform probability, the only non-vanishing (quadratic) statistical averages are
⟨x2n(i
=
1, 2)⟩stat
=
d2x(t) 3 n2η
⟨yn2 (i
=
1, 2)⟩stat
=
d2y (t) 3 n2η
(85)
Here, the exponent η ensures finite statistical averages ⟨vx2(t)⟩stat and ⟨vy2(t)⟩stat for an arbitrarily large number of Fourier components. In our simulations, between the two possible alternatives η = 5/6 and η = 1 of ref.,73 we have chosen η = 1 that
corresponds to the Lagrangian picture in which the point where the fluid velocity
is measured is a wandering material point in the fluid.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
54 Authors Names
Finally, the connection with the Earth cosmic motion is obtained by identifying dx(t) = v˜x(t) and dy(t) = v˜y(t) as given in Eqs. (76)(79). If, however, we require statistical isotropy, the relation
v˜x2(t) + v˜y2(t) = v˜2(t)
(86)
requires the identification
v˜(t)
dx(t) = dy(t) =
√ 2
(87)
For such isotropic model, by combining Eqs.(83)(87) and in the limit of an infinite statistics, one gets
⟨vx2 (t)⟩stat
=
⟨vy2(t)⟩stat
=
v˜2(t) 2
1 3
1 n2
v˜2(t) =
2
π2 18
n=1
⟨vx(t)vy(t)⟩stat = 0
(88)
and vanishing statistical averages
⟨C(t)⟩stat = 0
⟨S(t)⟩stat = 0
(89)
at any time t, see Eqs.(75). Therefore, by construction, this model gives a definite non-zero signal but, if the same signal were fitted with Eqs.(80) and (81), it would also give average values (Ck)avg = 0, (Sk)avg = 0 for the Fourier coefficients.
To understand how radical is the modification produced by Eqs.(89), we recall the traditional procedure adopted in the classical experiments. One was measuring the fringe shifts at some given sidereal time on consecutive days so that changes of the orbital velocity were negligible. Then, see Eqs.(20) and (74), the measured shifts at the various angle θ were averaged
∆λ(θ; t)
2D
⟨ λ ⟩stat = λ [2 sin 2θ ⟨S(t)⟩stat + 2 cos 2θ ⟨C(t)⟩stat]
(90)
and finally these average values were compared with models for the Earth cosmic
motion.
However if the signal is so irregular that, by increasing the number of mea-
surements, ⟨C(t)⟩stat → 0 and ⟨S(t)⟩stat → 0 the averages Eq.(90) would have no meaning. In fact, these averages would be non vanishing just because the statis-
tics is finite. In particular, the direction θ2(t) of the drift (defined by the relation tan 2θ2(t) = S(t)/C(t)) would vary randomly with no definite limit.
This is why one should concentrate the analysis on the 2nd-harmonic amplitudes
2D A2(t) = λ 2
S2(t) + C2(t)
2D λ
ϵ
vx2
(t)
+ c2
vy2
(t)
(91)
which are positive-definite and remain non-zero under the averaging procedure.
Moreover, these are rotational-invariant quantities and their statistical average
2D π2 V 2 sin2 z(t)
⟨A2(t)⟩stat
λ
· ·ϵ· 18
c2
(92)
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 55
would remain unchanged in the isotropic model Eq.(87) or with the alternative
choice dx(t) ≡ v˜x(t) and dy(t) ≡ v˜y(t). Analogous considerations hold for the mod-
ern
experiments
where
∆c¯θ (t) c
is
extracted
from
the
frequency
shift
of
two
optical
resonators. Again, the C(t) and S(t) obtained, through Eq.(74), from the very ir-
regular, measured signal (see e.g. Fig.19), are compared with the slowly varying
parameterizations Eqs.(80) and (81) to extract the Ck and Sk Fourier coefficients.
Then, by comparing with our simulation of the C(t) and S(t) in Fig.23, it is no
surprise if the average values (Ck)avg → 0, (Sk)avg → 0 by simply increasing the
number of observations.
References
1. A. Einstein, Zur Elektrodynamik bewegter Ko¨rper, Ann. der Physik, 17 (1905) 891;(English translation in: The Principle of Relativity by H. A. Lorentz et al., Methuen 1923)
2. H. A. Lorentz, Electromagnetic phenomena in a system moving with any velocity smaller than that of light, Proc. Acad. of Sci. Amsterdam, 6 1904.
3. A. Ungar,The relativistic composite-velocity reciprocity principle, Found. of Phys. 30, 331 (2000).
4. J. P. Costella, B. H. J. McKellar, A. A. Rawlinson, G. J. Stephenson Jr, The Thomas rotation, Am. J. Phys. 69, 837 (2001).
5. K. O Donnell and M. Visser,Elementary analysis of the special relativistic combination of velocities, Wigner rotation, and Thomas precession, Eur. J. Phys. 32, 1033 (2011).
6. Full reference to all papers by J. S. Bell, can be found in the Volume Collection, Speakable and Unspeakable in Quantum Mechanics, 2nd edition, Cambridge University Press, 2004.
7. M. Consoli, C. Matheson and A. Pluchino, The classical ether-drift experiments: a modern re-interpretation Eur. Phys. J. Plus, 128 (2013) 71.
8. M. Consoli and A. Pluchino, Cosmic Microwave Background and the issue of a fundamental preferred frame, Eur. Phys. Jour. Plus 133 (2018) 295.
9. M. Consoli and A. Pluchino, Michelson-Morley Experiments: an Enigma for Physics and the History of Science, World Scientific 2019, ISBN 978-981-3278-18-9.
10. M. Consoli and A. Pluchino, CMB, preferred reference system and dragging of light in the Earths frame, Universe 7 (2021) 311;arXiv:2109.03047 [physics.gen-ph].
11. T. Maudlin, Quantum Non-Locality and Relativity, Blackwell, Cambridge, 2011. 12. N. Gisin, Quantum Chance, Springer Int. Publ. 2014. 13. J. Bricmont, Making Sense of Quantum Mechanics, Springer Int. Publ. 2016. 14. A. Einstein, B. Podolski and N. Rosen, Can quantum-mechanical description of phys-
ical reality be considered complete?, Phys. Rev.47 (1935) 777. 15. D. Salart, A. Baas, C. Branciard, N. Gisin and H. Zbinden, Testing spooky action at
a distance, Nature 454 (2008) 861.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
56 Authors Names
16. J.-D. Bancal, S. Pironio, A. Acin, Y.-C. Liang, V. Scarani, and N. Gisin, Quantum nonlocality based on finite-speed causal influences leads to superluminal signaling, Nature Physics 8 (2012) 867.
17. P. A. M. Dirac, Development of the Physicists Conception of Nature, in The Physicists Conception of Nature, J. Mehra Ed., Reidel, Boston 1973.
18. A. Aspect, J. Dalibard, G. Roger, Experimental test of Bells inequalities using timevarying analyzers, Phys. Rev. Lett. 49, 1804 (1982).
19. J. F. Clauser, Experimental investigation of a polarization correlation anomaly, Phys. Rev. Lett. 36, 1223 (1976).
20. G. Weihs, T. Jennewein, C. Simon, H. Weinfurter, and A. Zeilinger, Violation of Bells Inequality under Strict Einstein Locality Conditions, Phys. Rev. Lett. 81, 5039 (1998)
21. A. Aspect, Speech at the Nobel Prize banquet, December 2022. 22. M. Consoli and E. Costanzo, Is the physical vacuum a preferred frame?, Eur. Phys.
Journ. C54 (2008) 285. 23. M. Consoli and E. Costanzo, Precision tests with a new class of dedicated ether-drift
experiments, Eur. Phys. Journ. C55 (2008) 469. 24. M. Consoli, Probing the vacuum of particle physics with precise laser interferometry,
Found. of Phys. 45, 22 (2015). 25. M. Consoli, A. Pluchino, P. Zizzi,Quantum Non-Locality and the CMB: what Experi-
ments say, Universe 8 (2022) 48; arXiv:2209.06858 [physics.gen-ph]. 26. R. Y. Chiao, Conceptual tensions between quantum mechanics and general relativ-
ity: Are there experimental consequences?, in “Science and Ultimate Reality: From Quantum to Cosmos”, honoring John Wheelers 90th birthday. J. D. Barrow, P. C. W. Davies, and C. L. Harper eds. Cambridge University Press (2003); arXiv:grqc/0303100. 27. G. t Hooft, In Search of the Ultimate Building Blocks, Cambridge Univ. Press 1997, p.70. 28. See, for instance, R. F. Streater and A. S. Wightman, PCT, Spin and Statistics, and all that, W. A. Benjamin, New York 1964. 29. I.E. Segal, Is the Physical Vacuum Really Lorentz-Invariant?, in Differential Geometry, Group Representations, and Quantization, J.D. Hennig, W.Lu¨cke, J. Tolar, Eds. Lecture Notes in Physics Vol. 379, Springer 1991. 30. M. Yoon and D. Huterer, Kinematic Dipole Detection With Galaxy Surveys: Forecasts And Requirements, Astrophys. J. Lett. 813 (2015) L18. 31. G. F. Smoot, Cosmic microwave background radiation anisotropies: Their discovery and utilization, Nobel Lecture, Rev. Mod. Phys. 79, 1349 (2007). 32. L. Hardy, Quantum mechanics, local realistic theories, and Lorentz-invariant realistic theories, Phys. Rev. Lett. 68 (1992) 2981. 33. A. A. Michelson and E. W. Morley, On the Relative Motion of the Earth and the
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 57
Luminiferous Ether, Am. J. Sci. 34, 333 (1887). 34. E. W. Morley and D. C. Miller, Phil. Mag. 9 (1905) 680. 35. D. C. Miller, The Ether-Drift Experiment and the Determination of the Absolute Mo-
tion of the Earth, Rev. Mod. Phys. 5, 203 (1933). 36. A. A. Michelson, et al., Conference on the Ether-Drift Experiments, Ap. J. 68 (1928)
p. 341-402. 37. K. K. Illingworth, A Repetition of the Michelson-Morley Experiment Using Kennedys
Refinement, Phys. Rev. 30, 692 (1927). 38. R. Tomaschek, About the Michelson experiment with fixed star light, Astron.
Nachrichten, 219, 301 (1923), English translation. 39. A. Piccard and E. Stahel, REALIZATION OF THE EXPERIMENT OF MICHEL-
SON IN BALLOON AND ON DRY LAND, Journ. de Physique et Le Radium IX (1928) No.2. 40. A. A. Michelson, F. G. Pease and F. Pearson,Repetition of the Michelson-Morley Experiment, Nature, 123, 88 (1929). 41. A. A. Michelson, F. G. Pease and F. Pearson, Repetition of the Michelson-Morley experiment, J. Opt. Soc. Am. 18, 181 (1929). 42. F. G. Pease, Ether-Drift Data, Publ. of the Astr. Soc. of the Pacific, XLII, 197 (1930). 43. G. Joos, Die Jenaer Wiederholung des Michelsonversuchs, Ann. d. Physik 7, 385 (1930). 44. H. Mu¨ller, S. Herrmann, C. Braxmaier, S. Schiller, A. Peters, Precision test of the isotropy of light propagation, Appl. Phys. B 77, 719 (2003). 45. H. A. Lorentz, The Theory of Electrons, Leipzig 1909, B. G. Teubner Editor. 46. J. M. Jauch and K. M. Watson, Phenomenological Quantum-Electrodynamics, Phys. Rev. 74, 950 (1948). 47. R. J. Kennedy,Simplified theory of the Michelson-Morley experiment, Phys. Rev. 47, 965 (1935). 48. J. C. Maxwell, Ether, Encyclopaedia Britannica, 9th Edition, 1878. 49. M. Born, Einsteins Theory of Relativity, Dover Publ., New York, 1962. 50. M. Consoli and E. Costanzo, Old and new ether-drift experiments: A sharp test for a preferred frame, N. Cimento B 119, 393 (2004). 51. W. M. Hicks, On the Michelson-Morley experiment relating to the drift of the aether, Phil. Mag. 3, 9 (1902). 52. M. Consoli, A. Pluchino and A. Rapisarda,Basic randomness of nature and ether-drift experiments, Chaos, Solitons and Fractals 44, 1089 (2011). 53. M. Consoli, A. Pluchino, A. Rapisarda and S. Tudisco, The vacuum as a form of turbulent fluid: motivations, experiments, implications, Physica A394, 61 (2014). 54. R. P. Feynman, R. B. Leighton and M. Sands, The Feynman Lectures on Physics, Addison Wesley Publ. Co. 1963.
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
58 Authors Names
55. L. Onsager, N. Cimento, Statistical hydrodynamics, Suppl. 6, 279 (1949). 56. G. L. Eyink and K. R. Sreenivasan, Onsager and the theory of hydrodynamic turbu-
lence, Rev. Mod. Phys. 78, 87 (2006). 57. E. T. Whittaker,A History of the Theories of Aether and Electricity, Dover Publ.,
New York 1989. 58. P. A. M. Dirac, Is there an Aether?, Nature 168, 906 (1951). 59. O. V. Troshkin, On wave properties of an incompressible turbulent fluid, Physica A
168 (1990) 881. 60. H. E. Puthoff,Linearized turbulent flow as an analog model for linearized General
Relativity, arXiv:0808.3401 [physics.gen-ph]. 61. T. D. Tsankov,Classical Electrodynamics and the Turbulent Aether Hypothesis,
Preprint February 2009, unpublished. 62. L. A. Saul, Spin Waves as Metric in a Kinetic Space-Time, Phys. Lett. A 314 (2003)
472. 63. M. Consoli, A kinetic basis for space-time symmetries, Phys. Lett. A376 (2012) 3377. 64. E. Nelson, A derivation of the Schr¨odinger Equation from Newtonian Mechanics, Phys.
Rev. 150 (1966) 1079. 65. P. Jizba and H. Kleinert, Superstatistics approach to path integral for a relativistic
particle Phys. Rev. D 82 (2010) 085016. 66. P. Jizba, F. Scardigli, Special Relativity induced by Granular Space, Eur. Phys. J. C73
(2013) 2491. 67. U. Frisch, Turbulence The Legacy of A. N. Kolmogorov, Cambridge University Press,
1995. 68. Q. Chen, E. Magoulakis, and S. Schiller,High-sensitivity crossed-resonator laser ap-
paratus for improved tests of Lorentz invariance and of space-time fluctuations, Phys. Rev. D 93 , 022003 (2016). 69. S. Rizzo, A. Rapisarda, Environmental Atmospheric Turbulence at Florence Airport, arXiv:cond-mat/0406684 [cond-mat.stat-mech]. 70. M. Nagel, S. R. Parker, E. V. Kovalchuk, P. L. Stanwix, J. G. Hartnett, E. N. Ivanov, A. Peters, M. E. Tobar, Direct terrestrial test of Lorentz symmetry in electrodynamics to 1018, Nature Comm.6, 8174 (2015). 71. J. J. Nassau and P. M. Morse,A Study of Solar Motion by Harmonic Analysis, Ap. J. 65, 73 (1927). 72. L. D. Landau and E. M. Lifshitz,Fluid Mechanics, Pergamon Press 1959. 73. J. C. H. Fung et al., Kinematic simulation of homogeneous turbulence by unsteady random Fourier modes, J. Fluid Mech. 236, 281 (1992). 74. R. S. Shankland, S. W. McCuskey, F. C. Leone, G. Kuerti, New Analysis of the Interferometer Observations of Dayton C. Miller, Rev. Mod. Phys.27, 167 (1955). 75. D. C. Miller, Comments on Dr. Georg Jooss Criticism of the Ether-Drift Experiment,
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
Instructions for typing manuscripts (papers title) 59
Phys. Rev. 45 (1934) 114. 76. L. S. Swenson Jr., The Ethereal Aether, A History of the Michelson-Morley-Miller
Aether-Drift Experiments, 1880-1930. University of Texas Press, Austin 1972. 77. Loyd S. Swenson Jr.,The Michelson-Morley-Miller Experiments before and after 1905,
Journ. for the History of Astronomy, 1, 56 (1970). 78. J. Shamir and R. Fox, A New Experimental Test of Special Relativity, N. Cim. 62B,
258 (1969). 79. M. Consoli, A. Pluchino and A. Rapisarda, Cosmic Background Radiation and ether-
drift experiments, Europhysics Lett. 113, 19001 (2016). 80. G. Joos, Note on the Repetition of the Michelson-Morley Experiment, Phys. Rev. 45,
114 (1934). 81. E. R. Farkas and W. W. Webb, Precise and millidegree stable control for fluorescense
imaging, Rev. Scient. Instr. 81, 093704 (2010). 82. Y. Zhao, D. L. Trumper, R. K. Heilmann, M. L. Schattenburg, Optimizatiom and tem-
perature mapping of an ultra-high thermal stability enviromental enclosure, Precision Engin. 34, 164 (2010). 83. I. P. Prikhodko, A. A. Trusov, A. M. Shkel, Compensation of drifts in high-Q MEMS gyroscopes using temperature self-sensing, Sensors and Actuators A 201, 517 (2013). 84. S. W. W. Manley, A novel Michelson-Morley experiment testing for anisotropic light propagation in gas without violation of local Lorentz invariance, Eur. Phys. J. Plus 138, 206 (2023). 85. R. V. Pound, Electronic frequency stabilization of microwave oscillators, Rev. Sci. Instr. 17, 490 (1946). 86. R. W. P. Drever, J. L. Hall, F. V. Kowalski, J. Hough, G. M. Ford, A. J. Munley, H. Ward, Laser phase and frequency stabilization using an optical resonator, Appl. Phys. B 31, 97 (1983). 87. E. D. Black, An introduction to Pound-Drever-Hall laser frequency stabilization, Am. J. Phys.69, 79 (2001). 88. A. Brillet and J. L. Hall, Improved Laser Test of the Isotropy of Space, Phys. Rev. Lett. 42, 549 (1979). 89. G. F. Smoot, M. V. Gorenstein, and R. A. Muller, Detection of Anisotropy in the Cosmic Blackbody Radiation, Phys. Rev. Lett.39, 898 (1977). 90. Ch. Eisele, M. Okhapkin, A. Nevsky, S. Schiller, A crossed optical cavities apparatus for a precision test of the isotropy of light propagation, Opt. Comm. 281, 1189 (2008). 91. S. Herrmann, A. Senger, K. M¨ohle, M. Nagel, E. V. Kovalchuk, A. Peters, Rotating optical cavity experiment testing Lorentz invariance at the 1017 level, Phys.Rev. D 80, 10511 (2009). 92. M. Nagel, K. Mo¨hle, K. D¨oringshoff, S. Schikora, E.V. Kovalchuk, A. Peters, Ultra-stable Cryogenic Optical Resonators For Tests Of Fundamental Physics,
February 21, 2024 2:5 WSPC/INSTRUCTION FILE Consoli˙review˙arxiv
60 Authors Names
arXiv:1308.5582[physics.optics]. 93. K. Numata, A, Kemery and J. Camp, Thermal-Noise Limit in the Frequency Stabi-
lization of Lasers with Rigid Cavities, Phys. Rev. Lett. 93, 250602 (2004). 94. M. Consoli and L. Pappalardo, Emergent gravity and ether-drift experiments, Gen.
Rel. and Grav. 42, 2585 (2010). 95. C. Barcelo, S. Liberati and M. Visser, Analog gravity from Bose-Einstein condensates,
Class. Quantum Grav. 18, 3595 (2001). 96. M. Visser, C. Barcelo and S. Liberati, Analogue models of and for gravity, Gen. Rel.
Grav. 34, 1719 (2002). 97. G. E. Volovik, Superfluid analogies of cosmological phenomena, Phys. Rep. 351, 195
(2001). 98. H. Yilmaz, New Approach to General Relativity, Phys. Rev. 111, 1417 (1958). 99. R. P. Feynman, in Superstrings: A Theory of Everything? , P. C. W. Davies and J.
Brown Eds., Cambridge University Press, 1997, pag. 201. 100. E. G. Adelberger, New tests of Einsteins equivalence principle and Newtons inverse-
square law, Class. Quantum Grav. 18, 2397 (2001). 101. R. DE. Atkinson,General relativity in Euclidean terms Proc. R. Soc. 272, 60 (1963). 102. K. Thorne, Black Holes and Time Warps: Einsteins Outrageous Legacy, W. W.
Norton and Co. Inc, New York and London, 1994, see Chapt. 11 “What is Reality?”. 103. R. J. Cook, Physical time and physical space in general relativity, Am. J. Phys. 72,
214 (2004). 104. A. S. Eddington,Space, Time and Gravitation, Cambridge University Press, 1920. 105. A. M. Volkov, A. A. Izmestev, and G. V. Skrotski, The propagation of electromag-
netic waves in a riemannian space, Sov. Phys. JETP 32, 686 (1971). 106. J. Broekaert, A Spatially-VSL Gravity Model with 1-PN limit of GRT, Found. of
Phys. 38, 409 (2008). 107. L. D. Landau and E. M. Lifshitz,The Classical Theory of Fields, Pergamon Press,
1971, p.257. 108. D. A. Jennings, R. E. Drullinger, K. M. Evenson, C. R. Pollock, J. S. Wells, The
Continuity of the Meter: The Redefinition of the Meter and the Speed of Visible Light, Journ. of Res. Nat. Bur. Stand. 92, 11 (1987). 109. M. Nagel, Design of a next-generation modern Michelson-Morley experiment, PhD Thesis, (Humboldt U., Berlin), March 2022. 110. C. La¨mmerzahl et al., OPTIS: a satellite-based test of special and general relativity, Class. Quantum Gravity 18, 2499 (2001).