Skip to content
Publicly Available Published by De Gruyter November 14, 2018

Einstein’s “Clock Hypothesis” and Mössbauer Experiments in a Rotating System

  • Alexander Kholmetskii EMAIL logo , Tolga Yarman , Ozan Yarman and Metin Arik

Abstract

An extra energy shift between emitted and received radiation on a rotating disc – next to the conventionally recognised second-order Doppler shift – has been revealed in a series of recent Mössbauer experiments, where a radioactive source is fixed at the centre and an absorber is attached to the rim of the rotating disc. This disclosure gives indication to a possible violation of the “clock hypothesis” by Einstein: i.e. the independence of the rate of a clock on its acceleration. At the moment, there seem to be two plausible interpretations of this result: (i) the deviation of the geometry of the rotating disc from that predicted by the general theory of relativity (GTR), or (ii) the existence of a specific maximal acceleration in nature, when transformation between two accelerated frames differs from the corresponding transformation of the relativity theory. We take a closer look at both ways leading to the violation of the clock hypothesis; particularly, by analysing the outcomes of recent experiments in rotating systems and by suggesting a new Mössbauer rotor experiment to determine the most feasible mechanism for testing the dependence of the rate of a clock on its acceleration.

1 Introduction

It is known that the “clock hypothesis” by Einstein (hereafter abbreviated as CH) implies the dependence of the time rate of a clock on just its velocity – thus signifying the independence of such a clock rate on its acceleration – and that this hypothesis played a fundamental role in the development of Einstein’s relativity theory (e.g. [1]). Remarkable agreement of both special relativity and general relativity with the findings from numerous astronomical observations and laboratory experiments on space-time physics supported the belief in the validity of the CH for a long time.

However, doubts with respect to the validity of the CH had been expressed in the second half of the past century where the existence of a maximal acceleration in nature, am, was posited (e.g. [2]). Under such an assumption, the time rate of a clock experiencing acceleration a becomes dependent on the ratio a/am, which therefore turns out to be at odds with the CH.

The value of the maximal acceleration estimated in [2] has the order of magnitude

(1)am=c2/lp5.5×1051m/s2,

where c = 3.0 × 108 m/s is the speed of light in vacuum, and lp = 1.62 × 10−35 m is the Planck length, which is rather huge. Thus, if the indicated value (1) is actual, any subsequent violation of the CH due to an accelerational upper bound can hardly be detected in laboratory-scale experiments, because the ratio a/am then becomes quite miniscule for any realistic situation. Furthermore, positing an accelerational upper bound would have detectable consequences (however scarce) with regard to the interpretation of some catastrophic cosmological phenomena; and this is most likely why the work in [2] did not alter the established belief in the validity of Einstein’s CH under usual circumstances.

The first indication of a possible violation of the CH in the laboratory scale became manifest at the beginning of the 21st century [3] upon our re-analysis of known past experiments that attempted to measure the Mössbauer effect in rotating systems. These were, in particular, the experiments conducted by Kündig [4] and Champeney et al. [5] – which were the most careful among the other known experiments on this subject – who had originally reported the confirmation of the standard relativistic prediction to within about 1 % measurement accuracy.

Following the prediction by the present second co-author that there should, in effect, turn up a severe violation of the CH in classical and quantum bound systems (see, e.g. [6]), our team decided to re-examine the sets of raw data fortunately presented by the experimenters in [4], [5]. We revealed that significant data processing errors were made by the aforementioned authors, and we immediately reported this disclosure [3]. It has been concluded in [3] that the observed energy shift between the resonant lines of the source and the absorber having different radial coordinates on a rotor contains a remarkable additional component next to the usual relativistic dilation of time, so much so that the determination of its origin became a vital problem.

As is known, the relative energy shift ΔE/E between the emitted and absorbed resonant radiation on a rotating disc – where a source is mounted on the rotor axis and a receiver is located at the rotor rim – can be expressed to the accuracy c−2 via the relationship

(2)ΔEE=ku2c2,

where u is the tangential velocity of the resonant absorber, and k is some coefficient which should be equal to 0.5 because of the time dilation effect of Einstein’s relativity theory. However, as we have shown in [3], the data published by Kündig [4] and Champeney et al. [5] based on the Mössbauer effect yielded, in fact, the inequality

(3)k0.6

once we re-analysed them, which points to the presence of an additional energy shift between the resonant lines of the source and the absorber. Note that the divergence in question amounts to more than 20 % in comparison with the relativistic prediction, whereas the experimental precision was reported to be about 1 %.

Our findings stimulated the realisation of modern Mössbauer experiments in rotating systems [7], [8], [9], which well confirmed the inequality (3), thus yielding

(4)k=0.66±0.03

in [7], [8], and

(5)k=0.69±0.02

in [9].

As can be seen, the value of k in (4) and (5) exceeds by 30 % the classical relativistic prediction (k = 0.5), and is much larger than the measurement uncertainty for the aforementioned experimental set-ups.

These results motivated further research into the investigation of the extra energy shift in Mössbauer rotor experiments, culminating in both experimental [10], [11] and theoretical (e.g. [12], [13], [14], [15]) approaches. In Section 2, we consider the possible violation of the CH in the view of these recent works, which proffer two different theoretical solutions to the conundrum at hand:

  1. Under the framework of a generalised version of special relativity developed by Friedman et al. (e.g. [12], [15]), the presence of a specific maximal acceleration amax in nature is assumed, whose value could be much smaller than the estimation (1) given in [2]. In such a case, the violation of the CH could be observable via the highly sensitive Mössbauer spectroscopy when it is applied to rotating systems.

  2. Within the framework of the Yarman-Arik-Kholmetskii (YARK) theory of gravity (e.g. [13], [14], [16], [17], [18], [19], [20], [21], [22], where space-time metric of a rotating disc does in fact depend on its centrifugal acceleration, this must lead to the violation of the CH as hinted by the inequality (3).

We also ought to mention at this point the recent attempts [23], [24] to explain the origin of the inequality (3) under the framework of general theory of relativity (GTR) – which, however, are erroneous, and the proposed solution is based on a misunderstanding by Corda of the basic points of the Mössbauer effect methodology as we had explained in our publications [25], [14]. Thus, the problem of the actual interpretation of the origin of the extra energy shift between emission and absorption lines in GTR remains open at the moment.

Further, we highlight the principal difference of mechanisms of possible violation of CH in generalised relativity by Friedman et al. and in the YARK theory.

The Friedman theory, being purely kinematical in its essence (see Section 2.1), implies a direct influence of acceleration on the tick of a clock, and thus its predictions with respect to the Mössbauer rotor experiments can be directly compared with the accelerator experiments with muons (e.g. [26], [27]), where the CH was tested via the measurement of the decay rate of positive and negative muons moving in circular orbits.

The application of YARK theory to the Mössbauer rotor experiments is based on the fundamental statement that in a rotating system, the tick of the clock depends on the work done to this clock against the centrifugal force when the clock is moving in the radial direction (see Section 2.2). Therefore, the related violation of CH via this mechanism should be specific for mechanically bound systems and cannot be realised in the accelerator experiments, where the magnetic Lorentz force, responsible for the circular motion of charged particles, as is known, does not make any work.

In Section 2, we analyse the data from Mössbauer rotor experiments in the light of theoretical predictions of the theory by Friedman et al. and YARK theory. We discuss the results obtained in comparison with the accelerator experiments with muons [26], [27], where the CH can be also tested. In Section 3, we suggest carrying out a new Mössbauer rotor experiment that is capable of discriminating between the predictions of generalised relativity by Friedman et al. and of YARK theory. Finally, we conclude in Section 4.

2 The Origin of the “Extra Energy Shift” between Emission and Absorption Lines in Mössbauer Rotor Experiments in the Light of Two Competing Viewpoints

2.1 The Generalised Relativity byFriedman et al.

Insight into the theory by Friedman and his collaborators can be found in [12], [15]. The unorthodox generalisation of special relativity by Friedman et al. is essentially based on the negation of the CH and the postulation of a maximal acceleration am. It is important to note that this upper bound is not derived a priori but is merely contingent upon the a posteriori inputting of an arbitrary magnitude by hand based on empirical measurements. In fact, the true measured value is the ratio a/am, where a is the acceleration achieved in a given experiment. It is known that, under laboratory conditions, extreme values of a for macroscopic objects can be reached in rotor experiments where the centrifugal acceleration a=ω2r can delineate large numbers (106–107) g for modern rotor systems, with ω being the angular rotation frequency, r the rotor radius, and g the acceleration of free fall on the surface of the earth. Therefore, Mössbauer experiments in rotating systems, which combine the feasibility of obtaining large ratios of ω2r/am with an exceptionally high sensitivity of the Mössbauer effect to the relative energy shifts of resonant lines, represent a promising tool in the search for effects violating the CH under laboratory conditions.

Hence, if one is to give credence to the approach by Friedman et al., it is possible to “reverse-engineer” the value of the actual energy shift between emission and absorption lines in a rotating system as expressed by the inequality (3) by simply relying on a maximal acceleration am with a value much lower than its original estimation in (1).

In order to evaluate the appropriate value of am, which could seemingly explain the inequality (3), we can refer to the transformation between two accelerated frames. In the generalised relativity by Friedman et al., it has the form of a usual Lorentz transformation with the replacement of the ratio v/c by a/am [12]. One can show that, in this case, the relationship between the energy of an emitted radiation E0 (for a source fixed on the rotational axis) and the energy of the absorbed radiation E (for an absorber located at the rotor edge) is given by the relation [12]

E=(1+ω2ram)(1ω2r2c2)1/2E0.

Therefore, the relative energy shift between the resonant lines of the source and the absorber takes the form

(6)ΔEE=E0EE0ω2ramω2r22c2=u2c2(12+c2ram),

where u = ωr.

Comparing (2) and (6), one can straightforwardly determine the coefficient k as the function of a maximal acceleration am:

(7)k=1/2+c2/ram.

Trying to estimate the value of maximal acceleration am on the basis of (7), Friedman et al. originally used the result of the experiment by Kündig [4], which we have re-analysed and corrected in [3], claiming that it is much more reliable than the results of modern experiments [7], [8], [9], yielding (4) and (5). As we have shown in [13], the argument by Friedman et al. against the results (4) and (5) rests on their fallacious assumption about the influence of the linear Doppler effect on the shape of the measured resonant line in a rotating system. However, one can easily demonstrate that, in the case where both the source and the absorber are fixed on a rotor, the linear Doppler effect totally vanishes [13]. At the same time, we do not look more closely at this problem at the present stage, since the difference between the rectified datum of the experiment by Kündig [4] upon our re-analysis in [3] (k = 0.596 ± 0.05) and the results (4) and (5) does not matter so much in the evaluation of the postulated maximal acceleration am. In particular, substituting in (7) the radial coordinate of the resonant absorber in the experiment by Kündig [4] (r = 9.3 cm), we obtain at k = 0.596

(8)am1.01×1019m/s2.

One can see that the value (8) is much smaller than the initial estimation of a plausible maximal acceleration as given in (1).

In order to verify this result, recently Friedman et al. carried out two Mössbauer experiments in a rotating system, where they applied resonant synchrotron radiation focused as a narrow beam on a spinning disc [10], [11].

Such experiments, where a source of resonant radiation rests in the laboratory, have the principal disadvantage in comparison with the traditional scheme (where a point-like source of resonant radiation and the resonant absorber are both fixed to a spinning disc) because of the emergence of a linear Doppler effect between the immovable synchrotron source and orbiting absorber, which leads to substantial broadening of the measured resonant line. This, in turn, results in a proportional decrease of the measurement sensitivity with respect to the energy shifts of such a line.

Besides, it is obvious that, for a synchrotron source, the rotor vibrations as a whole distort the measurement results, whereas in the traditional measurement scheme, where both a compact source and a resonant absorber are fixed on a rotor, only relative vibrations between them should be taken into account – which are anyway much smaller than the absolute rotor vibrations.

As a result, the width of the resonant line of a rotating absorber, when measured with a resonant synchrotron beam, exceeds the proper width of the line tens of times at the tangential velocities of around 100 m/s [10], [11]. In contrast, in traditional experiments with a compact resonant source fixed at the centre of the rotor along with a co-rotating resonant absorber at the rotor’s rim, the measured width of the resonant line remains comparable to its proper width even at near-sonic tangential velocities of 300 m/s because of the absence of the linear Doppler effect.

These facts explain the failure of the authors of the experiments in [10], [11] in measuring the relative energy shift between emitted and absorbed resonant lines as a function of the tangential velocity of the absorber, so much so that they could not even estimate the coefficient k in (2). Under these conditions, Friedman et al. focused their efforts on another method for the determination of possible influences of acceleration on the relative energy shift between emitted and received resonant radiation: namely, they compared the Mössbauer spectra of the resonant absorber at different angular positions [designated in [10], [11] as positions (a) and (b)], where the projection of centrifugal acceleration on the axis of the synchrotron beam has the same values but opposite signs. Hence, according to (6), the relative energy shift of the resonant lines between these positions – as tagged in [10], [11] under the designation of the “alignment shift” (AS) – is determined by the equation

(9)(ΔEE)AS=(ΔEE)(a)(ΔEE)(b)=2u2Ram=2ω2Ram.

An advantage of this method is the independence of AS on the finite distance between the axis of the synchrotron beam and the rotational axis [10], [11], so that it might seem, at first glance, that AS should be fully defined by the supposed maximal acceleration according to (9).

As is usually done in Mössbauer spectroscopy, the relative energy shift between emission and absorption lines (in our case, the relative alignment shift) can be expressed via the conditional relative velocity between the source and the absorber, corresponding to the same linear Doppler shift, i.e.

(ΔEE)AS=uASc

with

(10)uAS=2ω2Rcam.

The convenience of definition (10) is the fact that uAS can be directly obtained from the measured Mössbauer spectra in positions (a) and (b) plotted in the velocity scale.

In the last experiment by Friedman et al. [11], which differs from their earlier experiment [10] by some methodological improvements, the authors actually observed a quadratic dependence of uAS on the angular frequency ω, as predicted by (9), up to 200 rev/s. In particular, at the rotor radius of R = 5.0 cm, which was used in the experiments in [10], [11], and the rotational frequency of 200 rev/s, the estimated value of AS came out to be uAS = 0.41 mm/s, with the relative measurement uncertainty being less than 10 %. Concurrently, Friedman et al. noticed that the contribution of any effects related to the non-random character of rotor vibrations had a much smaller value. Thus, Friedman and his collaborators concluded that the observed AS betokens the violation of the CH as a result of the existence of a maximal acceleration in nature, which can be estimated from (9) as

(11)am=2ω2RcuAS.

Hence, at ω = 2π × 200 s−1, R = 0.05 m, and via the measured value uAS = 0.41 mm/s, we obtain [11]

(12)am=1.2×1017m/s2.

The value in (12) is two orders of magnitude smaller than the maximal acceleration (8) derived earlier by Friedman et al., which we have rectified based on the experiment by Kündig [3].

One should mention another estimation of this alleged maximal acceleration via the Mössbauer effect, which is shifted all the way up to at least

(13)am>1.5×1021m/s2

as obtained in [28] in the temperature-shift measurements with Mössbauer spectroscopy of 67Zn, which is characterised by the extremely narrow width of its resonant lines.

Nevertheless, in the opinion of Friedman et al., the three different results (8), (12), and (13) still do not create any doubts as to the correctness of their methodological approach, but rather indicate that am is not a universal quantity, as was also contended in [2], and, according to these authors, can be much smaller than that originally postulated in (1).

At the same time, there appears strictly no way to attribute any physical meaning to such conflicting values for a supposed accelerational upper bound in experiments using the same tool – Mössbauer spectroscopy – under similar conditions and rotor parameters.

What is more, in our recent paper [29] we have shown that the latest result by Friedman et al. (12) is far from convincing, and can be explained by the fluctuation δr of the position of the rotational axis, which is always present for any real rotor system. In particular, as is estimated in [29], the value of uAS at the level 0.41 mm/s (which is used in the derivation of (12)) is observed at δr ≃ 0.3 μm, which cannot be well controlled in the experiments [10], [11].

Next, we emphasise that the estimations of possible values of maximal acceleration by Friedman et al. [(8) and (12)] are anyway much smaller than the limited values of maximal acceleration evaluated via the experiments on accelerators of charged particles. First of all, we refer to the experiment [26] aimed at measuring the lifetime for negative and positive muons. In this experiment, the muons with the Lorentz factor 29.3 rotated on a circular orbit with the diameter of 14 m in the presence of a magnetic field of suitable magnitude, which was orthogonal to the rotating plane. In this case, the centripetal acceleration of muons due to the magnetic Lorentz force was about a = 1.3 × 1016 m/s2. Further, taking into account that the relative measurement uncertainly of the proper lifetime of muons was about 2 × 10−4, one can estimate a possible maximal acceleration (which could affect the measured lifetime of muons beyond the indicated uncertainty):

am7.0×1019m/s2.

We add that the analysis of the muon experiments on accelerators performed in [27] gives even stronger limitation with respect to the influence of acceleration on the particle’s lifetime: with a centripetal acceleration of muon near 1018 m/s2, the estimated deviation from the behaviour of an ideal clock does not exceed 10−25.

This and other results presented above create serious doubts about the validity of the generalised version of relativistic kinematics by Friedman et al., with the value of maximal acceleration of the order 1017–1019 m/s2. On the contrary, the accelerator experiments with positive and negative muons, orbiting in the presence of a suitable magnetic field, rather indicate the independence of the clock rate on its acceleration, as in effect originally assumed by Einstein.

At the same time, as we have already mentioned and will detail below, we anticipate that for bound systems of particles (where composing of such systems requires delivering a finite work to its constituents), there appears one more mechanism of violation of CH, which exhibits itself in the Mössbauer rotor experiments as the extra energy shift between emission and absorption resonant lines on a rotating disc. Here, it is important to stress that the accelerated charged particles, orbiting in the presence of the magnetic field, do not represent bound particles in the meaning defined above, because the magnetic Lorentz force does not constitute any work. Therefore, the indicated way of possible violation of CH for mechanically bound systems (like resonant source and resonant absorber fixed on a spinning disc) cannot be realised in the experiments where charged particles are accelerated in a magnetic field.

A consistent way to describe this mechanism of violation of CH is given within the YARK gravitation theory, given in the following.

2.2 Mössbauer Rotor Experiments from the Point of View of YARK Theory

An introductory framework of YARK theory of gravity has been drawn, e.g. in [16], [17], [18], [19], [20], [21], [22]. Here, we present some principal points appertaining to our approach.

The main motivating factor for the development of YARK theory was to achieve a symbiosis between gravitation and quantum mechanics, as well as the elimination of any difficulties with respect to the definition of gravitational energy. One can realise that the necessary condition for meeting the aforementioned criteria is the denial of a purely metric approach as had been applied in GTR and in extended theories of gravity.

Needless to say, it is known that purely dynamic theories of gravity (where the notion of “force” is defined independently of the space-time metric; like, e.g. in post-Newtonian theories) are incompatible with available experimental facts.

These difficulties are overcome in YARK theory, which can be classified neither as a purely metric theory nor as a purely dynamic theory: it rather successfully combines the properties of both of them. In particular, in YARK, the force exerted on a test particle in the presence of gravity represents a non-vanishing entry in any frame of observation (just like in dynamic theories); while, at the same time, the origin of this force is explained via the variation of the metric of four-space, which makes YARK similar to metric theories in its many applications.

The combination of dynamic and metric properties is achieved via the principal postulate of YARK theory, which defines the overall energy of an object moving in the presence of gravity as [16], [17], [18]

(14)E=γm0c2(1+EB/m0c2).

Here, m0 is the rest mass of the object in the absence of gravity, γ is its Lorentz factor, and EB is the static binding energy of the object at the given location (i.e. this is the work delivered to the object – assumed initially at rest at the given location – in order to carry it quasi-statically to an infinity away).

In fact, (14) states that the rest mass of the object is not of a constant magnitude but is rather altered by a gravitational environment by the value EB/c2.

Further, because of the close quantum mechanical relationship between the quantities “mass”, “energy”, “frequency”, “time”, and “size”, the variation of the rest mass of a test particle by the static binding energy affects the time rate for the particle and induces a corresponding transformation of spatial intervals in the presence of gravity [16], [17], [18]. Thus, the gravitational field alters the metric of space-time, so that the expression for the space-time interval in spherical coordinates and in the radially symmetric case reads as

(15)ds2=e2α(c2dt2dr2(1+α)2r2(dθ2+dϕ2sin2θ)),

where the factor α=GM/rc2, with G being the gravitational constant and r the distance from the centre of the immobile host body M to the point of observation. Here, r is the radial coordinate, θ is the polar coordinate, and φ is the azimuthal coordinate, and all of them are defined with respect to a distant observer located outside of gravity.

Further, the action of a particle in YARK theory, defined in a usual way as S=m0cds, yields the following Lagrangian in the metric (15) [19], [20], [21], [22]:

(16)L=m0eαc2/γ0,

where γ0=(1v02/c2)1/2, with v0 = dl/dτ being the velocity of the particle measured by a local observer, dl the path element, and τ the proper time. Hence, the force F acting on the particle, its momentum p, and its energy E are defined by the corresponding relationships:

(17)F=Lr=m0c2γ0(eα)r=GMm0eαrγ0r3,
(18)p=Lv=γ0m0eαv0,
(19)E=pvL=γ0m0eαc2.

Comparing now (19) with the known expression for the energy of the test particle in GTR [30]

(20)EGTR=γm0c212α,

we see that the terms describing the effect of gravity in (19) and (20) coincide with each other to the accuracy c−3; i.e. 12α1αeα. Therefore, GTR and YARK do converge in the limit of a weak gravitational field, and both provide a successful explanation of the classical astrophysical observations (e.g. gravitational redshift, precession of the perihelion of Mercury [31], [32], gravitational lensing [33], and Shapiro delay [34]).

In addition, YARK theory also achieved considerable successes in explaining modern observations where the weak relativistic limit is abandoned (e.g. derivation of the alternating sign for the accelerated expansion of the Universe without the need to involve a notion of “dark energy”, presentation of the Hubble constant in an analytical form, and elimination of the “information paradox” for black holes of the YARK type [18], [19]).

In the available publications about YARK theory, we still did not consider the topical problem of rotation of galaxies and its connection to the existence of dark matter, where significant modifications of GTR are possible (see, e.g. [35], [36]). This will be done in a separate contribution, but now we address the most recent experimental results in space-time physics, which, among other things, can also be considered as crucial tests of YARK theory. In this regard, it is important to emphasise that YARK theory remains the only alternative to GTR, which provides an adequate account of the GW150914 and GW151226 signals announced by LIGO beyond the hypothesis about gravitational waves [22]. In addition, the recent indications with respect to the practically null bending of high-energy γ-quanta under Earth’s gravity [37] find a successful explanation in the framework of YARK theory, too [21]. Last, but not least, the significant achievement of YARK theory remains its quantitative agreement with the experimental results (3) and (4) obtained in modern Mössbauer rotor experiments.

A possible divergence between the predictions of GTR and YARK theory at a measurable level in this kind of experiments stems from the fact that, for an accelerated motion of particles in the absence of gravity, one cannot separate (19) and (20) into particular contributions describing the effects of inertial and non-inertial motion, so that for such a motion, the specific terms distinguishing GTR and YARK from each other might emerge already in the order (u/c), where u is a typical velocity characterising the motion of a system in question. For example, under rotational motion, the factor α entering into (17)–(19) is defined as α=acentrifugalr/c2=u2/c2. The substitution of this value into (19) and (20) reveals their disparity already in the order (u/c)2, which corresponds to the accuracy in the presentation of (2). Hence, the coefficient k calculated in this latter equation can be different in GTR and in YARK even though both theories agree well with respect to numerous astrophysical observations dealing with commonly observed gravitational effects.

We emphasise the differing predictions made by GTR and YARK theory with respect to the coefficient k in (2) as indicative of how the respective geometries framed by both theories in a rotating disc should also be different.

The spatial geometry of the rotating disc in GTR is well known (see, e.g. [30]), and it yields k = 1/2 in (2), which, for a laboratory observer, is explained by the second-order Doppler effect and leaves no room for the violation of the clock hypothesis by Einstein.

The spatial geometry of the rotating disc in YARK theory is defined to the accuracy c−2 by the equation [13]

(21a)dxL=dx0,
(21b)dyL=dy0/(1u2/2c2),
(21c)dzL=dz0(1u2/2c2)

at the time moment when the tangential velocity u in the considered point of the disc r is collinear to the x-axis.

The physical meaning of (21a–c) has been disclosed in [13]. Because of centrifugal acceleration, the spatial intervals are stretched homogeneously in the x, y, and z directions by the conformal factor (1u2/2c2), in YARK theory. However, the Lorentz contraction of scale along the tangential velocity (directed along the x-axis at the considered time moment) exactly counterbalances the stretching effect along the x-direction, so that the x-coordinate remains unchanged within the adopted accuracy of calculation, c−2. We notice that this eliminates the Ehrenfest paradox. Similarly, the increase of the mass of the rotating particle by the Lorentz factor is counterbalanced by the decrease of this mass by (1u2/2c2) times because of the acceleration effect, so that m = m0 up to the adopted accuracy of the calculations, c−2 [13].

At this stage, it is important to emphasise the full compatibility of YARK theory with quantum mechanics the way we had substantiated it in [16], [17], [18], [38]. As pointed out in [13], this circumstance is profoundly crucial in the analysis of the Mössbauer effect where a resonant nucleus is to be considered analogous to a quantum particle localised in a crystal cell. On that account, the energy levels of such a nucleus should be defined via the Dirac equation in the spatial metric (21a–c), where the crystal cell can be modelled as a three-dimensional potential hole with infinite walls. As had been pointed out in [13], when we calculate the difference of the energy levels between a co-rotating resonant nuclear source and a resonant absorber characterised by different radial coordinates, it is sufficient to replace the Dirac equation with the Schrödinger equation, alongside the replacement of the variables according to (21a–c).

Thus, presenting a resonant nucleus in a crystal cell as a particle in a box with sizes a, b, and d along the respective coordinate axes, we obtain the solution of the Schrödinger equation in the form

(22)E=nx2h28m0a2+ny2h2(1u2/c2)8m0b2+nz2h2(1u2/c2)8m0d2,

where nx, ny, and nz are the corresponding principal quantum numbers [13].

Therefore, for a resonant absorber on the rotor rim, and for a resonant source on a rotating axis (where u = 0), we obtain

(23)E(absorber)=nx2h28m0a2+ny2h2(1u2/c2)8m0b2+nz2h2(1u2/c2)8m0d2,
(24)E(source)=nx2h28m0a2+ny2h28m0b2+nz2h28m0d2.

So, taking for simplicity a = b = d (which is actually the case for the majority of resonant absorbers used in the known Mössbauer rotor experiments), we derive the relative energy shift between emission and absorption resonant lines as

(25)E(absorber)E(source)E(source)=2u23c2.

Hence, (25) yields

(26)k=2/3.

Thus, the prediction of YARK theory (26) perfectly agrees with the recent measurement data (4) and (5), and is also in conformity with the Kündig [4] and Champeney et al. [5] outcome (3) the way we corrected them in [3].

One may add that the coefficient k in (26) can be presented as the sum k=12+16, where the term 1/2 stands for the usual transverse Doppler effect and the term 1/6 emerges as a result of the effect of acceleration as seen by a laboratory observer. Thus, (26) exhibits the violation of the CH as the specific effect for mechanically bound systems (in our case, resonant source and resonant absorber fixed on a rotating disc) predicted and explained by YARK theory, beyond the hypothesis about a possible existence of a maximal acceleration in nature.

3 Discussion

At the moment we have two plausible alternative hypotheses with respect to the violation of the CH:

  1. The novel framework of YARK theory of gravity, where the violation of the CH happens as a result of the dependence of a space-time metric of the rotating disc on the centrifugal acceleration of its points.

  2. The framework of the generalised relativity by Friedman et al., where the violation of the CH happens because of the existence of a suppositional maximal acceleration in nature am, so that the time rate of any clock depends on the ratio a/am, with a being its acceleration.

We have seen in Section 2 that the particular predictions of both of the mentioned theories with respect to the outcomes of Mössbauer rotor experiments are different:

  1. In YARK theory, the coefficient k in (2), which defines the relative energy shift between emitted and absorbed resonant radiation to the accuracy c−2, is always equal to 2/3 regardless of the radial coordinates of the radioactive source and the absorber and their centrifugal accelerations.

  2. In the generalised relativity by Friedman et al., the coefficient k in (2) depends on the centrifugal acceleration of an orbiting resonant absorber a = ω2r through the ratio (a/am), which leads to (7). This presumed maximal acceleration does not point to any universal upper bound, as evidenced by conflicting estimations of its value in (8), (12), and (13), as well as in the accelerator experiments with muons [26], [27].

Nevertheless, let us objectively state that these observations do not mean to abolish the hypothesis about the possible existence of a maximal acceleration in nature, even perhaps with a value much smaller than the estimation (1). Thus, if such an accelerational upper bound exists and remains at a reasonable magnitude, it is conceivable to reveal its presence via the Mössbauer effect in a rotating system.

In this respect, the unambiguous determination of the origin of the inequality (3) – i.e. either YARK’s prediction based on a specific dependence of the geometry of a rotating disc on its centrifugal acceleration or the existence of a maximal acceleration in nature as intimated by the generalised relativity by Friedman et al. – remains topical.

In order to resolve this issue, we suggest carrying out a new Mössbauer rotor experiment where both the source and the absorber are fixed on a rotor, and the radial coordinate of the source rs represents a variable parameter under the fixed radial position of the absorber r at the rotor rim, thus with rs < r. Under these conditions, the result of this experiment will allow us to distinguish the hypothesis about the possible existence of a limited acceleration in nature from the prediction (26) of YARK theory.

Indeed, if we consider first the case where the source of a resonant radiation is located on the rotational axis (rs = 0), YARK theory yields

(27)ΔEE=23u2c2=23ω2r2c2.

On the other hand, this result can be interpreted in terms of the hypothesis about the maximal acceleration [see (7)] as the equality

(28)am=6c2/r.

Therefore, at a finite radial coordinate for the source rs, we will obtain in the case of the validity of the YARK’s interpretation of (27)

(29)ΔEE=23ω2c2(r2rs2),

whereas within the framework of the hypothesis about maximal acceleration, we shall derive according to (7) and (28)

(30)ΔEE=12ω2c2(r2rs2)ω2am(rrs)=12ω2c2(r2rs2)ω26c2(r2rrs).

Thus, at rs ≠ 0, (29) and (30) yield different results. For example, at rs = r/2, (29) gives

(31)ΔEE=12ω2r2c2,

which is a YARK prediction, whereas the updated (30) yields

(32)ΔEE=1124ω2r2c2,

akin to the prediction of the generalised relativity by Friedman et al.

The relative difference between (31) and (32) is more than 4 %, and at the measurement uncertainty of less than 1 % – which is achievable in the proposed Mössbauer rotor experiment – it will be reliably detected. Thus, the outcome of the proposed experiment will shed light on the origin of the inequality k > 0.6, with the realisation of a tangible mechanism explaining the violation of the CH.

4 Conclusion

Results of modern experiments that have measured the Mössbauer effect in rotating systems confirm the inequality (3) and indicate the violation of the clock hypothesis by Einstein.

Nowadays two alternative explanations for the observed violation of the CH exist:

  1. The explanation under the framework of YARK theory, where the coefficient k in (2) should be exactly equal to 2/3 because of the specific dependence of the spatial geometry of a rotating disc on its centrifugal acceleration. Here we stress a perfect agreement between the stated prediction k = 2/3 of YARK theory and the outcomes of the recent experiments (4) and (5).

  2. The explanation within the framework of the generalised relativity by Friedman et al., where the inequality (3) is justified by the existence of a maximal (albeit arbitrarily varying) acceleration am in nature, which affects the coefficient k via (7). At the same time, if this mechanism is actually true, then the range of the values of maximal acceleration inducing measurable variations of the coefficient k in (7) must be many orders of magnitude smaller than its initial estimation (1) based on the fundamental quantities of light speed and Planck’s constant.

Concurrently, we notice that the sensitivity of modern Mössbauer rotor experiments using synchrotron radiation [10], [11] is still inadequate to carry out a direct measurement of the coefficient k in (2). The attempt by the authors of [11] to measure instead the AS between two angular positions of orbiting resonant absorber – characterised by equal in value but opposite in sign centrifugal acceleration – and their claim that the non-vanishing AS indicates the influence of acceleration on the tick of orbiting clock are questioned in our recent paper [29], where we disclosed non-accounted systematic errors in the experiment [11].

Nevertheless, the failure of the synchrotron Mössbauer experiments performed to date to check the hypothesis about the possible existence of a maximal acceleration in nature [10], [11] still leaves room for a test of this hypothesis via Mössbauer rotor experiments with usual resonant sources spinning on the rotor axis. The traditional technique is, at the moment, much more sensitive to the energy shift of resonant lines in comparison with the case of using a synchrotron source. In the present paper, we showed that the measurement of the Mössbauer effect in a rotating system with the fixed radial coordinate of the resonant absorber located at the rotor rim and the variable coordinate of a resonant source is sufficient to discriminate between the predictions of both theories (YARK and the generalised relativity by Friedman and his collaborators). Thus, the experiment proposed in Section 3 can shed more light on the actual mechanism behind the empirically substantiated violation of the CH.

References

[1] A. Einstein, The Meaning of Relativity, Princeton University Press, Princeton 1953.Search in Google Scholar

[2] E. R. Caianiello, Lett. Nuovo Cim. 32, 65 (1981).10.1007/BF02745135Search in Google Scholar

[3] A. L. Kholmetskii, T. Yarman, and O. V. Missevitch, Phys. Scr. 77, 035302 (2008).10.1088/0031-8949/77/03/035302Search in Google Scholar

[4] W. Kündig, Phys. Rev. 129, 2371 (1963).10.1103/PhysRev.129.2371Search in Google Scholar

[5] D. C. Champeney, G. R. Isaak, and A. M. Khan, Proc. Phys. Soc. 85, 583 (1965).10.1088/0370-1328/85/3/317Search in Google Scholar

[6] T. Yarman, Found. Phys. Lett. 19, 675 (2006).10.1007/s10702-006-1057-7Search in Google Scholar

[7] A. L. Kholmetskii, T. Yarman, O. V. Missevitch, and B. I. Rogozev, Phys. Scr. 79, 065007 (2009).10.1088/0031-8949/79/06/065007Search in Google Scholar

[8] A. L. Kholmetskii, T. Yarman, and O. V. Missevitch, Int. J. Phys. Sci. 6, 84 (2011).Search in Google Scholar

[9] T. Yarman, A. L. Kholmetskii, M. Arik, B. Akkus, Y. Öktem, et al., Can. J. Phys. 94, 780 (2016).10.1139/cjp-2015-0063Search in Google Scholar

[10] Y. Friedman, I. Nowik, I. Felner, J. M. Steiner, E. Yudkin, et al. EPL 114, 50010 (2016).10.1209/0295-5075/114/50010Search in Google Scholar

[11] Y. Friedman, I. Nowik, I. Felner, J. M. Steiner, E. Yudkin, et al. J. Synchrotron Radiat. 24, 661 (2017).10.1107/S1600577517002405Search in Google Scholar PubMed

[12] Y. Friedman and Y. Gofman, Phys. Scr. 82, 015004 (2010).10.1088/0031-8949/82/01/015004Search in Google Scholar

[13] T. Yarman, A. L. Kholmetskii, and M. Arik, Eur. Phys. J. Plus 130, 191 (2015).10.1140/epjp/i2015-15191-4Search in Google Scholar

[14] A. L. Kholmetskii, T. Yarman, O. Yarman, and M. Arik, Ann. Phys. 374, 247 (2016).10.1016/j.aop.2016.08.016Search in Google Scholar

[15] Y. Friedman, Ann. Phys. 523, 408 (2011).10.1002/andp.201000135Search in Google Scholar

[16] T. Yarman, Ann. Fond. de Broglie 29, 459 (2004).10.1016/j.ijhydene.2004.02.006Search in Google Scholar

[17] T. Yarman, Int. J. Theor. Phys. 48, 2235 (2009).10.1007/s10773-009-0005-2Search in Google Scholar

[18] T. Yarman and A. L. Kholmetskii, Eur. Phys. J. Plus 128, 8 (2013).10.1140/epjp/i2013-13008-2Search in Google Scholar

[19] T. Yarman, A. L. Kholmetskii, M. Arik, and O. Yarman, Can. J. Phys. 94, 271 (2016).10.1139/cjp-2015-0689Search in Google Scholar

[20] T. Yarman, A. L. Kholmetskii, M. Arik, and O. Yarman, Can. J. Phys. 94, 558 (2016).10.1139/cjp-2016-0059Search in Google Scholar

[21] M. Arik, T. Yarman, A. L. Kholmetskii, and O. Yarman, Can. J. Phys. 94, 616 (2016).10.1139/cjp-2015-0291Search in Google Scholar

[22] T. Yarman, A. L. Kholmetskii, M. Arik, and O. Yarman, Can. J. Phys. 95, 963 (2017).10.1139/cjp-2016-0699Search in Google Scholar

[23] C. Corda, Ann. Phys. 355, 360 (2015).10.1016/j.aop.2015.02.021Search in Google Scholar

[24] C. Corda, Ann. Phys. 368, 258 (2016).10.1016/j.aop.2016.02.011Search in Google Scholar

[25] A. L. Kholmetskii, T. Yarman, and M. Arik, Ann. Phys. 363, 556 (2015).10.1016/j.aop.2015.09.007Search in Google Scholar

[26] J. Bailey, K. Borer, F. Combley, H. Drumm, F. Krienen, et al. Nature 268, 301 (1977).10.1038/268301a0Search in Google Scholar

[27] A. M. Eisele, Helvetica Phys. Acta 60, 1024 (1987).Search in Google Scholar

[28] W. Potzel, Hyperfine Interact. 237, 38 (2016).10.1007/s10751-016-1212-xSearch in Google Scholar

[29] A. L. Kholmetskii, T. Yarman, O. Yarman, and M. Arik, Eur. Phys. J. Plus 133, 261 (2018).10.1140/epjp/i2018-12089-7Search in Google Scholar

[30] L. D. Landau and E. M. Lifshitz, The Classical Theory of Fields, Butterworth & Heinemann, Oxford 1999.Search in Google Scholar

[31] T. Yarman, Int. J. Phys. Sci. 5, 2679 (2010).Search in Google Scholar

[32] T. Yarman, Int. J. Phys. Sci. 6, 2117 (2011).Search in Google Scholar

[33] T. Yarman, A. L. Kholmetskii, M. Arik, and O. Yarman, Phys. Essays 27, 558 (2014).10.4006/0836-1398-27.4.558Search in Google Scholar

[34] T. Yarman, Superluminal Interaction as the Basis of Quantum Mechanics: A Whole New Unification of Micro and Macro Worlds, Lambert Academic Publishing, Saarbrucken 2011.Search in Google Scholar

[35] S. S. McGaugh, F. Lelli, and J. M. Schombert, Phys. Rev. Lett. 117, 201101 (2016).10.1103/PhysRevLett.117.201101Search in Google Scholar PubMed

[36] S. Hossenfelder, Phys. Rev. D95, 124018 (2017).10.1103/PhysRevD.95.124018Search in Google Scholar

[37] V. Gharibyan, Available at: http://arxiv.org/pdf/1401.3720.pdf [Accessed 12 July, 2014].Search in Google Scholar

[38] T. Yarman, The Quantum Mechanical Framework Behind the End Results of the General Theory of Relativity: Matter Is Built on a Matter Architecture, Nova Publishers, New York 2010.Search in Google Scholar

Received: 2018-07-23
Accepted: 2018-10-22
Published Online: 2018-11-14
Published in Print: 2019-01-28

©2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 12.6.2024 from https://www.degruyter.com/document/doi/10.1515/zna-2018-0354/html
Scroll to top button