Skip to main content
Log in

Evidence for an Effective Gravitational Constant from the Laboratory Measurements

  • Published:
Gravitation and Cosmology Aims and scope Submit manuscript

Abstract

We consider, within the framework of an improved \(5D\) Kaluza-Klein theory (KK\(\psi\)), the set of twenty most precise values ever published of Earth-based gravitational constant measurements. We take into account the possibility of a jump discontinuity into the geomagnetic potential in connection with the spin crossover of the iron (III) complex ions migrating from the Earth’s outer core into the deep Earth mantle as well as the spin crossover of the copper (III) complex ions migrating downwards within either the Earth’s crust or deep in the Earth’s mantle. Moreover, it turns out that the boundary conditions assigned to both internal, \(\phi\), and external, \(\psi\), scalar fields are sensitive to the aforementioned jump discontinuity too. Again, the fits to the experimental data are found in good agreement with the predictions of the linearized KK\(\psi\) theory. An estimate around the vacuum expectation values \(\psi=v\) and \(\phi=1\) of the second derivatives \({\partial^{2}f_{\textrm{EM}}}/{\partial\phi^{2}}(v,1)\), \({\partial^{2}f_{\textrm{EM}}}/{\partial\phi\partial\psi}(v,1)\) and \({\partial^{2}f_{\textrm{EM}}}/{\partial\psi^{2}}(v,1)\) of the coupling function, \(f_{\textrm{EM}}(\psi,\phi)\), of the electromagnetic field (EM) with \(\psi\) and \(\phi\), is provided for the first time, thereby enabling the determination of the product \(v\times f_{\textrm{EM}}(v,\phi)\) and \(v\times f_{\textrm{EM}}(\psi,1)\) in the second-order approximation from the experimental data.

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

Access this article

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

Instant access to the full article PDF.

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

Similar content being viewed by others

Notes

  1. e.g., two simple pendulums with separation determined by a laser interferometer, a torsion balance-pendulum in the time-of-swing mode, a torsion balance pendulum using the angular-acceleration feedback, a torsion strip balance in the Cavendish mode, a torsion strip balance in electrostatic servo mode, a vertical atomic interferometer, a beam-balance (gravitational force of a large mercury mass on two copper cylinders with a modified commercial mass comparator).

  2. By comparing the gravitational constant values between the two groups that we have distinguished in our study.

  3. Two equations can be inferred: \(\Delta V=0\) since \(\textrm{div}\textrm{curl}\vec{A}=0\) and \(\Delta\vec{A}=\vec{0}\) since \(\textrm{curl}\vec{\nabla}V=\vec{0}\) and on account of the Coulomb gauge \(\textrm{div}\vec{A}\)=0. Besides, let us notice that \(A^{2}+V^{2}=[{\mu_{0}M}/(4\pi r^{2})]^{2}\)

  4. This particularity leads us to conjecture that there might be a hidden copper deposit sufficiently rich in copper (III) coordination compounds either in the Earth’s crust or deep in the Earth’s mantle at the longitude and latitude of the BIPM. There, the spin crossover of the copper (III) complex ion should occur, unlike the other laboratories mentioned in this study. Recently, large quantities of copper, lead and zinc have been unveiled within 200 kilometres of the transition between thick and thin lithosphere [31] and possibly in Earth’s deep mantle [32].

  5. Let us notice that this force resembles the Poynting-Robertson light drag, \(\vec{F}={-}({P}/c^{2})\vec{v}\), where

    $$P=\frac{1}{2}mc^{2}\left|\frac{dN}{dt}\right|=\frac{1}{2}|V_{0}|cI$$

    replaces the incident EM power. Also, as one can notice, this power \(P\) is akin to the Peltier heat per unit time generated at the junction of two conductors A and B whose Peltier coefficients would satisfy \(q(\Pi_{A}-\Pi_{B})=\frac{1}{2}mc^{2}\) when the current, \(I\), flows from A to B. By arranging one obtains

    $$\vec{F}=-\frac{1}{2}\frac{|\dot{N}|}{N}\vec{p}={-}\frac{1}{2}\frac{mI}{|q|}\vec{v}={-}\frac{1}{2}|V_{0}|I\frac{\vec{v}}{c},$$

    where

    $$\vec{p}=M\vec{v},\quad M=Nm,\quad\dot{N}=\frac{dN}{dt},\quad I=|q\dot{N}|.$$
  6. The work-energy theorem states that \(\Delta E_{c}=W\). Now, the change in the kinetic energy reads \(\Delta E_{c}=\frac{1}{2}Mv^{2}\), and the infinitesimal work

    $$\delta W=\pm\frac{1}{2}\frac{V_{0}}{c}v^{2}dQ\quad\Rightarrow\quad W=\pm\frac{1}{2}\frac{QV_{0}}{c}v^{2}.$$

    Thus one obtains

    $$\frac{1}{2}Mv^{2}=\pm\frac{1}{2}\frac{QV_{0}}{c}v^{2}\quad\Rightarrow\quad V_{0}=\pm\frac{Mc}{Q}=\pm\frac{mc}{q}.$$

REFERENCES

  1. J. Luo and Z. K. Hu, Class. Quantum Grav. 17, 2351 (2000).

    Article  ADS  Google Scholar 

  2. V. N. Melnikov, Arxiv prepint gr-qc/9903110.

  3. J.P Mbelek and M. Lachièze-Rey, Arxiv prepint gr-qc/0012086.

  4. J.P. Mbelek and M. Lachièze-Rey, Grav. Cosmol. 8, 331 (2002).

    ADS  Google Scholar 

  5. J.P. Mbelek, Grav. Cosmol. 25, 250 (2019).

    Article  ADS  MathSciNet  Google Scholar 

  6. E. Thébault et al., Earth planet and space 67, 79 (2015).

  7. J.P. Mbelek, IJMPA 35, 2040027 (2020).

    Article  ADS  Google Scholar 

  8. Z. K. Hu et al., Phys. Rev. D 71, 127505 (2005);

    Article  ADS  Google Scholar 

  9. Z. K. Hu et al., Phys. Rev. D 71, 127505 (2005); J. Luo et al., Phys. Rev. D 59, 042001 (1999).

    Article  ADS  Google Scholar 

  10. H. V. Parks and J. E. Faller, Phys. Rev. Letters 105, 110801 (2010);

    Article  ADS  Google Scholar 

  11. H. V. Parks and J. E. Faller, Phys. Rev. Letters 105, 110801 (2010); Erratum: H. V. Parks and J. E. Faller, Phys. Rev. Letters 122, 199901 (2019).

  12. Q. Li et al., Nature 560, 582 (2018).

    Article  ADS  Google Scholar 

  13. T. J. Quinn, C. C. Speake, S. J. Richman, R. S. Davis, and A. A. Picard, Phys. Rev. Letters 87, 111101 (2001).

    Article  ADS  Google Scholar 

  14. T. Quinn, C. Speake, H. Parks, and R. Davis, Phil. Trans. R. Soc. A 372, 20140032 (2014);

    Article  ADS  Google Scholar 

  15. T. Quinn, C. Speake, H. Parks, and R. Davis, Phil. Trans. R. Soc. A 372, 20140032 (2014); T. Quinn, H. Parks, C. Speake, and R. Davis, Phys. Rev. Letters 111, 101102 (2013).

    Article  ADS  Google Scholar 

  16. G. G. Luther and W. Towler, Phys. Rev. Letters 48, 121 (1982).

    Article  ADS  Google Scholar 

  17. O. V. Karagioz and V. P. Izmailov, Measurement techniques 39, 979 (1996).

    Article  Google Scholar 

  18. G. Rosi et al., Nature 510, 518 (2014).

    Article  ADS  Google Scholar 

  19. J. Luo et al., Phys. Rev. Letters 102, 240801 (2009).

    Article  ADS  Google Scholar 

  20. S. Schlamminger, E. Holzschuh and W. Kündig, Phys. Rev. Letters 89, 161102 (2002).

    Article  ADS  Google Scholar 

  21. S. Schlamminger et al., Phys. Rev. D 74, 082001 (2006).

    Article  ADS  Google Scholar 

  22. T. R. Armstrong and M. P. Fitzgerald, Phys. Rev. Letters 91, 201101 (2003).

    Article  ADS  Google Scholar 

  23. S. Schlamminger, J. H. Gundlach, and R. D. Newman, Phys. Rev. D 91, 121101 (2015);

    Article  ADS  Google Scholar 

  24. S. Schlamminger, J. H. Gundlach, and R. D. Newman, Phys. Rev. D 91, 121101 (2015); J. H. Gundlach and S. M. Merkowitz, Phys. Rev. Letters 85, 2869 (2000).

    Article  ADS  Google Scholar 

  25. M. U. Sagitov et al., Dok. Acad. Nauk SSSR 245, 567 (1977).

    ADS  Google Scholar 

  26. U. Kleinevoss, Bestimmung der Newtonschen Gravitationskonstanten G, PhD thesis dissertation (Universität Wuppertal, Wuppertal, Germany, 2002).

  27. C. H. Bagley and G. G. Luther, Phys. Rev. Letters 78, 3047 (1997).

    Article  ADS  Google Scholar 

  28. R. Newman et al., Phil. Trans. R. Soc. A 372, 20140025 (2014).

    Article  ADS  Google Scholar 

  29. V. V. Zelentsov, Russian Journal of Coordination Chemistry 29, 425 (2003).

    Article  Google Scholar 

  30. Y. Ide et al., Dalton Trans. 46, 242 (2017).

    Article  Google Scholar 

  31. S. Shankar et al., Nature Communications 9, 4750 (2018).

    Article  ADS  Google Scholar 

  32. C. Krebs et al., Angewandte Chemie international Edition 38, 359 (1999).

    Article  ADS  Google Scholar 

  33. B.S. Manhas et al., Synthesis and Reactivity in Inorganic and Metal-Organic Chemistry 29, 1009 (1999).

    Article  Google Scholar 

  34. J. Meija et al., Pure and Applied Chemistry 88, 265 (2016).

    Article  Google Scholar 

  35. M. J. Hoggard et al., Nature Geoscience 13, 504 (2020).

    Article  ADS  Google Scholar 

  36. C. T. A. Lee et al., Science 336, 64 (2012).

    Article  ADS  Google Scholar 

  37. P. R. Bevington and D. K. Robinson, Data Reduction and Error Analysis for the Physical Sciences (McGraw-Hill, New York, 2003), 3rd ed., pp. 204–208 and pp. 260–264.

  38. S. S. Lobanov et al., arXiv: 1610.01048.

  39. S. S. Lobanov et al., J. Geophys. Research: Solid Earth 122, 3565 (2017).

    Article  ADS  Google Scholar 

  40. S. Layek et al., Phys. Rev. B 94, 125129 (2016).

    Article  ADS  Google Scholar 

  41. M.-L. Boillot at al., New J. Chem. 26, 313 (2002).

    Article  Google Scholar 

  42. X.-D. Che et al., Journal of Physical Chemistry Letters 8, 5587 (2017).

    Article  Google Scholar 

  43. S. Speziale et al., PNAS 102, 17918 (2005).

    Article  ADS  Google Scholar 

  44. Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959).

    Article  ADS  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. P. Mbelek.

Appendices

APPENDIX A

As one knows, the central ion of a complex ion can make a spin crossover (or spin transition/spin equilibrium) from a low spin state to a high spin state (or conversely) by absorbing (or emitting) a photon [3436] or by migrating from a high-pressure area to a low-pressure area (or conversely). So, let us consider a distribution of identical monocinetic ions that constitute a current \(I\). We argue that these charged carriers, by making a pressure induced spin crossover at the Earth’s core-mantle boundary, will be subject to a drag force,

$$\vec{F}={-}\frac{1}{2}\frac{\left|V_{0}\right|I}{c}\vec{v},$$
(A.1)

while creating in turn a jump discontinuity equal to \(\pm V_{0}\) in the geomagnetic potentialFootnote 5 As such, this implies an infinitesimal workFootnote 6 equal to

$$\delta W=\pm\frac{1}{2}V_{0}\frac{v^{2}}{c}dQ.$$
(A.2)

Hence the expression of the infinitesimal variation of the internal energy is

$$dU=\sum_{i}{Y_{i}dX_{i}}+\mu dN\pm\frac{1}{2}\frac{v^{2}}{c}V_{0}dQ$$
$${}+\mu^{\prime}dN^{\prime}\pm\frac{1}{2}\frac{v^{\prime 2}}{c}V_{0}dQ^{\prime},$$
(A.3)

and consequently, the expression of the infinitesimal variation of the total energy

$$dE_{\textrm{total}}=d\left[N(E_{c}+E_{p})+N^{\prime}(E_{c}^{\prime}+E_{p}^{\prime})\right]+dU$$
$${}=Nd(E_{c}+E_{p})+N^{\prime}d(E_{c}^{\prime}+E_{p}^{\prime})$$
$${}+\sum_{i}{Y_{i}dX_{i}}+\left[\mu+E_{p}+\frac{1}{2}(mc\pm qV_{0})\frac{v^{2}}{c}\right]dN$$
$${}+\left[\mu^{\prime}+E_{p}^{\prime}+\frac{1}{2}(mc\pm qV_{0})\frac{v^{\prime 2}}{c}\right]dN^{\prime},$$
(A.4)

where \(E_{c}=\frac{1}{2}mv^{2}\), \(E_{c}^{\prime}=\frac{1}{2}mv^{\prime 2}\), \(dQ=qdN\), \(dQ^{\prime}=qdN^{\prime}\), \(Y_{i}\) is the intensive quantity associated to the extensive quantity \(X_{i}\), \(N\) and \(N^{\prime}\) denote the numbers of particles of mass \(m\) and electric charge \(q\) all identical up to the spin state, where the prime (respectively the absence of prime) stands for the population of those particles either in the high spin state (respectively. the low spin state) which undergoes a pressure-induced spin-crossover (see [3639]) thereby increasing the number of population in the low spin state (respectively, the high spin state). Furthermore, \(dE_{c}=0\) and \(dE_{c}^{\prime}=0\), for ions in uniform motion. By neglecting the variation of the altitude, Eq. (A.4) reduces to

$$dE_{\textrm{total}}=\sum_{i}{Y_{i}dX_{i}}$$
$${}+\left[\mu+E_{p}+\frac{1}{2}(mc^{2}\pm qcV_{0})(\frac{v^{2}}{c^{2}})\right]dN$$
$${}+\left[\mu^{\prime}+E_{p}^{\prime}+\frac{1}{2}(mc^{2}\pm qcV_{0})(\frac{v^{\prime 2}}{c^{2}})\right]dN^{\prime}$$
(A.5)

Moreover, for an isolated system in thermodynamic equilibrium, \(\mu^{\prime}=\mu\), \(N^{\prime}+N=\textrm{const}\), the total energy \(E_{\textrm{total}}\) is conserved, and all \(X_{i}\) (including the entropy) are assumed constant. Consequently, since \(E_{p}^{\prime}=E_{p}\) at the boundary where the spin crossover occurs, Eq. (A.5) reduces in turn to

$$(mc\pm qV_{0})(v^{2}-v^{\prime 2})=0.$$
(A.6)
Fig. 5
figure 5

The curve \(v\times f_{\textrm{EM}}(\psi,1)\) versus \(\psi\) in the second-order approximation, derived from the experimental data.

Let us emphasize that Eq. (A.6) holds only if either the cations are subject to no drag force and no spin crossover, in which case \(v^{\prime}=v\) with respect to the terrestrial reference frame, or otherwise \(mc\pm qV_{0}=0\) during the spin crossover and \(v^{\prime}\neq v\) in accordance with the observed velocity gradient. This means that the spin transition of iron (III) cations from the high spin state to the low spin state (or converserly) in the deep mantle of the Earth above (or below) a certain pressure and temperature, will also involve a jump discontinuity of the geomagnetic potential of an amount, \(V_{0}\), such that

$$V_{0}=\pm\frac{mc}{q}.$$
(A.7)

We argue that the above formula corresponds solely to a spin transition \(\Delta s=\pm 1\). For \(\Delta s\neq\pm 1\), it generalizes either to

$$V_{0}={-}\Delta s\frac{mc}{q},$$
(A.8)

or to

$$\Delta V=\frac{mc}{q}\Delta s.$$
(A.9)

Peculiarly, when the spin crossover does not occur, \(\Delta V=V_{0}=0\) since \(\Delta s=0\). When \(n\) various species undergo the spin crossover at the same time at a point, according to the principle of superposition, Eq. (A.8) rewrites as

$$V_{0}=-\sum_{i=1}^{i=n}{\Delta s_{i}\frac{m_{i}c}{q_{i}}},$$
(A.10)

where \(\Delta s_{i}=s_{i}({\textrm{final}})-s_{i}({\textrm{initia}}l)\), hence \(\Delta s_{i}= s_{i}({\textrm{low}})-s_{i}({\textrm{high}})\) for a spin transition from high spin to low spin and \(\Delta s_{i}=s_{i}(h{\textrm{igh}})-s_{i}({\textrm{low}})\) for a spin transition from low to high spin.

APPENDIX B

Let us recall that the (pseudo)-scalar or vector electromagnetic potentials have no direct physical meaning in classical physics because of their arbitrary choice due to the gauge invariance. However, things change drastically in quantum mechanics, as can be seen in the Aharonov–Bohm effect [40]. Throughout, based on quantum-mechanical considerations not requiring any quantum gravity treatment still lacking hitherto, as a first attempt we provide a semiclassical justification of the relation between jump discontinuities in the geomagnetic potential and subsequently in the scalar fields \(\psi\) and \(\phi\) through Eqs. (8, (9). Indeed, the vector potential \(\vec{A}\) created in space by a magnetic dipole moment \(\vec{M}\) reads (see Section 3)

$$\vec{A}=\frac{\mu_{0}}{4\pi r^{3}}\vec{M}\times\vec{r}=\vec{M}\times\frac{\vec{E}}{Nqc^{2}},$$
(B.1)

where \(\vec{E}=\dfrac{Nq\vec{r}}{4\pi\epsilon_{0}r^{3}}\) denotes the electric field created radially by a set of \(N\) identical central ions of electric charge \(q\) and mass \(m\) at a point defined by the position vector \(\vec{r}\) within a domain \((\Omega)\) of volume \(V_{\Omega}\). Now, the magnetic flux \(\Phi\) reads

$$\Phi=\iint{\vec{B}\cdot\vec{dS}}=\oint{\vec{A}\cdot\vec{dl}}$$
$${}=\oint{\bigg{(}\vec{M}\times\frac{\vec{E}}{Nqc^{2}}\bigg{)}\vec{dl}}=\frac{ME\sin\Theta}{Nqc^{2}}l,$$
(B.2)

where we have set \(\Theta=(\vec{M};\vec{r})\). Moreover, the pseudoscalar geomagnetic potential reads (see Section 3)

$$V=\frac{\mu_{0}}{4\pi r^{3}}\vec{M}\cdot\vec{r}=\vec{M}.\frac{\vec{E}}{Nqc^{2}}=\frac{ME\cos\Theta}{Nqc^{2}}.$$
(B.3)

Now, the standard deviations

$$\sigma_{V}=\left(\int\limits_{-\pi}^{\pi}{\iiint_{\Omega}{(V-\bar{V})^{2}\frac{d^{3}r}{V_{\Omega}}}\frac{d\Theta}{2\pi}}\right)^{1/2}$$

and

$$\sigma_{\Phi}=\left(\int\limits_{-\pi}^{\pi}{\iiint_{\Omega}{(\Phi-\bar{\Phi})^{2}\frac{d^{3}r}{V_{\Omega}}}\frac{d\Theta}{2\pi}}\right)^{1/2},$$

respectively, of \(V\) and \(\Phi\) reduce to

$$\sigma_{V}=\dfrac{M\left\langle E\right\rangle}{\sqrt{2}Nqc^{2}},\quad\sigma_{\Phi}=\frac{M\left\langle E\right\rangle}{\sqrt{2}Nqc^{2}}l$$

on account that both means \(V\) and \(\Phi\) are equal to zero, where we have set

$$\left\langle E\right\rangle=\left(\iiint_{\Omega}{E\frac{d^{3}r}{V_{\Omega}}}\right)^{1/2}.$$

So, by combining Eqs. (B.2) and (B.3), it follows

$$\sigma_{V}=\frac{\sigma_{\Phi}}{l}.$$
(B.4)

Analogously to the case of the Aharonov–Bohm effect, the standard deviation of the fluctuating magnetic flux associated to random jump discontinuities in the geomagnetic potential due to the spin crossover of identical central ions of mass \(m\), should be quantified through circular loops of radius \(l/(2\pi)=\hbar/(mc)\) (Compton wavelength of a particle of mass \(m\)). Thus,

$$\sigma_{\Phi}=k\frac{h}{q},$$
(B.5)

and accordingly,

$$\sigma_{V}=k\frac{mc}{q},$$
(B.6)

where \(k\) is an integer.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mbelek, J.P., Ungem, L.B. Evidence for an Effective Gravitational Constant from the Laboratory Measurements. Gravit. Cosmol. 27, 54–62 (2021). https://doi.org/10.1134/S0202289321010138

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0202289321010138

Navigation