Skip to main content
Log in

Equivalent Quantum Equations in a System Inspired by Bouncing Droplets Experiments

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

In this paper we study a classical and theoretical system which consists of an elastic medium carrying transverse waves and one point-like high elastic medium density, called concretion. We compute the equation of motion for the concretion as well as the wave equation of this system. Afterwards we always consider the case where the concretion is not the wave source any longer. Then the concretion obeys a general and covariant guidance formula, which leads in low-velocity approximation to an equivalent de Broglie-Bohm guidance formula. The concretion moves then as if exists an equivalent quantum potential. A strictly equivalent free Schrödinger equation is retrieved, as well as the quantum stationary states in a linear or spherical cavity. We compute the energy (and momentum) of the concretion, naturally defined from the energy (and momentum) density of the vibrating elastic medium. Provided one condition about the amplitude of oscillation is fulfilled, it strikingly appears that the energy and momentum of the concretion not only are written in the same form as in quantum mechanics, but also encapsulate equivalent relativistic formulas.

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

Similar content being viewed by others

Notes

  1. Let, for instance, the mass density expressed by a 3D Gaussian function, \(\rho (\mathbf {r},t)\!=\! \frac{m_0}{(\sigma \sqrt{2\pi })^3}\exp \left[ -\frac{(\mathbf {r}-\mathbf {\xi })^2}{2\,\sigma ^2}\right] \), where \(\sigma \rightarrow 0\). Hence, Eq. (6) becomes \(\left( \frac{\gamma _m}{c_m^2}\frac{\partial \frac{1}{\gamma _m}}{\partial t}\frac{\partial \varphi }{\partial t} + \left[ \!\frac{\mathbf {v}}{c_m^2} \, \frac{\partial \varphi }{\partial t}(\mathbf {\xi },t) + \mathbf {\nabla }\varphi (\mathbf {\xi },t)\right] \cdot \frac{\mathbf {r}-\mathbf {\xi }}{\sigma ^2}\!\right) \rho (\mathbf {r},t) = 0.\) This leads, when \(\mathbf {r}=\mathbf {\xi }\), to \(\frac{\partial }{\partial t} (\frac{1}{\gamma _m})=0\).

  2. Eq. (7) becomes in the low-velocity approximation, \(\mathbf {v}\,\psi (\mathbf {\xi },t) = \frac{c_m^2}{\mathrm {i}\,\Omega _m}\mathbf {\nabla }\psi (\mathbf {\xi },t)\), whose real part yields \(\mathbf {\nabla }{F}(\mathbf {\xi },t)=\mathbf {0}\). Put into the real part of Eq (7), we get \(\frac{\partial F}{\partial t}(\mathbf {\xi },t)=0\).

  3. In any reference frame where the concretion has the velocity \(\mathbf {v}\), the particle derivative is written as \(\frac{\mathrm {d}}{\mathrm {d}t}= \frac{\partial }{\partial t} + \mathbf {v}\cdot \mathbf {\nabla }\).

  4. The rest mass energy has no importance in quantum mechanics à la Schrödinger.

  5. Let \(\omega _{\mathrm {circ}, v} = \frac{v}{r_c\,\sin \theta _c}\) the angular velocity of the concretion. The velocity of the concretion derived from the guidance formula yields \(\omega _{\mathrm {circ}, v} = \frac{v^2}{c_m^2}\frac{\Omega _m}{m}\). This leads in the low-velocity approximation (note \(v = \omega _{\mathrm {circ}, v} = 0\) when \(m= 0\)) to \(\omega _{\mathrm {circ}, v}\ll \Omega _m\).

  6. The angular momentum density derives from the stress-energy tensor. Spatial integration around the location of the concretion and using time-averaged values (in the same manner as for \(W_\mathrm {conc}\) and \(\mathbf {p}_\mathrm {conc}\)) lead to \(\mathbf {L}_\mathrm {conc}\).

  7. \(\langle \partial _\mu \varphi \partial ^\mu \varphi \rangle = \left( \frac{(\Omega _m + \omega )^2}{c_m^2}-\frac{v^2}{c_m^4}(\Omega _m+\omega )^2\right) \,\langle \varphi ^2\rangle \), by using the guidance formula (7) and \(\frac{\partial F}{\partial t}=0\) at the location of the concretion. Next, the low-velocity approximation (related to \(\omega \ll \Omega _m\)), Eq. (19) and the condition (14) are used.

  8. A paper including an external potential is in preparation.

  9. Note that this difference vanishes when the concretion is in symbiosis with the wave (cf. Section 2.3).

References

  1. Couder, Y., Protière, S., Fort, E., Boudaoud, A.: Walking and orbiting droplets. Nature 437, 208 (2005)

    Article  ADS  Google Scholar 

  2. Bush, J.W.M.: Pilot-wave hydrodynamics. Annu. Rev. Fluid Mech. 47, 269–292 (2015)

    Article  ADS  Google Scholar 

  3. de Broglie, L.:La mécanique ondulatoire et la structure atomique de la matière et du rayonnement. J. de Phys. Radium, série VI, t. VIII, no 5 (1927)

  4. Couder, Y., Fort, E.: Probabilities and trajectories in a classical wave-particle duality probabilities and trajectories in a classical wave-particle duality. J. Phys. Conf. Ser. 361, 012001 (2012)

    Article  Google Scholar 

  5. Bush, J.W.M.: The new wave of pilot-wave theory. Phys. Today 68(8), 47 (2015)

    Article  Google Scholar 

  6. Borghesi, C.:Dualité onde-corpuscule formée par une masselotte oscillante dans un milieu élastique: étude théorique et similitudes quantiques. Ann. Fond. de Broglie 42(1), 161 (2017). English translation: Wave-particle duality coming from a bead oscillator in an elastic medium, theoretical study and quantum similarities; [physics.class-ph]

  7. Boudaoud, A., Couder, Y., Ben Amar, M.: A self-adaptative oscillator. Eur. Phys. J. B 9, 159–165 (1999)

    Article  ADS  Google Scholar 

  8. Holland, P.R.: Hamiltonian theory of wave and particle in quantum mechanics I: Liouville‘s theorem and the interpretation of the de Broglie-Bohm theory. Nuovo Cim. B 116, 1043–1070 (2001)

    ADS  Google Scholar 

  9. Holland, P.R.: Hamiltonian theory of wave and particle in quantum mechanics II: Hamilton-Jacobi theory and particle back-reaction. Nuovo Cim. B 116, 1143–1172 (2001)

    ADS  Google Scholar 

  10. Holland, P.R.: Quantum back-reaction and the particle law of motion. J. Phys. A 39, 559 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Durt, T.: Generalized guidance equation for peaked quantum solitons and effective gravity. Europhys. Lett. 114(1), 10004 (2016)

    Article  ADS  Google Scholar 

  12. Durt, T.: L. de Broglie’s double solution and self-gravitation. Ann. Fond. de Broglie 42(1), 73 (2017)

  13. de Broglie, L.: Une tentative d’interprétation causale et non linéaire de la mécanique ondulatoire. In: Gauthier-Villars, (ed.) Non-linear Wave mechanics— A causal Interpretation. Elsevier, Amsterdam (1960)

  14. de Broglie, L.: Interpretation of quantum mechanics by the double solution theory. Ann. fond. de Broglie 12, no 4 (1987). English translation from a paper originally published in the book Foundations of Quantum Mechanics—Rendiconti della Scuola Internazionale di Fisica “Enrico Fermi”

  15. Fort, E., Eddi, A., Boudaoud, A., Moukhtar, J.: Couder, Yves: Path-memory induced quantization of classical orbits. Proc. Natl. Acad. Sci. USA 107(41), 17515–17520 (2010)

    Article  ADS  Google Scholar 

  16. Perrard, S., Labousse, M., Miskin, M., Fort, E., Couder, Y.: Self-organization into quantized eigenstates of a classical wave-driven particle. Nat. Commun. 5, 3219 (2014)

    Article  ADS  Google Scholar 

  17. de Broglie, L.: L’univers à cinq dimensions et la mécanique ondulatoire. J. de Phys. Radium VI, t. VIII, no 2 (1927)

  18. Moláček, J., Bush, J.W.M.: Drops bouncing on a vibrating bath. J. Fluid Mech. 727, 582–611 (2013)

    Article  ADS  MATH  Google Scholar 

  19. Moláček, J., Bush, J.W.M.: Drops walking on a vibrating bath: towards a hydrodynamic pilot-wave theory. J. Fluid Mech. 727, 612–647 (2013)

    Article  ADS  MATH  Google Scholar 

  20. Fargue, D.: Louis de Broglie‘s “double solutio” a promising but unfinished theory. Ann. Fond. de Broglie 42(1), 9 (2017)

    Google Scholar 

  21. Colin, S., Durt, T., Willox, R.: L. de Broglie’s double solution program: 90 years later. Ann. Fond. de Broglie 42(1), 19 (2017)

  22. Fort, E., Couder, Y.: Trajectory eigenmodes of an orbiting wave source. Europhys. Lett. 102, 16005 (2013)

    Article  ADS  Google Scholar 

  23. Labousse, M., Perrard, S., Couder, Y., Fort, E.: Self-attraction into spinning eigenstates of a mobile wave source by its emission back-reaction. Phys. Rev. E 94, 042224 (2016)

    Article  ADS  Google Scholar 

  24. Holland, P.R.: The quantum theory of motion. Cambridge University Press, Cambridge (1995)

    MATH  Google Scholar 

  25. Oriols, X., Mompart, J.: Overview of Bohmian mechanics. In: Oriols, X., Mompart, J. (eds.) Nanoscale Systems to Cosmology. Pan Stanford, Singapore (2012)

    Google Scholar 

  26. Bohm, D.: A suggested interpretation of the quantum theory in terms of “hidde” variables. I. Phys. Rev. 85, 166–179 (1952)

    Article  ADS  MATH  Google Scholar 

  27. Borghesi, C., Moukhtar, J., Labousse, M., Eddi, A., Fort, E., Couder, Y.: Interaction of two walkers: wave-mediated energy and force. Phys. Rev. E 90, 063017 (2014)

    Article  ADS  Google Scholar 

  28. L. Landau et E. Lifchitz, Mécanique quantique. Ed. Mir, Moscow, 1975

  29. H. Poincaré, Sur la dynamique de l’électron. Rendiconti del Circolo Matematico di Palermo (1905 July 23), 1906. English translation and modernised presentation: H. M. Schwartz, Poincaré’s Rendiconti Paper on Relativity; Am. J. Phys., 39 1287–1294 (1971), 40 862–871 & 1282–1287(1972)

  30. Pierseaux, Y.: La“structure fine”de la relativité restreinte. L’Harmattan, Paris (1999)

    MATH  Google Scholar 

  31. Zee, A.: Quantum Field Theory in a Nutshell. Princeton University Press, Princeton (2003)

    MATH  Google Scholar 

  32. Perrard, S., Labousse, M., Fort, E., Couder, Y.: Chaos driven by interfering memory. Phys. Rev. Lett. 113, 104101 (2014)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank Hervé Mohrbach and Alain Bérard for help and discussions. We would like to thank the Reviewer for his constructive suggestions and insightful comments on the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Borghesi.

Appendix

Appendix

1.1 Lagrangian Density of a System with a Bead Versus One with an Elastic Medium Concretion and Some Consequences

The system presented in [6] consists of (i) a bead oscillator, i.e., a punctual mass with an “internal clock”, by which the mass tends to oscillate at a proper pulsation \(\Omega _0\) through a quadratic potential, and (ii) the same (homogeneous) elastic medium as here. The Lagrangian density of this system is thus written as

$$\begin{aligned} \mathcal {L}_{\mathrm {syst.\, with\, bead}} = \frac{1}{2}\,\rho _0 \left[ \left( \frac{\mathrm {d}\varphi }{\mathrm {d}\tau }\right) ^2 \; -\; \Omega _0^2\,\varphi ^2 \right] \, + \frac{1}{2}\, \mathcal {T}\,\left[ \partial _\mu \varphi \, \partial ^\mu \varphi \,-\, \frac{\Omega _m^2}{c_m^2} \varphi ^2 \right] , \end{aligned}$$
(24)

where \(\frac{\mathrm {d}}{\mathrm {d}\tau }\) is the particle derivative (cf. Footnote 3) expressed with \(\tau \) the natural time of the bead, \(\tau \), and \(\rho _0\) denotes its natural mass density. \(\rho _0\) of a bead of mass \(m_0\) is again written as in Eq. (1) in the proper reference frame, \(\mathcal {R}_0\), of the bead.

The difference between the Lagrangian density of a system with a bead oscillator (24) and the one with a point-like high density medium (2) naturally lies in the part dedicated to the ‘particle’—i.e., the bead or the concretion. A basic comparison between the two Lagrangian densities emphasises the following points.

  • The system with a bead has a wave-particle duality nature, which is de facto imposed from the beginning—from the density Lagrangian. In contrast, the system with a concretion is wave monistic.

  • The system with a bead has two reference pulsations, the one of the elastic medium or wave (\(\Omega _m\)) and the one proper to the bead oscillator (\(\Omega _0\)). Of course, the latter one does not exist in the monistic system with the concretion—and only remains the reference pulsation of the wave, \(\Omega _m\), which also describes the pulsation of the quadratic potential acting on the ‘particle’ concretion.

  • The bead involves in the Lagrangian density \(\left( \frac{\mathrm {d}\varphi }{\mathrm {d}\tau }\right) ^2\), while the concretion \(c_m^2\,\partial _\mu \varphi \, \partial ^\mu \varphi \) (recall in the Lagrangian density (2) that \(\mathcal {T}=\mu _0\,c_m^2\)). For the sake of simplicity we place ourself in the proper reference frame \(\mathcal {R}_0\) of the ‘particle’, where the subscript 0 means this reference frame. In one side \(\left( \frac{\mathrm {d}\varphi }{\mathrm {d}\tau }\right) ^2=\left( \frac{\partial \varphi }{\partial t_0}\right) ^2\), while on the other side \(c_m^2\,\partial _\mu \varphi \, \partial ^\mu \varphi = \left( \frac{\partial \varphi }{\partial t_0}\right) ^2 - c_m^2\,\left( \mathbf {\nabla }_0\varphi \right) ^2\). The concretion behaves thus as the bead, but with one important difference: on the contrary to the bead, the concretion takes into account the wave slope (at its location).Footnote 9 This means that the concretion, due to its elastic medium nature, is ‘more’ than a macroscopic point mass as the bead.

The concretion is admittedly ‘more’ than an usual point mass as in macroscopic systems, but it is associated for all practical purposes to a material point. The concretion, as a macroscopic point mass, has indeed well characterised location, velocity, energy and linear momentum (cf. Sect. 2.4) and angular momentum (cf. Sect. 3.4). However, in comparison to the bead, the elastic-medium nature of the concretion provides some specificities as follows.

  • The concretion can only be piloted by the wave. On the contrary to the bead, the usual equation of motion of the concretion can be ‘cancelled’ (under certain conditions), which leads to the concretion to be only piloted by the wave (cf. Section 2.3).

  • Conditions for which the ‘particle’ is no longer a wave source are less drastic for the concretion than for the bead (cf. Sect. 2.3). (It suffice for the concretion to be located at a local extremum of the transverse wave in its reference frame, whereas the bead must oscillate in its reference frame at its proper pulsation.) On the contrary to a system with a bead, a system with a concretion can have several stationary states as in equivalent quantum mechanics.

1.2 Calculation of the Wave Equation and the Equation of Motion for the Concretion

The equation of motion for the concretion comes from a principle of least action, when the four-position of the concretion is subjected to a small change, \(\xi ^\alpha \rightarrow \xi ^\alpha + \delta \xi ^\alpha \), while the wave field \(\varphi \) is fixed. This is applied to the Lagrangian of the concretion \(L_c\). From the Lagrangian density (2) and the mass density of the concretion (1), the Lagrangian of the concretion expressed in a reference frame where the concretion has a velocity \(\mathbf {v}\) is written as

$$\begin{aligned} L_c = \frac{m_0\,c_m^2}{2\,\gamma _m} \left( \partial _\mu \varphi \partial ^\mu \varphi - \frac{\Omega _m^2}{c_m^2}\,\varphi ^2 \right) , \end{aligned}$$
(25)

where \(\gamma _m=\frac{1}{\sqrt{1-v^2/c_m^2}}\). \(L_c\) is here the only Lagrangian which depends on the location of the concretion.

\(\xi ^\alpha \rightarrow \xi ^\alpha + \delta \xi ^\alpha \) leads at first order to \(\varphi (\xi ^\alpha ) \rightarrow \varphi (\xi ^\alpha ) + \partial _\alpha \varphi (\xi ^\alpha ) \, \delta \xi ^\alpha \), \(\partial _\mu \varphi \rightarrow \partial _\mu \varphi + \partial _\mu \left( \partial _\alpha \varphi \,\delta \xi ^\alpha \right) \) and \(\mathrm {d}\tau = \sqrt{\mathrm {d}\xi ^\mu \mathrm {d}\xi _\mu }/c \rightarrow \mathrm {d}\tau + \delta (\mathrm {d}\tau )\) where \(\delta (\mathrm {d}\tau ) = \frac{U_\alpha }{c_m^2}\, \mathrm {d}(\delta \xi ^\alpha )\) and \(U^\alpha =\frac{\mathrm {d}\xi ^\alpha }{\mathrm {d}\tau }\). Recall that \(\tau \) is the proper time of the concretion (so \(\mathrm {d}t = \gamma _m\,\mathrm {d}\tau \)). Thus, the conditions for which the small change \(\delta \xi ^\alpha \) implies at the first order \(\delta (\int L_c\, \mathrm {d}t) = 0\), where boundary values are fixed, are such that

$$\begin{aligned} \begin{aligned}&\int m_0\,c_m^2\,\partial _\mu \left( \partial _\alpha \varphi \,\delta \xi ^\alpha \right) \partial ^\mu \varphi \;\mathrm {d}\tau - \int m_0\, \Omega _m^2\,\varphi \,\partial _\alpha \varphi \,\delta \xi ^\alpha \;\mathrm {d}\tau \\&\quad + \int \frac{m_0}{2}\left( \partial _\mu \varphi \, \partial ^\mu \varphi - \frac{\Omega _m^2}{c_m^2} \varphi ^2 \right) \, U_\alpha \frac{\mathrm {d}(\delta \xi ^\alpha )}{\mathrm {d}\tau }\mathrm {d}\tau = 0 . \end{aligned} \end{aligned}$$
(26)

Integrating by parts the first and the third integrals, with fixed end points, and since the small change \(\delta \xi ^\alpha \) is arbitrary, lead to the equation of motion (3).

The wave equation comes from a principle of least action, when the wave field is subjected to a small change, \(\varphi \rightarrow \varphi + \delta \varphi \), while the four-position of the concretion is fixed.

\(\varphi \rightarrow \varphi + \delta \varphi \) leads to \(\delta (\varphi ^2) = 2\varphi \,\delta \varphi \) and \(\delta (\partial _\mu \varphi \,\partial ^\mu \varphi )=2\partial _\mu (\delta \varphi )\,\partial ^\mu \varphi \). According to Eq. (2) the conditions for which the small change \(\delta \varphi \) implies at first order \(\delta (\int \mathcal {L}\, \mathrm {d}t\, \mathrm {d}^3 \mathbf {r}) = 0\), where boundary values are fixed, are written as

$$\begin{aligned} \int \mathcal {T}\left( 1+\frac{\rho _0(\mathbf {r},t)}{\mu _0}\right) \left[ \partial _\mu (\delta \varphi )\,\partial ^\mu \varphi - \frac{\Omega _m^2}{c_m^2}\,\varphi \,\delta \varphi \right] \mathrm {d}t\, \mathrm {d}^3 \mathbf {r} = 0 . \end{aligned}$$
(27)

Integrating by parts the term with \(\partial _\mu (\delta \varphi )\), with fixed end points, and since the small change \(\delta \varphi \) is arbitrary, provide the wave equation (4). (Note that the generalised Euler-Lagrange equation leads to the same result, as expected.)

1.3 Calculation of the Wave Potential

Calculations in this Appendix are near for example to the ones in [25]. The particle derivative of the guidance formula (10) is written as

$$\begin{aligned} \frac{\mathrm {d}\mathbf {v}}{\mathrm {d}t} = \frac{c_m^2}{\Omega _m}\left( \frac{\partial \, \mathbf {\nabla } \Phi }{\partial t} + (\mathbf {v}\cdot \mathbf {\nabla })\mathbf {\nabla }\Phi \right) . \end{aligned}$$
(28)

Using Eq. (10) for the velocity and some basic vector calculus identities yield:

$$\begin{aligned} \frac{\mathrm {d}\mathbf {v}}{\mathrm {d}t} = \frac{c_m^2}{\Omega _m} \, \mathbf {\nabla }\left( \frac{\partial \, \Phi }{\partial t} + \frac{c_m^2}{2\,\Omega _m} (\mathbf {\nabla } \Phi )^2 \right) . \end{aligned}$$
(29)

The term in brackets in the right side is evaluated from the wave equation (4) without any source wave – since the concretion and the wave are assumed in these calculations to be in symbiosis. By using the complex notation of \(\varphi \) (8) with \(\psi = F\,\mathrm {e}^{\mathrm {i}\,\Phi }\), the real part of Eq. (4) leads to:

$$\begin{aligned} \frac{2\,\Omega _m}{c_m^2}\,\frac{\partial \, \Phi }{\partial t} \,F - \Delta F + (\mathbf {\nabla } \Phi )^2 \, F = 0 , \end{aligned}$$
(30)

in which terms with \((\frac{\partial \, \Phi }{\partial t})^2\) and \(\frac{\partial ^2 \, F}{\partial t^2}\) are neglected in the low-velocity approximation. Put into Eq. (29) it appears that the concretion moves under the influence of the potential Q given in Eq. (12).

1.4 From the Klein–Gordon-like Equation Without Source to the Equivalent Free Schrödinger Equation

The derivation of the free Schrödinger from the Klein–Gordon equation (without source) in the low-velocity approximation is mentioned by de Broglie (see e.g. [13] §II.7) and is well-known in the literature (see e.g., [31] §III.5). Here, by using the modulating wave \(\psi \) contained in \(\varphi \) (8), the Klein–Gordon-like equation (4) without source yields

$$\begin{aligned} \mathrm {Re}\left[ \left( \frac{1}{c_m^2}\left[ -\,\Omega _m^2\,\psi - 2\,\mathrm {i}\,\Omega _m\,\frac{\partial \psi }{\partial t} + \frac{\partial ^2 \psi }{\partial t^2}\right] - \, \Delta \psi \, + \, \frac{\Omega _m^2}{c_m^2}\,\psi \right) \mathrm {e}^{-\mathrm {i}\,\Omega _m t}\right] = 0 .\qquad \ \end{aligned}$$
(31)

Since this equation holds for any t, we get:

$$\begin{aligned} - \frac{1}{c_m^2}\frac{\partial ^2 \psi }{\partial t^2} - 2\,\mathrm {i}\frac{\Omega _m}{c_m^2}\frac{\partial \psi }{\partial t} - \Delta \psi = 0 . \end{aligned}$$
(32)

In the low-velocity approximation, the term \(\frac{\partial ^2 \psi }{\partial t^2}\) is negligible with respect to the other terms.

To convince ourselves of the validity of this approximation, we consider a free concretion in symbiosis with a plane wave \(\varphi \) (cf. also Sect. 3.2). By using \(\varphi = A\, \cos (\mathbf {k}\cdot \mathbf {r}-\Omega \,t)\), the Klein–Gordon-like equation (4) leads to \(c_m^2\,k^2=\Omega ^2-\Omega _m^2\); and the \(\varphi \)-guidance formula (7) leads to \(\mathbf {v}=\frac{c_m^2 \, \mathbf {k}}{\Omega }\). Thus \(\Omega = \frac{1}{\sqrt{1-v^2/c_m^2}}\,\Omega _m\). According to Eq. (8), the modulating wave is written as \(\psi = A\,\mathrm {e}^{\mathrm {i}(\mathbf {k}\cdot \mathbf {r}-\omega \,t)}\), where \(\omega = \Omega - \Omega _m\). This leads to \(\omega =\frac{1}{2}\frac{v^2}{c_m^2}\,\Omega _m\) in the low-velocity approximation. Then, the term \(\frac{1}{c_m^2}\frac{\partial ^2 \psi }{\partial t^2}\) is proportional to \(\frac{\Omega _m^2}{c_m^2}\frac{v^4}{c_m^4}\,\psi \), while the two other terms in Eq. (32) are proportional to \(\frac{\Omega _m^2}{c_m^2}\frac{v^2}{c_m^2}\,\psi \).

In the low-velocity approximation, the modulating wave \(\psi \) is thus governed by the equivalent free Schrödinger equation (17).

1.5 Energy and Linear Momentum of the Concretion

As usual, the energy density, \(\rho _e\), and the momentum density, \(\mathbf {g}\), are evaluated from the stress–energy tensor T of the system (\(T^{\mu \nu }=\frac{\partial \,\mathcal {L}}{\partial (\partial _\mu \varphi )}\,\partial ^\nu \varphi - \eta ^{\mu \nu }\,\mathcal {L}\), where \(\eta ^{\mu \nu }\) denotes the Minkowski metric, with the signature \((+,-,-,-)\) adopted throughout this article). The Lagrangian density (2), where \(\mathcal {T} = \mu _0\,c_m^2\), leads to

$$\begin{aligned} \rho _e= & {} \frac{\mu _0}{2}(1 + \frac{\rho _0}{\mu _0})\left[ \left( \frac{\partial \varphi }{\partial t}\right) ^2 + \left( c_m\,\mathbf {\nabla }\varphi \right) ^2 + \Omega _m^2\,\varphi ^2\right] \nonumber \\ \mathbf {g}= & {} -\,\mu _0\,(1 + \frac{\rho _0}{\mu _0})\,\frac{\partial \varphi }{\partial t}\, \mathbf {\nabla }\varphi . \end{aligned}$$
(33)

In the following we use their time-averaged values during one oscillation (written as \(\langle \cdots \rangle \)). Spatial integration around the location of the concretion provides the energy, \(W_\mathrm {conc}\), and the momentum, \(\mathbf {p}_\mathrm {conc}\), of the concretion. The following expressions are given in a reference frame, \(\mathcal {R}\), where the concretion has the velocity \(\mathbf {v}\).

We evaluate first the case for which the condition (11) is satisfied, in addition to the \(\varphi \)-guidance formula. Eqs. (1), (7), (11) and a little bit of algebra yield:

$$\begin{aligned} W_\mathrm {conc}= & {} \gamma _m \, m_0\, \Omega _m^2\, \left\langle \, \varphi ^2(\mathbf {\xi },t)\, \right\rangle \nonumber \\ \mathbf {p}_\mathrm {conc}= & {} \gamma _m \, m_0\, \frac{\Omega _m^2}{c_m^2} \, \left\langle \, \varphi ^2(\mathbf {\xi },t)\, \right\rangle \, \mathbf {v} . \end{aligned}$$
(34)

We have notably used \(\langle \frac{1}{\gamma _m^2} (\frac{\partial \varphi }{\partial t}(\mathbf {\xi },t))^2 \, \rangle = \Omega _m^2\,\langle \varphi ^2(\mathbf {\xi },t) \rangle \). Since the amplitude of oscillation of the concretion, \(F_c\), remains constant in time, it follows equations (13).

Now we consider that the concretion has just its velocity given by the \(\varphi \)-guidance formula (7). Let \(\Omega \) (where \(\Omega =\Omega _m+\omega \)) the pulsation of the wave \(\varphi \) in \(\mathcal {R}\). In the same manner as above, Eq. (33) leads to

$$\begin{aligned} W_\mathrm {conc}= & {} \frac{1}{2}\,\frac{m_0}{\gamma _m}\, \left[ \left( 1+\frac{v^2}{c_m^2}\right) \, \left\langle \, \left( \frac{\partial \varphi (\mathbf {\xi },t)}{\partial t}\right) ^2\; \right\rangle \;+\; \Omega _m^2\, \langle \, \varphi ^2(\mathbf {\xi },t)\, \rangle \right] \nonumber \\ \mathbf {p}_\mathrm {conc}= & {} \frac{m_0}{\gamma _m\, c_m^2}\, \left\langle \,\left( \frac{\partial \varphi (\mathbf {\xi },t)}{\partial t}\right) ^2\; \right\rangle \, \mathbf {v}. \end{aligned}$$
(35)

In the low-velocity approximation, \(\frac{v^2}{c_m^2}\) and \(\frac{\omega }{\Omega _m}\) have the same order of magnitude. (To convince us, consider how a pulsation \(\Omega _m\) in \(\mathcal {R}_0\) becomes in \(\mathcal {R}\), for example by using Lorentz-Poincaré transformation.) Moreover \(\langle \, (\frac{\partial \varphi }{\partial t}(\mathbf {\xi },t))^2 \, \rangle = \Omega ^2\,\langle \, \varphi ^2(\mathbf {\xi },t) \,\rangle \) – because the magnitude F at the location of the concretion is \(F_c\), constant in time. Taking into account the condition (14), equations (35) become Eqs. (15) and (16) at first-order approximation in \(\frac{v^2}{c_m^2}\) and \(\frac{\omega }{\Omega _m}\).

1.6 Superposition of Eigenstates in the Linear Cavity

We study the superposition of eigenstates for the concretion in the linear cavity (cf. Sect. 3.3). To have a deeper meaning, we deal with the transverse wave, \(\varphi \), rather than the modulating wave, \(\psi \). According to the Klein–Gordon-like equation (4) without source and the boundary conditions, an eigenstate is written as \(\varphi _n(x,t)=A_n\,\sin (K_n\,x)\,\cos (\Omega _n\,t -\theta _n)\), where \(K_n = \frac{n\,\pi }{L}\), \(c_m^2\,K_n^2 = \Omega _n^2-\Omega _m^2\), \(A_n\) denotes an amplitude of transverse oscillations and \(\theta _n\) a phase shift, irrelevant in this study. (This expression is in agreement with the one of \(\psi \) written in Sect. 3.3 when \(\omega _n(=\Omega _n-\Omega _m) \ll \Omega _m\).) For the sake of simplicity we consider two eigenstates, \(n=1\) and 3 (two odd numbers), such that the superposition is written as

$$\begin{aligned} \varphi (x,t) = \frac{A}{\sqrt{2}}\,\left[ \sin (K_1\,x)\,\cos (\Omega _1\,t) + \sin (K_3\,x)\,\cos (\Omega _3\,t)\right] , \end{aligned}$$
(36)

where A is an amplitude of transverse oscillations.

In \(x=L/2\), the wave slope \(\mathbf {\nabla }\varphi (x=L/2,t) = 0\) for any t. According to the \(\varphi \)-guidance formula (7), the concretion can be thus located at this point and remain here—as for any eigenstate with an odd number n.

Let us now study the energy of the concretion remaining at \(x=L/2\). According to Eq. (35), after averaging over a transverse oscillation period and when \(\omega _n \ll \Omega _m\), the energy of the concretion is

$$\begin{aligned} W_\mathrm {conc}= \frac{1}{2}\,m_0\,A^2\,\Omega _m^2\left[ \left( 1+\cos ( \Delta \omega \, t)\right) \left( 1 + \frac{\omega _1+\omega _3}{2\,\Omega _m}\right) \right] , \end{aligned}$$
(37)

where \(\Delta \omega = \omega _3-\omega _1\). Rather than calculating the average value of \(W_\mathrm {conc}\) over time, it appears that the total energy of the concretion, \(W_\mathrm {conc}\), is periodically equal to zero. The analogy with quantum mechanics seems no longer to hold. Furthermore a total energy of the concretion equal to zero does not seem to be realistic, in particular if the ‘particle’ concretion is a simplified representation of a soliton. Then, the toy model suggested in this paper should not be suitable for dealing with the superposition of eigenstates. If more complex considerations demand that the kind of concretion has a total energy near to \(m_0\,c_m^2\), we could expect that same phenomena as for walkers appear: the system could exhibit very short transitions between eigenstates [32].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Borghesi, C. Equivalent Quantum Equations in a System Inspired by Bouncing Droplets Experiments. Found Phys 47, 933–958 (2017). https://doi.org/10.1007/s10701-017-0103-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-017-0103-6

Keywords

Navigation