This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\fnm

David \surDarrow

\orgname

Massachusetts Institute of Technology, \orgdivDepartment of Mathematics, \cityCambridge, \stateMA, \countryUSA

Convergence to Bohmian mechanics in a de Broglie-like pilot-wave system

Abstract

Bohmian mechanics supplements the quantum wavefunction with deterministic particle trajectories, offering an alternate, dynamical language for quantum theory. However, the Bohmian particle does not affect its guiding wave, so the wave field must instead be prescribed by the system geometry. While this property is widely assumed necessary to ensure agreement with quantum mechanics, much work has recently been dedicated to understanding classical pilot-wave systems, which feature a two-way coupling between particle and wave. These systems—including the “walking droplet” system of Couder and Fort [1] and its various abstractions [2, 3, 4, 5]—allow us to investigate the limits of classical systems and offer a touchstone between quantum and classical dynamics. In this work, we present a general result that bridges Bohmian mechanics with this classical pilot-wave theory. Namely, Darrow and Bush [5] recently introduced a Lagrangian pilot-wave framework to study quantum-like behaviours in classical systems; with a particular choice of particle-wave coupling, they recover key dynamics hypothesised in de Broglie’s early double-solution theory [6, 7]. We here show that, with a different choice of coupling, their de Broglie-like system reduces exactly to single-particle Bohmian mechanics in the non-relativistic limit. Finally, we present an application of the present work in developing an analogue for position measurement in a de Broglie-like setting.

keywords:
Klein–Gordon equation, Hydrodynamic quantum analogues, Pilot-wave theory, Zitterbewegung, Lagrangian mechanics

1 Introduction

In his 1924 thesis [8], Louis de Broglie predicted that massive particles would behave like waves of the eponymous de Broglie wavenumber:

k=p/,k=p/\hbar,

where p=γmup=\gamma mu is the relativistic momentum of a particle of mass mm and velocity uu. His vision deviated somewhat from the prevailing theories of Schrödinger, Bohr, and Heisenberg, however; in the subsequent years, he worked on his double-solution theory, which posited that a quantum particle consists of both particle and wave, evolving in unison [6, 7]. The wave field would contain a singularity at the particle’s position, though its form was unspecified by de Broglie, and the interplay between the particle and wave would give rise to emergent statistical behaviour in line with quantum theory. De Broglie famously presented a version of his theory at the 1927 Solvay conference, to mixed reception, and (temporarily) abandoned the pilot-wave approach shortly thereafter [9].

In 1952, David Bohm arrived at a different pilot-wave model of quantum mechanics [10, 11], providing an important counterexample to the “impossibility proofs” of von Neumann and others, which had suggested the impossibility of any hidden variable theory [12]. Alongside the Schrödinger wavefunction ψ\psi, Bohm posits a deterministic, physical trajectory for each particle in the system [10, 11, 13]. Restricting to the non-relativistic, one-particle case for simplicity, the wavefunction remains that of Schrödinger,

itψ=22m2ψ+V(x)ψ,i\hbar\partial_{t}\psi=-\frac{\hbar^{2}}{2m}\nabla^{2}\psi+V(x)\psi, (1)

and the particle position qpq_{p} is guided along as

q˙p=mIm(ψψ).\dot{q}_{p}=\frac{\hbar}{m}\operatorname{Im}\left(\frac{\nabla\psi}{\psi}\right). (2)

Notably, if Born’s rule (i.e., P(qp=q)|ψ(q)|2P(q_{p}=q)\propto|\psi(q)|^{2}) holds at one time, the trajectory equation (2) ensures that it holds at arbitrary times—as such, Bohmian predictions align exactly with those of standard quantum theory [10].

Bohm’s work inspired de Broglie to return to pilot-wave dynamics [14], and it also caught the attention of John Bell, who identified the mathematical gaps in von Neumann’s proof [15]. Bell also proved his own no-go theorem [16], demonstrating that any hidden variable approach to quantum mechanics must be inherently non-local; his work was corroborated by a famous experiment of Alain Aspect in 1982 [17], lending credence to Bohm’s non-local theory of particle trajectories.

A key property of the Bohmian approach, however, is that the particle’s Bohmian position has no influence on the wavefunction. The wavefunction evolution is prescribed by the system geometry, and the particle navigates its guiding wave without altering it. While this property has thus far been assumed necessary for agreement with quantum mechanics (with a few exceptions, e.g., [18, 19]), it marks a significant departure both from de Broglie’s earlier pilot-wave approach and from classical field theories, and it raises real consequences for how one can interpret the theory. As an example, one can see the one-way coupling as equivalent to the problem of “empty waves” in Bohmian mechanics; if the particle cannot affect the wave field, it is difficult to distinguish a field that is carrying a particle from one that is not.

In a different direction, there has been a great deal of recent interest in understanding classical pilot-wave systems, largely inspired by the “walking droplet” system of Couder and Fort [1]. In their system, a millimetric droplet moves about the surface of a vibrating fluid by interacting with a wave of its own creation. Walking droplets have been shown to engage in a variety of quantum-like behaviours: among many other examples [20], they have been found to undergo single-particle slit diffraction and interference [1, 21, 22], attain quantised orbits and stable spin states [23, 24, 25, 26], give rise to quantum-like statistics in an analogue of the quantum corral [27], and follow “surreal trajectories” similar to those predicted by Bohmian mechanics [28]. While the majority of these analogues are only qualitative, recent authors have attempted to construct classical field theories to achieve quantitative agreement with certain quantum results [3, 4], furthering the theoretical program initiated by Louis de Broglie [6, 7].

In fact, like de Broglie’s double-solution program before them, these classical systems are defined by their two-way coupling; the highly nonlinear interplay between particle and field gives rise to interesting dynamical behaviours, seemingly intractably at odds with the simple wave-guiding-particle approach of Bohmian mechanics.

In the present work, we work to connect these two classes of pilot-wave theory. We make use of the Lagrangian pilot-wave framework put forward by Darrow and Bush [5], which couples a Klein–Gordon field to a relativistic point particle. With a particular choice of particle-wave coupling, they recover many of the key behaviours hypothesised in de Broglie’s double-solution theory, as we describe in Section 2.

We investigate their framework with a different choice of coupling, recovering a distinct set of quantum-like behaviours. Namely, we show that, in the non-relativistic limit, the system converges exactly to single-particle Bohmian mechanics, where the guiding wave (less a particle-centred wavepacket) plays the role of the Schrödinger wavefunction. This result holds in a highly general setting of the pilot-wave framework, with arbitrary domain, initial conditions, and potential. Moreover, it extends the relation between Klein–Gordon and Schrödinger waves [29] to the case of a particle-wave system, and it provides a means by which Bohmian mechanics can emerge from a pilot-wave theory closer to de Broglie’s original double-solution program.

A related program was considered by Sutherland [18, 19], who showed that a certain particle-wave coupling could be “gauged away” to reveal a match with Bohmian mechanics. However, the current work undertakes a different mechanism to decouple the wave field from particle feedback, and which cannot be added or removed with an appropriate gauge transformation. This difference ensures that, in limits where it does not converge to Bohmian mechanics, our system is able to undergo the same nonlinear evolution characteristic of the classical pilot-wave theories discussed above.

As a demonstration of the latter, we introduce an analogue of wavefunction collapse in our de Broglie-like pilot-wave system. Under an appropriately-defined analogue of a (single-particle) position measurement, we show that the wavefunction collapses exactly to its projection in the measured region containing the particle.

The present work helps to bridge the gap between two different perspectives on pilot-wave dynamics. Bohmian mechanics exactly recovers the predictions of quantum mechanics, but at the cost of a one-way coupling from wave to particle; classical pilot-wave systems maintain a two-way coupling, and they can recover convincing analogues of quantum dynamics in certain cases, but it is far from clear whether they could replicate even single-particle quantum predictions completely. Our work answers this latter question in the affirmative, and it demonstrates how the pilot-wave framework of Darrow and Bush neatly interpolates between Bohmian mechanics and classical—or at least de Broglie-like—dynamics, suggesting a stronger link between these systems than previously thought.

2 De Broglie-like pilot-waves

We briefly review the Lagrangian pilot-wave framework put forward by Darrow and Bush [5], before moving on to the specific limit under consideration. Using a joint particle-wave action, they couple a complex Klein–Gordon field ϕ\phi to a relativistic particle of mass mm at the point qp3\vec{q}_{p}\in\mathbb{R}^{3}. Extremising against both objects, they find that the particle trajectory and field satisfy the coupled equations

(μμ+kc2)ϕ=kc1γ1δ3(qqp)σϕ,dt(mσγu)=γ1mc2σ(qp),\begin{gathered}(\partial_{\mu}\partial^{\mu}+k_{c}^{2})\phi=k_{c}^{-1}\gamma^{-1}\delta^{3}(\vec{q}-\vec{q}_{p})\frac{\partial\sigma}{\partial\phi^{*}},\\ d_{t}\left(m\sigma\gamma\vec{u}\right)=\gamma^{-1}mc^{2}\nabla\sigma(q_{p}),\end{gathered} (3)

where u\vec{u} is the velocity of the particle, γ=(1(uc)2)1/2\gamma=(1-(\frac{\vec{u}}{c})^{2})^{-1/2} its Lorentz factor, kc=mc/k_{c}=mc/\hbar its Compton wavenumber, and σ=σ(ϕ)\sigma=\sigma(\phi) a real-valued coupling function111In fact, σ\sigma should be a function of the continuous part ϕ¯{\overline{\phi}}, which we define shortly. This distinction does not play into the current discussion. between particle and wave.

Darrow and Bush focus on a particular limit of interest, where

σ=1+b2/4π+bReϕ\sigma=1+b^{2}/4\pi+b\operatorname{Re}\phi

for a real coupling parameter bb. This limit formalises many of the properties de Broglie hypothesised for his double-solution program [6, 7], including a synchronised Compton-scale oscillation in both particle and field, the de Broglie relation p=k\vec{p}=\hbar\vec{k} in the vicinity of the particle, and a Compton-scale wavepacket that traces out the particle’s trajectory.

In the present work, we consider a new limit of (3), where the particle is coupled to the wave’s phase rather than its amplitude. Recall from Darrow and Bush [5] the definition of the continuous component of ϕ\phi:

ϕ¯(q)=limr014πS2𝑑ξr(rϕ(q+rξ)).{\overline{\phi}}(q)=\lim_{r\to 0}\frac{1}{4\pi}\int_{S^{2}}d\xi\;\partial_{r}(r\phi(q+r\xi)).

Where ϕ\phi is continuous, we find ϕ=ϕ¯\phi={\overline{\phi}}; however, if

ϕ=ϕ1+a/qqp\phi=\phi_{1}+a/\|q-q_{p}\| (4)

for a continuous function ϕ1\phi_{1} near a point qpq_{p}, we instead find ϕ¯(qp)=ϕ1(qp){\overline{\phi}}(q_{p})=\phi_{1}(q_{p}). With this in mind, define the continuous wave phase to be

θ:=Imlogϕ¯=12i(logϕ¯logϕ¯)\theta:=\operatorname{Im}\log{\overline{\phi}}=\frac{1}{2i}\left(\log{\overline{\phi}}-\log{\overline{\phi}}^{*}\right)

at the point qpq_{p}. This phase is well-defined (up to a period of 2π2\pi) and finite.

Now, we define our particle-wave coupling by the following choice of sawtooth wave for σ\sigma:

σ(ϕ)=a+bθ,0<θ2π,\sigma(\phi)=a+b\theta,\qquad 0<\theta\leq 2\pi, (5)

where σ/b<0\sigma/b<0 for all θ\theta; we can ensure this condition by choosing aa and bb to have opposite signs with |a|>2π|b||a|>2\pi|b|.

We make this choice of σ\sigma to ensure that dynamics are independent of global phase shifts, but much of our analysis can be extended to a generic function σ=σ(θ)\sigma=\sigma(\theta). We will briefly comment on alternate choices in the following section.

Finally, applying (3) to the coupling (5), we find

(μμ+kc2)ϕ=δ3(qqp)bγ12ikcϕ¯,dt(γmσ(θ)u)=γ1mc2σ(θ).\begin{gathered}\left(\partial^{\mu}\partial_{\mu}+k_{c}^{2}\right)\phi=-\delta^{3}(q-q_{p})\frac{b\gamma^{-1}}{2ik_{c}{\overline{\phi}}^{*}},\\ d_{t}(\gamma m\sigma(\theta)u)=-\gamma^{-1}mc^{2}\nabla\sigma(\theta).\end{gathered} (6)

The means by which these equations reduce to (1) and (2) will be detailed in Section 3.

There are several immediate differences we can identify from the limit discussed by Darrow and Bush [5]. Two features are familiar from quantum mechanics: ϕ\phi is now necessarily complex-valued, and the particle dynamics are now independent of field amplitude. For a more subtle distinction, recall that a Klein–Gordon field decomposes in the non-relativistic limit as ϕψeiωct\phi\sim\psi e^{-i\omega_{c}t}, taking on a global phase oscillation at the Compton frequency. In their limit, this oscillation imparts a particle vibration at the same frequency, as ϕ\nabla\phi itself oscillates—this Zitterbewegung plays a critical role in recovering aspects of de Broglie’s double-solution theory. With the current choice of coupling, however, the force σθ\nabla\sigma\sim\nabla\theta no longer changes under a global change of phase, and we expect (and will demonstrate shortly) that the Zitterbewegung should disappear in the non-relativistic limit of the present system.

3 Reduction to Bohmian Mechanics

In parallel with the non-relativistic limit of the classical Klein–Gordon theory [29], we introduce an oscillating ansatz of the form

ϕ=ψeiωctbeiωct8πikcψ(qp)qqp=:ψeiωct+ϕwav,\phi=\psi e^{-i\omega_{c}t}-\frac{be^{-i\omega_{c}t}}{8\pi ik_{c}\psi^{*}(q_{p})\|q-q_{p}\|}=:\psi e^{-i\omega_{c}t}+\phi_{\text{wav}}, (7)

with ωc=mc2/\omega_{c}=mc^{2}/\hbar the Compton frequency of our particle, kc=mc/k_{c}=mc/\hbar the corresponding wavenumber, and bb as in (5). Notably, the continuous component of our wave is simply ϕ¯=ψeiωct{\overline{\phi}}=\psi e^{-i\omega_{c}t}. We show here that, in the non-relativistic limit, the evolution of ψ\psi approaches that of a Schrödinger wavefunction, and the evolution of qpq_{p} approaches that of a Bohmian particle guided by ψ\psi.

Specifically, we demonstrate the following result:


Claim.

Suppose ϕ\phi and qpq_{p} evolve under (6) in a connected, open set Ω3\Omega\subset\mathbb{R}^{3}, with zero Dirichlet conditions on Ω\partial\Omega (if the boundary is non-empty). Fix initial conditions ψ0\psi_{0}, ψ˙0\dot{\psi}_{0}, qp,0q_{p,0}, and q˙p,0\dot{q}_{p,0}, such that ψ0(qp,0)0\psi_{0}(q_{p,0})\neq 0, and let ϕ\phi be given by (7). Let (ψB,qpB)(\psi^{B},q_{p}^{B}) be the free Bohmian process in Ω\Omega with the same initial conditions.

In the non-relativistic limit ucu\ll c, |ψ˙|ωc|ψ||\dot{\psi}|\ll\omega_{c}|\psi|, the quantities ψ\psi and qpq_{p} converge to ψB\psi^{B} and qpBq_{p}^{B}, respectively.


Remark.

We present our argument only in the free (or wall-bounded) case for mathematical clarity. However, at the end of this section, we describe how the same argument extends to a non-zero potential V(q)V(q).


To show this result, we instead demonstrate that such a convergence occurs for a short time, with t(0,τ]t\in(0,\tau] for some

τ=τ(|ψ0(qp,0)|,|ψ0(qp,0)|,|q˙p,0|).\tau=\tau(|\psi_{0}(q_{p,0})|,|\nabla\psi_{0}(q_{p,0})|,|\dot{q}_{p,0}|).

In particular, if we define

M=|ψ0(qp,0)|1+|ψ0(qp,0)|+|q˙p,0|,M=|\psi_{0}(q_{p,0})|^{-1}+|\nabla\psi_{0}(q_{p,0})|+|\dot{q}_{p,0}|,

we argue that τ\tau is a strictly decreasing function of MM. In Bohmian mechanics, however, all of these terms are locally uniformly bounded, since the particle never intersects the ψ=0\psi=0 locus; that is, for any T>0T>0, we can fix

MT=sup(0,T](|ψB(qpB)|1+|ψB(qpB)|+|q˙pB|)M_{T}=\sup_{(0,T]}\left(|\psi^{B}(q^{B}_{p})|^{-1}+|\nabla\psi^{B}(q^{B}_{p})|+|\dot{q}^{B}_{p}|\right)

and choose a corresponding τT=τT(MT)\tau_{T}=\tau_{T}(M_{T}) that ensures convergence for any subprocess t(t0,t0+τ](0,T]t\in(t_{0},t_{0}+\tau]\subset(0,T]. Piecing these processes together, we see that the process itself converges over the full interval—and since TT was arbitrary, that it converges for all time.

Trajectory equation.

We focus first on the trajectory equation, as it allows us to rule out fast particle oscillations. Specifically, we show that, if the particle does not currently satisfy the guidance equation (2), the “error” in its velocity is quickly dampened until it does. This property guarantees convergence to the Bohmian trajectory even if the particle does not begin with the correct, Bohmian velocity. Moreover, this is a significant difference from the limit studied by Darrow and Bush [5], where the particle undergoes fast Zitterbewegung oscillations about its mean trajectory.

As a note, because the singular component of ϕ\phi is symmetric about qpq_{p}, we self-consistently define

θ:=Imlog(ψeiωct)\nabla\theta:=\nabla\operatorname{Im}\log(\psi e^{-i\omega_{c}t})

at the point qpq_{p}. At the level of derivatives, this is simply a mean gradient over a small ball centred at qpq_{p}.

The time derivative of σ\sigma is

σ˙(θ)\displaystyle\dot{\sigma}(\theta) =dt(Imlogψωct)σ(θ)\displaystyle=d_{t}(\operatorname{Im}\log\psi-\omega_{c}t)\sigma^{\prime}(\theta)
=ωcσ(θ)(1+O(|ψ˙|/|ψ|ωc)),\displaystyle=-\omega_{c}\sigma^{\prime}(\theta)(1+O(|\dot{\psi}|/|\psi|\omega_{c})),

which transforms our guiding equation—to leading order in u/cu/c—into

ωcσ(θ)u+σ(θ)u˙=c2σ(θ)θ+O(u2/c2).-\omega_{c}\sigma^{\prime}(\theta)u+\sigma(\theta)\dot{u}=-c^{2}\sigma^{\prime}(\theta)\nabla\theta+O(u^{2}/c^{2}).

Now, note that

θ=Imlogψ=Im(ψψ),\nabla\theta=\nabla\operatorname{Im}\log\psi=\operatorname{Im}\left(\frac{\nabla\psi}{\psi}\right),

which allows us to recover

uσωcσu˙=mIm(ψψ)+O(u2/c2),u-\frac{\sigma}{\omega_{c}\sigma^{\prime}}\dot{u}=\frac{\hbar}{m}\operatorname{Im}\left(\frac{\nabla\psi}{\psi}\right)+O(u^{2}/c^{2}),

or, plugging in our expression (5),

uσbωcu˙=mIm(ψψ)+O(u2/c2).u-\frac{\sigma}{b\omega_{c}}\dot{u}=\frac{\hbar}{m}\operatorname{Im}\left(\frac{\nabla\psi}{\psi}\right)+O(u^{2}/c^{2}). (8)

There are two possibilities with these dynamics.

Naïvely, it appears that the u˙\dot{u} term vanishes in the non-relativistic limit, and this is indeed a solution of (8). In this case, the dynamics reduce to

u=mIm(ψψ),u=\frac{\hbar}{m}\operatorname{Im}\left(\frac{\nabla\psi}{\psi}\right), (9)

identical to the equation (2) guiding Bohmian particles.

As a second possibility, we can take a cue from the model of Darrow and Bush [5] and note that these dynamics might support a transient motion about the Bohmian trajectory (9). In particular, the full trajectory can generically take the form

utot=uBohm+ufluc,u_{\text{tot}}=u_{\text{Bohm}}+u_{\text{fluc}},

where uflucu_{\text{fluc}} satisfies the unforced equation

uflucσ(ωct)bωcu˙fluc=0.u_{\text{fluc}}-\frac{\sigma(-\omega_{c}t)}{b\omega_{c}}\dot{u}_{\text{fluc}}=0. (10)

With the particle-wave coupling we have fixed, however, we have that σ/b<0\sigma/b<0; then the equation (10) dampens uflucu_{\text{fluc}} to zero within a Compton timescale 1/ωc\sim 1/\omega_{c}.

Field equation.

Following a similar argument as carried out by Darrow and Bush [5] in their limit of interest, the ansatz (7) ensures that ψ\psi is continuous at the point qpq_{p}, and that ϕ¯(qp)=ψ(qp)eiωct{\overline{\phi}}(q_{p})=\psi(q_{p})e^{-i\omega_{c}t}. As such, we can write the nonlinear forcing in the wave equation as

(μμ+kc2)ϕ=δ3(qqp)γ1beiωct2ikcψ.\left(\partial^{\mu}\partial_{\mu}+k_{c}^{2}\right)\phi=-\delta^{3}(q-q_{p})\frac{\gamma^{-1}be^{-i\omega_{c}t}}{2ik_{c}\psi^{*}}. (11)

To understand how the particle-to-wave and wave-to-particle effects might decouple in the non-relativistic limit, first consider a simple case where qpq_{p} and ψ\psi are unchanging in time. Then the singular component of ϕ\phi satisfies

(μμ+kc2)ϕwav=2ϕwav=δ3(qqp)beiωct2ikcψ,(\partial^{\mu}\partial_{\mu}+k_{c}^{2})\phi_{\text{wav}}=-\nabla^{2}\phi_{\text{wav}}=-\delta^{3}(q-q_{p})\frac{be^{-i\omega_{c}t}}{2ik_{c}\psi^{*}}, (12)

which would exactly account for our nonlinear forcing. As we will see, this simple case reflects the system in general; in the non-relativistic limit, the Compton-frequency oscillation dominates the evolution of this source term, and the fixed-time solution ϕwav\phi_{\text{wav}} becomes a true solution to the forced Klein–Gordon equation. The remainder, ψeiωct\psi e^{-i\omega_{c}t}, then evolves according to the unforced equation.

More rigorously, the above argument identifies

ϕwav,t0:=beiωct8πikcψ(qp,t0)qqp\phi_{\text{wav},t_{0}}:=-\frac{be^{-i\omega_{c}t}}{8\pi ik_{c}\psi^{*}(q_{p},t_{0})\|q-q_{p}\|} (13)

as the unique wavepacket solution to the fixed-ψ\psi problem

(μμ+kc2)ϕwav,t0=δ3(qqp)beiωct2ikcψ(qp,t0).(\partial^{\mu}\partial_{\mu}+k_{c}^{2})\phi_{\text{wav},t_{0}}=-\delta^{3}(q-q_{p})\frac{be^{-i\omega_{c}t}}{2ik_{c}\psi^{*}(q_{p},t_{0})}.

In moving to the generic case, where ψ\psi varies in time, we proceed to show that the inhomogeneous portion of the solution remains close to (13).

To this end, recall the form of the retarded Klein–Gordon Green’s function [30]:

G(q,t)\displaystyle G(q,t) =θ(t)4πqδ(tq/c)\displaystyle=\frac{\theta(t)}{4\pi\|q\|}\delta(t-\|q\|/c)
kc22πθ(ctq)J1(ωct2q2/c2)ωct2q2/c2.\displaystyle\qquad-\frac{k_{c}^{2}}{2\pi}\theta(ct-\|q\|)\frac{J_{1}(\omega_{c}\sqrt{t^{2}-\|q\|^{2}/c^{2}})}{\omega_{c}\sqrt{t^{2}-\|q\|^{2}/c^{2}}}.

Here, θ\theta is the Heaviside step function, which ensures that data travels forward in time, and J1J_{1} is a Bessel function of the first kind. The first term in this expression communicates a simple Coulomb potential; the second term corrects this potential with an internal vibration at the Compton frequency, characteristic of any massive field.

An exact solution to the time-varying problem (11) is thus given by

ϕ~wav\displaystyle\tilde{\phi}_{\text{wav}} =t𝑑sG(qqp(s),s)beiωcs2ikcψ(s)\displaystyle=-\int_{-\infty}^{t}ds\;G(q-q_{p}(s),s)\frac{be^{-i\omega_{c}s}}{2ik_{c}\psi^{*}(s)} (14)
=t𝑑sG(qqp(t),s)beiωcs2ikcψ(t)\displaystyle=-\int_{-\infty}^{t}ds\;G(q-q_{p}(t),s)\frac{be^{-i\omega_{c}s}}{2ik_{c}\psi^{*}(t)}
+t𝑑sG(qqp(t),s)beiωcs2ikc(ψ(t)1ψ(s)1)\displaystyle\qquad\qquad+\int_{-\infty}^{t}ds\;G(q-q_{p}(t),s)\frac{be^{-i\omega_{c}s}}{2ik_{c}}(\psi^{*}(t)^{-1}-\psi^{*}(s)^{-1})
+t𝑑s(G(qqp(t),s)G(qqp(s),s))beiωcs2ikcψ(s)\displaystyle\qquad\qquad+\int_{-\infty}^{t}ds\;\left(G(q-q_{p}(t),s)-G(q-q_{p}(s),s)\right)\frac{be^{-i\omega_{c}s}}{2ik_{c}\psi^{*}(s)}
=:ϕwav,t+ε1(q,t)+ε2(q,t),\displaystyle=:\phi_{\text{wav},t}+\varepsilon_{1}(q,t)+\varepsilon_{2}(q,t),

where the fixed-time solution ϕwav,t\phi_{\text{wav},t} is defined by (13). The problem thus reduces to bounding ε1\varepsilon_{1} and ε2\varepsilon_{2}.

Recall from the derivation of the trajectory equation that the particle velocity is quickly dampened to the Bohmian form; in particular, we can bound u/c=O(U/c)u/c=O(U/c) with a velocity scale UU dependent only on its initial conditions q˙p,0\dot{q}_{p,0} and the Bohmian data |ψ0(qp,0)||\psi_{0}(q_{p,0})|, |ψ0(qp,0)||\nabla\psi_{0}(q_{p,0})|. With this in mind, we can choose a time

τ=τ(|ψ0(qp,0)|,|ψ0(qp,0)|,|q˙p,0|)\tau=\tau(|\psi_{0}(q_{p,0})|,|\nabla\psi_{0}(q_{p,0})|,|\dot{q}_{p,0}|)

such that the particle remains bounded away from the zero locus of ψ\psi for t[0,τ]t\in[0,\tau]; say that α<|ψ(qp)|<α1\alpha<|\psi(q_{p})|<\alpha^{-1} for this interval, for a fixed value α\alpha. Because of the zero Dirichlet conditions, this ensures that qpq_{p} remain bounded away from Ω\partial\Omega for tτt\leq\tau.

Reducing τ\tau and increasing UU as necessary, we similarly estimate |ψ˙(qp)|=O((mU2/)α1)|\dot{\psi}(q_{p})|=O((mU^{2}/\hbar)\alpha^{-1}) over this interval; this is simply the Schrödinger energy scale EmU2E\sim mU^{2}, encoded in our non-relativistic hypothesis, put together with our estimate |ψ(qp)|<α1|\psi(q_{p})|<\alpha^{-1}.

We translate and nondimensionalise the integrand of (14) as s=ωc(st)s^{\prime}=\omega_{c}(s-t), as tc=1/ωct_{c}=1/\omega_{c} is the dominant timescale in the Green’s function. Then we expand ψ\psi about tt as

ψ(s)=ψ(t)+sdsψ+O((s)2(U2/c2)2α1),\psi(s)=\psi(t)+s^{\prime}d_{s^{\prime}}\psi+O((s^{\prime})^{2}(U^{2}/c^{2})^{2}\alpha^{-1}),

or equivalently,

ψ(s)1=ψ(t)1sψ2dsψ+O((s)2(U2/c2)2α1).\psi(s)^{-1}=\psi(t)^{-1}-s^{\prime}\psi^{-2}d_{s^{\prime}}\psi+O((s^{\prime})^{2}(U^{2}/c^{2})^{2}\alpha^{-1}).

With this in mind, we can refine our estimate of ε1\varepsilon_{1} as follows; writing r=qqcr=\|q-q_{c}\|, we find

ε1\displaystyle\varepsilon_{1} =0𝑑sG(qqp(t),s)beis2ikcωc(ψ(t)1ψ(s)1)\displaystyle=\int_{-\infty}^{0}ds^{\prime}\;G(q-q_{p}(t),s)\frac{be^{-is^{\prime}}}{2ik_{c}\omega_{c}}(\psi^{*}(t)^{-1}-\psi^{*}(s)^{-1})
=0𝑑sG(qqp(t),s)beis2ikcωcsψ2dsψ+O(α1U4/c4)\displaystyle=-\int_{-\infty}^{0}ds^{\prime}\;G(q-q_{p}(t),s)\frac{be^{-is^{\prime}}}{2ik_{c}\omega_{c}}s^{\prime}\psi^{-2}d_{s^{\prime}}\psi+O(\alpha^{-1}U^{4}/c^{4})
=bkcωc2π0𝑑sθ(ctqqp(t))J1(ωct2r2/c2)ωct2r2/c2eis2ikcωcsψ2dsψ\displaystyle=\frac{bk_{c}\omega_{c}}{2\pi}\int_{-\infty}^{0}ds^{\prime}\;\theta(ct-\|q-q_{p}(t)\|)\frac{J_{1}(\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}})}{\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}}}\frac{e^{-is^{\prime}}}{2ik_{c}\omega_{c}}s^{\prime}\psi^{-2}d_{s^{\prime}}\psi
kcωcr4πrbeikcrdsψ2ikcωcψ2+O(α1U4/c4)\displaystyle\qquad\qquad-\frac{k_{c}\omega_{c}r}{4\pi r}\cdot\frac{be^{-ik_{c}r}d_{s^{\prime}}\psi}{2ik_{c}\omega_{c}\psi^{2}}+O(\alpha^{-1}U^{4}/c^{4})
=b0𝑑sθ(ctqqp(t))J1(ωct2r2/c2)ωct2r2/c2eis4πisψ2dsψ+O(α1U2/c2)\displaystyle=b\int_{-\infty}^{0}ds^{\prime}\;\theta(ct-\|q-q_{p}(t)\|)\frac{J_{1}(\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}})}{\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}}}\frac{e^{-is^{\prime}}}{4\pi i}s^{\prime}\psi^{-2}d_{s^{\prime}}\psi+O(\alpha^{-1}U^{2}/c^{2})

Finally, introducing r:=kcrr^{\prime}:=k_{c}r allows us to calculate

0𝑑sθ(ctq)\displaystyle\int_{-\infty}^{0}ds^{\prime}\;\theta(ct-\|q\|) J1(ωct2r2/c2)ωct2r2/c2eis4πis\displaystyle\frac{J_{1}(\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}})}{\omega_{c}\sqrt{t^{2}-r^{2}/c^{2}}}\frac{e^{-is^{\prime}}}{4\pi i}s^{\prime}
=0𝑑sθ(sr)J1((s)2(r)2)(s)2(r)2eis4πis=O(1),\displaystyle=\int_{-\infty}^{0}ds^{\prime}\;\theta(s^{\prime}-r^{\prime})\frac{J_{1}(\sqrt{(s^{\prime})^{2}-(r^{\prime})^{2}})}{\sqrt{(s^{\prime})^{2}-(r^{\prime})^{2}}}\frac{e^{-is^{\prime}}}{4\pi i}s^{\prime}=O(1),

as the integral itself converges—to see this, note that the Bessel function decreases as J1(x)=O(x1/2)J_{1}(x)=O(x^{-1/2}). But then our estimate for ψ1\psi^{-1} gives

ε1=O(α1U2/c2),\varepsilon_{1}=O(\alpha^{-1}U^{2}/c^{2}),

which tells us that feedback from ϕwav\phi_{\text{wav}} on ψ\psi vanishes in the non-relativistic limit. A similar argument demonstrates that ε2=O(α1U/c)\varepsilon_{2}=O(\alpha^{-1}U/c).

In full, these estimates on ε1\varepsilon_{1} and ε2\varepsilon_{2} demonstrate that the error in (12) is at most O(U/c)O(U/c) in the non-relativistic limit, and thus that our ansatz (7) for ϕwav\phi_{\text{wav}} exactly cancels out the nonlinear wave forcing in this limit. The remaining component ψeiωct\psi e^{-i\omega_{c}t} thus solves a homogeneous Klein–Gordon equation, and a standard reduction of the resulting unforced Klein–Gordon equation shows that ψ\psi converges to a solution of the free Schrödinger equation.

Putting this field result together with our derivation of the trajectory equation above, we recover the claimed short-time convergence; the long-time convergence follows as discussed at the beginning of this section.

Note on nonzero potentials.

As it stands, our result applies to free systems (i.e., V(q)0V(q)\equiv 0) and to systems with rigid walls (i.e., V(qΩc)=+V(q\in\Omega^{c})=+\infty). For instance, this encompasses several systems previously investigated in the case of walking droplets: the quantum corral (or quantum billiards in general) [27], slit diffraction apparatuses [1, 21, 22], and systems involving sharp walls and beam splitters [28, 31].

To adapt our result to the case of nonzero potentials, we modify the Klein–Gordon equation as

(μμ+(kc+V(q)/c)2)ϕ=δ3(qqp)bγ12ikcϕ¯.\left(\partial^{\mu}\partial_{\mu}+(k_{c}+V(q)/\hbar c)^{2}\right)\phi=-\delta^{3}(q-q_{p})\frac{b\gamma^{-1}}{2ik_{c}{\overline{\phi}}^{*}}.

Since the contributions of V(x)V(x) appear only at an order U2/c2U^{2}/c^{2} in this equation, the form of ϕwav\phi_{\text{wav}} does not change from the free case considered here. Moreover, since the Klein–Gordon Green’s function changes only at the same relative order, our earlier estimate on ε1\varepsilon_{1} and ε2\varepsilon_{2} holds as stated, and our result goes through unaffected.

Note on the choice of σ\sigma.

Different expressions for σ\sigma yield different transient behaviours according to their analogues for (10). As an example, note that the piecewise sawtooth wave

σ(θ)={θ2π0<θπθ+2ππ<θ2π\sigma(\theta)=\begin{cases}\theta-2\pi&0<\theta\leq\pi\\ \theta+2\pi&\pi<\theta\leq 2\pi\end{cases}

periodically changes the sign of σ/b\sigma/b in (10), forcing uflucu_{\text{fluc}} to jump periodically. In this way, a different choice of σ\sigma could excite Compton-scale oscillations, better matching the behaviour of the limit studied by Darrow and Bush [5]. We conjecture that ψ\psi still converges to the Schrödinger wavefunction with such a coupling, and that the particle oscillates rapidly about a mean Bohmian trajectory.

4 Analogue Position Measurements

We have demonstrated how, with an appropriate choice of particle-field coupling, the relativistic model of Darrow and Bush [5] reduces to traditional Bohmian mechanics in the non-relativistic limit. In this limit, the particle continues to carry a Compton-scale wavepacket, as noted by Darrow and Bush, but this wavepacket no longer radiates energy into the rest of the field.

We can force the system to deviate from Bohmian mechanics by introducing a thermodynamic element to the system’s evolution, however. In this section, we show how coupling the wave field to a set of heat sinks gives rise to a compelling analogue of wavefunction collapse. Along with greater insight into the energetics of the present pilot-wave system, this argument suggests how an experiment with, e.g., the walking droplet system of Couder and Fort [1] might give a macroscopic analogue of quantum measurements.

We model a position measurement as follows:

  1. 1.

    We fix a partition of the domain Ω3\Omega\subset\mathbb{R}^{3} into open cells UαΩU_{\alpha}\subset\Omega. During the course of measurement, we enforce that the particle cannot cross from one grid cell to the next, and that field dynamics are decoupled between neighbouring cells.

  2. 2.

    We let the pilot-wave system undergo an energy-minimising flow in each cell, corresponding to a heat sink attached to each cell.

  3. 3.

    Post-measurement, we lift the partition and allow the system to continue evolving according to (6).

The above process is depicted in Fig. 1.


Remark.

Our analogue of position measurement differs drastically from the usual quantum formalisms, such as that of von Neumann [32]. The primary reason for this is, our results apply only for a single particle222In fact, a several-particle framework would require nonlocality in order to converge to Bohmian mechanics, as per Bell’s theorem [16]. This would sacrifice a key similarity with classical pilot-wave systems, so it is not of present interest., so we cannot add new degrees of freedom for the measurement apparatus. As such, we attempt only to model the energy deposition by a single particle as it is measured.


To make the measurement process precise, we first derive an expression for the cell-averaged energy.

{tikzpicture}\node

[anchor=south west,inner sep=0] (image) at (0,1) Refer to caption; \node[anchor=south west,inner sep=0] (image1) at (4.5,1) Refer to caption; \node[anchor=south west,inner sep=0] (image2) at (9,1) Refer to caption; \node[text width=2cm] at (4.9,2) \rightsquigarrow; \node[text width=2cm] at (9.4,2) \rightsquigarrow;

Figure 1: A depiction of the analogue wave collapse we find in our de Broglie-like pilot wave system. In short, we split the domain into measured regions Uα3U_{\alpha}\subset\mathbb{R}^{3}, and we couple each region to a distinct heat sink; this process analogises the energy deposition of a particle into a measurement device. We show that, during such a process, the wave field ψ\psi collapses exactly to its projection in the cell containing the particle. We note that there is no conflict with the convergence to Bohmian mechanics we argued in Section 3; the presence of heat sinks violates the assumptions of those results, at least during the measurement process.

Deriving a local mean energy.

We write

ψwav:=18πikcψ(qp)qqp,\psi_{\text{wav}}:=-\frac{1}{8\pi ik_{c}\psi^{*}(q_{p})\|q-q_{p}\|},

so that the total field takes the form

ϕ=(ψ+ψwav)eiωct=:ψtoteiωct.\phi=(\psi+\psi_{\text{wav}})e^{-i\omega_{c}t}=:\psi_{\text{tot}}e^{-i\omega_{c}t}.

Recall that the function ψ\psi solves the homogeneous Schrödinger equation in the non-relativistic limit. The Klein–Gordon energy density takes the form

ρE(ϕ)=mkc|tϕ|2+mc2kcϕ2+mc2kc3|ϕ|2,\rho_{E}(\phi)=mk_{c}|\partial_{t}\phi|^{2}+mc^{2}k_{c}\|\nabla\phi\|^{2}+mc^{2}k_{c}^{3}|\phi|^{2},

which we write in terms of ψ\psi as

ρE\displaystyle\rho_{E} =mkc|tψtotiωcψtot|2+mc2kc3(|ψtot|2|ψwav|2)\displaystyle=mk_{c}|\partial_{t}\psi_{\text{tot}}-i\omega_{c}\psi_{\text{tot}}|^{2}+mc^{2}k_{c}^{3}\left(|\psi_{\text{tot}}|^{2}-|\psi_{\text{wav}}|^{2}\right)
+mc2kc(ψtot2ψwav2)+ρs/|ψ(qp)|2.\displaystyle\qquad+mc^{2}k_{c}\left(\|\nabla\psi_{\text{tot}}\|^{2}-\|\nabla\psi_{\text{wav}}\|^{2}\right)+\rho_{s}/|\psi(q_{p})|^{2}.

Here, we have grouped all terms proportional to 1/|ψ(qp)|21/|\psi(q_{p})|^{2} into the renormalised “singular energy density” ρs\rho_{s}; the magnitude of this energy will not affect particle dynamics post-measurement.

When we integrate over the full cell, cross terms of the form ψ2ψwav\psi\nabla^{2}\psi^{*}_{\text{wav}} (and thus ψwav2ψ\psi^{*}_{\text{wav}}\nabla^{2}\psi, integrating by parts) can be removed, as the expression

𝑑Vψ2ψwav\displaystyle\int dV\;\psi\nabla^{2}\psi^{*}_{\text{wav}} 𝑑Vψψ(qp)δ3(qqp)=1\displaystyle\propto\int dV\;\frac{\psi}{\psi(q_{p})}\delta^{3}(q-q_{p})=1

is independent of ψ\psi; it corresponds only to an overall shift of energy. Discarding these terms, as well as terms of order O(U4/c4)O(U^{4}/c^{4}), we find a total energy

E\displaystyle E =2mc2kc𝑑V(kc2|ψ|2+ψ2)+ρsV/|ψ(qp)|2\displaystyle=2mc^{2}k_{c}\int dV\;\left(k_{c}^{2}|\psi|^{2}+\|\nabla\psi\|^{2}\right)+\rho_{s}V/|\psi(q_{p})|^{2}
+𝑑Vmωc4πqqp(Reψtψ(qp)ψ(qp)2Imωcψψ(qp))\displaystyle\qquad+\int dV\;\frac{m\omega_{c}}{4\pi\|q-q_{p}\|}\left(\operatorname{Re}\frac{\psi\partial_{t}\psi(q_{p})}{\psi(q_{p})^{2}}-\operatorname{Im}\frac{\omega_{c}\psi}{\psi(q_{p})}\right)

in the cell containing the particle, defining the cell volume V=𝑑V 1V=\int dV\;1. The final term on the right-hand side approximately vanishes in a short-time average; we can see this rigorously if we assume that the constant contribution of ψ\psi dominates higher-energy terms, or that ψ\psi is nearly monochromatic. This leaves us to minimise

E~:=2mc2kc𝑑V(kc2|ψ|2+ψ2)+ρsV/|ψ(qp)|2.\displaystyle\tilde{E}:=2mc^{2}k_{c}\int dV\;\left(k_{c}^{2}|\psi|^{2}+\|\nabla\psi\|^{2}\right)+\rho_{s}V/|\psi(q_{p})|^{2}. (15)

To continue, recall that the Bohmian particle (which ours converges to) obeys Born statistics, as

P(qp=q)=|ψ(q)|2V1𝑑V|ψ|2.P(q_{p}=q)=\frac{|\psi(q)|^{2}}{V^{-1}\int dV\;|\psi|^{2}}.

Weighting the mean energy (15) by this distribution, we find a final estimate for the energy in each cell UαΩU_{\alpha}\subset\Omega:

E¯|Uα:=E~Uα\displaystyle{\overline{E}}|_{U_{\alpha}}:=\langle\tilde{E}\rangle_{U_{\alpha}} =2mc2kc𝑑V(kc2|ψ|2+ψ2)+𝟏qpUα𝑑VρsV2|ψ(q)|2|ψ(q)|2𝑑V|ψ|2\displaystyle=2mc^{2}k_{c}\int dV\;\left(k_{c}^{2}|\psi|^{2}+\|\nabla\psi\|^{2}\right)+\mathbf{1}_{q_{p}\in U_{\alpha}}\int dV^{\prime}\frac{\rho_{s}V^{2}}{|\psi(q^{\prime})|^{2}}\,\frac{|\psi(q^{\prime})|^{2}}{\int dV\;|\psi|^{2}} (16)
=2mc2kc𝑑V(kc2|ψ|2+ψ2)+𝟏qpUαρsV2𝑑V|ψ|2.\displaystyle=2mc^{2}k_{c}\int dV\;\left(k_{c}^{2}|\psi|^{2}+\|\nabla\psi\|^{2}\right)+\mathbf{1}_{q_{p}\in U_{\alpha}}\,\frac{\rho_{s}V^{2}}{\int dV\;|\psi|^{2}}.

Here, the notation 𝟏qpUα\mathbf{1}_{q_{p}\in U_{\alpha}} indicates that the last term is present only if the particle is contained in the cell UαU_{\alpha}.

Analogue measurements.

Now that we have the energy estimate (16) for each cell, we can define our analogue measurement appropriately:


Definition 1.

Fix a configuration space Ω3\Omega\subseteq\mathbb{R}^{3}, and suppose that (qp,ψ,ϕwav)(q_{p},\psi,\phi_{\text{wav}}) evolve according to the non-relativistic dynamics of Section 3.

A position measurement in this system is (a) a collection UαΩU_{\alpha}\subset\Omega of disjoint open sets such that ΩαU¯α\Omega\subset\bigcup_{\alpha}{\overline{U}}_{\alpha} and (b) a mean-energy-minimising flow in each cell

ψ|Uαψt|Uα,dtψt|UαψE¯(ψt)|Uα,\psi|_{U_{\alpha}}\mapsto\psi_{t}|_{U_{\alpha}},\qquad d_{t}\psi_{t}|_{U_{\alpha}}\propto-\nabla_{\psi}{\overline{E}}(\psi_{t})|_{U_{\alpha}},

with E¯{\overline{E}} as defined in (16).


Importantly, this is a dynamical process, and we will see that the system behaviour depends on whether such a process runs for a short or long time.

The first variation of this quantity, evaluated against a function hL2(Uα)h\in L^{2}(U_{\alpha}), is given by

δψE¯(h)|Uα\displaystyle\delta_{\psi}{\overline{E}}(h)|_{U_{\alpha}} =2mc2kc𝑑V(kc2ψ2ψ)h\displaystyle=2mc^{2}k_{c}\int dV\;\left(k_{c}^{2}\psi^{*}-\nabla^{2}\psi^{*}\right)h
ρsV2𝟏qpUα[𝑑V|ψ|2]2𝑑Vψh,\displaystyle\qquad-\frac{\rho_{s}V^{2}\mathbf{1}_{q_{p}\in U_{\alpha}}}{\left[\int dV\;|\psi|^{2}\right]^{2}}\int dV\;\psi^{*}h,

which identifies the energy gradient as

ψE¯|Uα\displaystyle\nabla_{\psi}{\overline{E}}|_{U_{\alpha}} =(2mc2kc3ρsV2𝟏qpUα[𝑑V|ψ|2]2)ψ2mc2kc2ψ.\displaystyle=\left(2mc^{2}k_{c}^{3}-\frac{\rho_{s}V^{2}\mathbf{1}_{q_{p}\in U_{\alpha}}}{\left[\int dV\;|\psi|^{2}\right]^{2}}\right)\psi-2mc^{2}k_{c}\nabla^{2}\psi. (17)

Now, consider an energy-minimising flow of the form dtψ=κ~ψE¯d_{t}\psi=-\tilde{\kappa}\nabla_{\psi}{\overline{E}}, as in Definition 1, with κ~\tilde{\kappa} the constant of proportionality. The final term of (17) is familiar from the imaginary-time Schrödinger equation. Identifying coefficients with the Schrödinger equation suggests writing

κ~=κ4m2c2kc,\tilde{\kappa}=\frac{\kappa\hbar}{4m^{2}c^{2}k_{c}},

where κ>0\kappa>0 is a dimensionless parameter. Then we find

dtψt=κ(ωc2(ωs/2)𝟏qpUαV2[𝑑V|ψ|2]2)ψ+κ2m2ψ,\displaystyle d_{t}\psi_{t}=-\kappa\left(\frac{\omega_{c}}{2}-\frac{(\omega_{s}/2)\mathbf{1}_{q_{p}\in U_{\alpha}}}{V^{-2}\left[\int dV\;|\psi|^{2}\right]^{2}}\right)\psi+\kappa\frac{\hbar}{2m}\nabla^{2}\psi, (18)

where ωs=ρs/4m2c2kc\omega_{s}=\rho_{s}\hbar/4m^{2}c^{2}k_{c} is a system-dependent frequency. If we renormalise ρs\rho_{s} by assuming a finite radius O(1/kc)O(1/k_{c}) for the particle, then ωs=O(ωc)\omega_{s}=O(\omega_{c}).

The two terms on the right-hand side of (18) accomplish different tasks, and at different timescales. Over a long timescale O(κ1τdB)O(\kappa^{-1}\tau_{dB}), where τdB=λdB/c\tau_{dB}=\lambda_{dB}/c, the final term relaxes the system into the Schrödinger ground state. This is the expected behaviour of a thermal sink, so it is not surprising that we recover this relaxation.

More importantly, the first term in (18) adjusts the normalisation of the wavefunction. Specifically, over the fast timescale O(κ1τc)=O(κ1τdB(U/c)2)O(\kappa^{-1}\tau_{c})=O(\kappa^{-1}\tau_{dB}(U/c)^{2}), the first term scales the wavefunction without modification until

V1𝑑V|ψ|2=2mc2kc3ρs=O(1)V^{-1}\int dV\;|\psi|^{2}=\sqrt{\frac{2mc^{2}k_{c}^{3}}{\rho_{s}}}=O(1)

in the cell containing the particle. However, any cells UαU_{\alpha} for which qpUαq_{p}\notin U_{\alpha} admit the simpler relaxation

dtψt|Uα∌qp=κωc2ψ+κ2m2ψ,d_{t}\psi_{t}|_{U_{\alpha}\not\ni q_{p}}=-\kappa\frac{\omega_{c}}{2}\psi+\kappa\frac{\hbar}{2m}\nabla^{2}\psi,

which reduces the norm of the wavefunction to zero over the same fast timescale κ1τc\kappa^{-1}\tau_{c}. This corresponds to the phenomenon of wavefunction collapse in quantum theory: the wavefunction is projected to its component in the cell containing the particle. Since this projection does not affect the particle dynamics in its own cell, the particle is able to evolve post-measurement according to the Bohmian mechanics derived in the preceding section.

5 Discussion

There are several strong similarities between Bohmian mechanics and de Broglie’s 1927 proposal [14]; in both, a pilot-wave guides a point-like particle, and Born’s probability rule emerges dynamically. Even still, de Broglie eventually decided that Bohmian mechanics deviated too far from his original picture; he believed that the underlying theory should be relativistic, that the pilot-wave should be a “physical one… which cannot be arbitrarily normed, and which is distinct from the [Schrödinger wavefunction],” and that the particle must “form on [the pilot-wave] a small region of high energy concentration, which may be likened… to a moving singularity” [33].

While de Broglie’s double-solution program was never able to fully explain quantum mechanics, it provided a striking physical picture that presaged some of the classical pilot-wave systems studied today. These systems—such as the “walking droplet” system of Couder and Fort [1] and its various abstractions [2, 3, 4, 5]—give rise to a variety of novel, classical behaviours, and they offer dynamical analogues of several key results from quantum theory [20]. However, it remains unclear whether such systems could replicate even single-particle quantum predictions completely.

Here, we have answered the latter question in the affirmative. Specifically, we have re-examined the de Broglie-like pilot-wave framework of Darrow and Bush [5], in the case in which a point-particle is coupled to the phase of a Klein–Gordon field. We have here demonstrated that its non-relativistic limit reduces exactly to Bohmian mechanics, in a highly general setting. In so doing, we see that de Broglie’s double-solution theory—modernised and formalised in the system studied by Darrow and Bush—can be connected to Bohmian mechanics using a single Lagrangian framework. Moreover, by pairing this system with an abstracted version of a measuring apparatus, we have shown that it can give rise to a compelling analogue of wavefunction collapse.

The present work further demonstrates the richness of the classical field theory proposed by Darrow and Bush [5], its potential for connecting different classes of pilot-wave systems, and its potential for capturing new classical quantum analogues.

\bmhead

Acknowledgments

I would like to express my sincere gratitude to Professor John Bush (MIT) for his invaluable guidance, support, and encouragement. His constructive dissent has consistently played a crucial role in shaping the direction of my research.

Declarations

Funding

The authors did not receive support from any organization for the submitted work.

Competing Interests

The authors have no financial or proprietary interests in any material discussed in this article.

References

  • \bibcommenthead
  • Couder and Fort [2006] Couder, Y., Fort, E.: Single-particle diffraction and interference at a macroscopic scale. Phys. Rev. Lett. 97, 154101 (2006) https://doi.org/10.1103/PhysRevLett.97.154101
  • Durey and Bush [2021] Durey, M., Bush, J.: Classical pilot-wave dynamics: The free particle. Chaos 31, 033136 (2021) https://doi.org/10.1063/5.0039975
  • Dagan and Bush [2020] Dagan, Y., Bush, J.: Hydrodynamic quantum field theory: the free particle. CR Mecanique 348, 555–571 (2020) https://doi.org/10.5802/crmeca.34
  • Durey and Bush [2020] Durey, M., Bush, J.: Hydrodynamic quantum field theory: The onset of particle motion and the form of the pilot wave. Front. Phys. 8, 300 (2020) https://doi.org/10.3389/fphy.2020.00300
  • Darrow and Bush [2024] Darrow, D., Bush, J.W.M.: Revisiting de Broglie’s double-solution pilot-wave theory with a Lorentz-covariant Lagrangian framework. Symmetry 16(2) (2024)
  • de Broglie [1956] Broglie, L.: Une Tentative D’interprétation Causale et Nonlinéaire de la Mécanique Ondulatoire: la Théorie de la Double solution. Gautier-Villars, Paris (1956)
  • de Broglie [1970] Broglie, L.: The reinterpretation of wave mechanics. Found. Phys. 1(1), 5–15 (1970)
  • de Broglie [1925] de Broglie, L.: Recherches sur la théorie des quanta. Ann. Phys. 10(3), 22–128 (1925) https://doi.org/10.1051/anphys/192510030022
  • Bacciagaluppi and Valentini [2009] Bacciagaluppi, G., Valentini, A.: Quantum Theory at the Crossroads: Reconsidering the 1927 Solvay Conference. Cambridge University Press, ??? (2009)
  • Bohm [1952a] Bohm, D.: A suggested interpretation of the quantum theory in terms of hidden variables, I. Phys. Rev. 85, 66–179 (1952)
  • Bohm [1952b] Bohm, D.: A suggested interpretation of the quantum theory in terms of hidden variables, II. Phys. Rev. 85, 180–193 (1952)
  • von Neumann and Beyer [2018] Neumann, J., Beyer, R.T.: Mathematical Foundations of Quantum Mechanics: New Edition. Princeton University Press, Princeton, NJ (2018). http://www.jstor.org/stable/j.ctt1wq8zhp Accessed 2024-01-11
  • Holland [1993] Holland, P.R.: The Quantum Theory of Motion: An Account of the de Broglie-Bohm Causal Interpretation of Quantum Mechanics. Cambridge University Press, Cambridge, U.K. (1993)
  • de Broglie [1952] Broglie, L.: La physique quantique restera-t-elle indéterministe? Rev. Hist. Sci. 5(4), 289–311 (1952). Accessed 2024-03-08
  • Bell [1966] Bell, J.S.: On the problem of hidden variables in quantum mechanics. Rev. Mod. Phys. 38, 447–452 (1966) https://doi.org/10.1103/RevModPhys.38.447
  • Bell [1964] Bell, J.S.: On the Einstein–Podolsky–Rosen paradox. Phys. Phys. Fiz. 1, 195–200 (1964) https://doi.org/10.1103/PhysicsPhysiqueFizika.1.195
  • Aspect et al. [1982] Aspect, A., Dalibard, J., Roger, G.: Experimental test of Bell’s inequalities using time-varying analyzers. Phys. Rev. Lett. 49, 1804–1807 (1982) https://doi.org/10.1103/PhysRevLett.49.1804
  • Sutherland [2019a] Sutherland, R.: Incorporating action and reaction into a particle interpretation for quantum mechanics – Dirac case (2019)
  • Sutherland [2019b] Sutherland, R.: Incorporating action and reaction into a particle interpretation for quantum mechanics – Schrödinger case (2019)
  • Bush and Oza [2021] Bush, J.W.M., Oza, A.U.: Hydrodynamic quantum analogs. Rep. Prog. Phys. 84(1) (2021)
  • Pucci et al. [2018] Pucci, G., Harris, D.M., Faria, L.M., Bush, J.W.M.: Walking droplets interacting with single and double slits. J. Fluid Mech. 835, 1136–1156 (2018)
  • Ellegaard and Levinsen [2020] Ellegaard, C., Levinsen, M.T.: Interaction of wave-driven particles with slit structures. Physical Review E 102(2), 023115 (2020) https://doi.org/10.1103/PhysRevE.102.023115 . Accessed 2021-08-27
  • Fort et al. [2010] Fort, E., Eddi, A., Boudaoud, A., Moukhtar, J., Couder, Y.: Path-memory induced quantization of classical orbits. P. Natl. Acad. Sci. USA 107 (2010) https://doi.org/10.1073/pnas.1007386107
  • Harris and Bush [2014] Harris, D.M., Bush, J.W.M.: Droplets walking in a rotating frame: from quantized orbits to multimodal statistics. J. Fluid Mech. 739, 444–464 (2014) https://doi.org/10.1017/jfm.2013.627
  • Perrard et al. [2014] 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) https://doi.org/10.1038/ncomms4219
  • Oza et al. [2018] Oza, A., Rosales, R., Bush, J.W.M.: Hydrodynamic spin states. Chaos: An Interdisciplinary Journal of Nonlinear Science 28, 096106 (2018) https://doi.org/10.1063/1.5034134
  • Harris et al. [2013] Harris, D.M., Moukhtar, J., Fort, E., Couder, Y., Bush, J.W.M.: Wavelike statistics from pilot-wave dynamics in a circular corral. Phys. Rev. E 88, 011001 (2013) https://doi.org/10.1103/PhysRevE.88.011001
  • Frumkin et al. [2022] Frumkin, V., Darrow, D., Bush, J.W.M., Struyve, W.: Real surreal trajectories in pilot-wave hydrodynamics. Phys. Rev. A 106, 010203 (2022) https://doi.org/10.1103/PhysRevA.106.L010203
  • Schoene [1979] Schoene, A.Y.: On the nonrelativistic limits of the Klein–Gordon and Dirac equations. Journal of Mathematical Analysis and Applications 71(1), 36–47 (1979) https://doi.org/10.1016/0022-247X(79)90216-6
  • [30] Green’s function for the Klein–Gordon equation. Wolfram Research (Accessed 2023). functions.wolfram.com/Bessel-TypeFunctions/BesselJ/31/02/
  • Frumkin and Bush [2023] Frumkin, V., Bush, J.W.M.: Misinference of interaction-free measurement from a classical system. Phys. Rev. A 108, 060201 (2023) https://doi.org/10.1103/PhysRevA.108.L060201
  • Mello [2014] Mello, P.A.: The von Neumann model of measurement in quantum mechanics. In: AIP Conference Proceedings. AIP Publishing LLC, Melville, NY (2014). https://doi.org/10.1063/1.4861702 . http://dx.doi.org/10.1063/1.4861702
  • de Broglie [1987] de Broglie, L.: Interpretation of quantum mechanics by the double solution theory. Ann. Fond. Louis de Broglie 12, 1–23 (1987)