Figure 1.
(
a) A depiction of the free particle in our system, in two dimensions and with the coupling constant
. In this simulation, we accelerate the particle from rest to a velocity
. Radiation is emitted from the point of acceleration, corresponding to the form predicted in
Section 3.4. Axes are given in units of
. (
b) The wave form adjoining the particle, as predicted in
Section 3.2, is visible as a high-amplitude region around the particle, of characteristic radius corresponding to the Compton scale
. As predicted in
Section 3.4, the local wave field has a characteristic wavelength
, where
v is the
instantaneous (rather than initial) speed of the particle. (
c) The same simulation at a later time. Because the wave travels out from the particle’s point of origin, the local curvature of the wavefront decreases as the particle moves forward. Because the domain is periodic in both directions, the wave field grows complex, and the particle experiences the radiation from its periodic images.
3.1. Conserved Currents in the AM System
We now consider the non-conservative derivation of the AM system detailed in
Appendix B.1. Although it does not fit into our general framework (
1), it allows us to recover two key conservation laws. The first and most important of these conservation laws is that of the
stress-energy tensor , which encodes system energy and linear momentum in a Lorentz-covariant 2-tensor. In a given reference frame,
can be identified with the system energy density, and similarly
with the momentum density. The remaining terms represent fluxes of these quantities, such that, in the absence of sources or sinks, we have
We are able to recover a stress-energy conservation law by applying Lemma 2 to spacetime translations. The full derivation is in
Appendix A.2. We define the stress-energy tensor
which is exactly the sum of a relativistic free particle and a free scalar field. Then we recover the balance equations
This demonstrates the key benefit of our non-conservative derivation: the momentum conservation is exact, and the energy balance is neatly encoded by the material derivative
of the field along the particle trajectory.
We can find a similar conservation law for the
relativistic angular momentum by examining spatial rotations and Lorentz boosts. Here, the spatial components
form the classical angular momentum bivector:
where
is the linear momentum of (
9). The temporal components
are somewhat less useful; we have
which represents a scaled centre-of-mass value.
Instead of applying Noether’s theorem, we deduce this conservation law more easily by leveraging (
10). We define the angular momentum currents
which gives the angular momentum
. Plugging in the balance Equation (
10) and using the symmetry of
, we recover
where
and
for
. For the spatial components
, which represent the classical angular momentum current, we deduce a true conservation law
We expect angular momentum conservation to play an important role in bound or orbiting states, which are outside the scope of this work. In the context of the linear acceleration of the free particle, this conservation law requires that the outgoing radiation have vanishing total angular momentum.
3.2. Steady States of the Free Particle, and the Local Wavepacket
As a first step towards identifying the local form of the wave field, note that one steady-state solution of (
7) and (
8) is
This corresponds to a Yukawa potential of range
or, in dimensional form, the Compton wavelength
.
Now consider a general trajectory
, and recall the (position-space) Green’s function for the forced Klein–Gordon equation [
48]:
with
the Heaviside step function and
a Bessel function of the first kind. We integrate this expression over the particle trajectory one term at a time, first defining
This term reduces to
where the sum is taken over times
such that
. Since the particle is traveling strictly slower than
, however, it can only cross this locus once—say, at
—and the sum reduces further to
Suppose we are in the instantaneous rest frame of the particle, so that
for some acceleration amplitude
. For
within the ball
,
, we find that
, and thus
, yielding
With this result in hand, we reduce the total field to the local expression (
11); using the expression (
12) and subtracting the components giving (
11), we find
using the expansion
. In particular, the expression (
11) holds in a neighborhood of the particle in the particle’s instantaneous frame of reference, up to a finite contribution of amplitude
. By performing a Lorentz boost in the
direction, we recover the more general form
for the wavepacket adjoining a particle of velocity
. As we discuss further in
Appendix B.3, this extension follows exactly from the approximate Lorentz-covariance of the AM system.
In short, the analysis above demonstrates that the particle is
dressed with a trajectory-independent Yukawa potential, constant up to a length contraction. In
Section 3.5, we will demonstrate that this “wavepacket” modifies the particle’s effective inertial mass and momentum. We see a numerical depiction of this wavepacket in
Figure 1, albeit in two rather than three dimensions. There, the local wavepacket corresponds to the high-amplitude region of radius
centered on the particle.
Another important inference may be made by re-examining (
14). Consider again the rest frame of the particle at
, and suppose as before that the particle is accelerating as
. Then the second term in (
14) corresponds precisely to the
radiation created by the particle’s motion between time
and
; that is, the change in the scalar value of the field
away from its steady state (
15). Our calculation shows that this value is at most
, where
is the magnitude of the velocity change in this interval. If the particle is not accelerating at all (and so
), it does not radiate any waves outward: at a fixed velocity
, the particle carries
only the wavepacket (
15). If the particle acceleration has a magnitude
, as in the above analysis, the value of
at a later time is changed at most
at the rate. The resulting radiation becomes significant only if
, as will arise if the particle rapidly changes from one velocity state to another. This calculation will be substantiated in our numerical study of wave radiation in
Section 3.4.
3.3. Zitterbewegung: Particle Oscillation at the Compton Frequency
Spontaneous particle oscillations have been shown to arise in several models of classical pilot-wave systems. In-line speed oscillations have been reported to arise in several settings in the walking droplet systems, including the hydrodynamic analogue of Friedel oscillations [
32]. In-line oscillations with amplitude comparable to the wavelength of the pilot wave have been shown to be a robust feature of the
generalized pilot-wave framework (GPWF) [
37], a parametric generalization of the walking droplet system [
20]. Moreover, one-dimensional motion of the free particle in HQFT [
43,
44] is marked by erratic in-line oscillations at the Compton frequency. Notably, in all of these examples, particle oscillations are restricted to the in-line direction, and the coupling strength between particle and wave is a
periodic function of time.
We proceed by demonstrating that both features of de Broglie’s harmony of phases, an internal particle oscillation at frequency
and an accompanying wave of wavelength
, emerge naturally from the time-invariant dynamics (
7) and (
8). Moreover, the oscillation frequency and wavelength update dynamically as the particle’s momentum changes, in order to preserve the de Broglie relation
. Finally, the particle vibrates in
all directions when interacting with a wall-bounded geometry.
In our first series of tests, we start a particle at rest and accelerate it quickly to a speed
, from which it settles quickly into a steady speed
. We discuss the form of
further in
Section 3.5, and in particular, we derive a nonzero
virtual mass imparted to the particle’s rest mass by the surrounding wavefield.
For
and
, a spectrogram of the resulting in-line position oscillations is shown in
Figure 2. We highlight two noteworthy effects:
Initially, the particle undergoes in-line oscillations with frequency
in response to outgoing radiation from the point of acceleration, effectively surfing over its own radiative wave field. The form of this radiation is discussed in
Section 3.4;
After Compton periods, the particle oscillates with amplitude at frequencies between and . This is an artifact of our periodic domain, and specifically the particle interacting with the wave form generated by its periodic images. However, we expect the same effect to occur any time the particle interacts with a wall-bounded geometry, and its wave form reflects off the boundaries.
We quantify the first effect in
Figure 3a, where we repeat this experiment across a wide range of velocities and
b values. We see that the oscillation frequency conforms closely to
in all cases. In all simulations, the dominant frequency is constant until the particle encounters radiation from the opposite side of the domain.
Figure 2.
A spectrogram of in-line oscillations for our two-dimensional system, with coupling constant . The shown color values are normalized by , where is the oscillation magnitude. Here, we give the particle an initial velocity , which quickly relaxes to a mean velocity . For Compton periods, the particle undergoes an oscillation at the frequency . Thereafter, waves cover the entire periodic domain, and the particle vibrates at frequencies between and . Note, the diminishing intensity of the yellow line at reflects the temporal decay of the in-line Zitter.
Figure 2.
A spectrogram of in-line oscillations for our two-dimensional system, with coupling constant . The shown color values are normalized by , where is the oscillation magnitude. Here, we give the particle an initial velocity , which quickly relaxes to a mean velocity . For Compton periods, the particle undergoes an oscillation at the frequency . Thereafter, waves cover the entire periodic domain, and the particle vibrates at frequencies between and . Note, the diminishing intensity of the yellow line at reflects the temporal decay of the in-line Zitter.
Figure 3.
(
a) Dominant oscillation frequencies at the beginning of each trajectory, for particles across the range of initial velocities
to
. We observe that the particle oscillates at the frequency
independent of the coupling constant and velocity. (
b) Amplitudes of in-line oscillations in the long-time limit of
Figure 2, i.e., after waves have covered the entire periodic domain. Curves of the form
are shown for reference, where
and
are least-squares fits (reported in the Table (
c)).
Figure 3.
(
a) Dominant oscillation frequencies at the beginning of each trajectory, for particles across the range of initial velocities
to
. We observe that the particle oscillates at the frequency
independent of the coupling constant and velocity. (
b) Amplitudes of in-line oscillations in the long-time limit of
Figure 2, i.e., after waves have covered the entire periodic domain. Curves of the form
are shown for reference, where
and
are least-squares fits (reported in the Table (
c)).
Numerical Simulation of the AM System.
Our numerical code is based on that developed by Faria [
49] for walking droplets. We model the space as a two-dimensional periodic domain, allowing us to use high-accuracy pseudo-spectral methods to resolve the wave field. In turn, we evolve the wave field using a fourth-order Runge–Kutta method, separately tracking
and
in order to break (
3) into two first-order equations.
Our algorithm takes the following form. At each timestep, we add an approximate delta function—taken to be a Gaussian of variance
—to the field
at the current particle position
, scaled by
. Suppose
and
are the discrete Fourier transforms of
and
, respectively, so that, for instance,
. We evolve the fields
and
for one time-step according to the equations
and calculate the gradient
using the field’s Fourier expansion:
We use this to evolve
for one time-step, and we move the particle accordingly.
As we argue in
Appendix B.3, we can deduce the same behavior after a Lorentz transformation. Namely, oscillations at
arise regardless of initial velocity. If the particle is moving at a velocity
and accelerates to a steady velocity
, it begins to vibrate at the frequency
. This demonstrates a relativistically-correct internal clock at the Compton frequency, as in de Broglie’s original model, but with two distinctions. First, it emerges naturally from a time-invariant dynamics, without reference to intrinsic particle oscillation. Second, the emergent particle clock updates dynamically to adjust to the particle’s
current velocity, preserving the synchrony of de Broglie’s clock even under the influence of applied forces. We detail the wave form complementing this clock in the subsequent section.
3.4. A Dynamical Harmony of Phases
We proceed to examine the wave form generated following a particle acceleration, first for the case of acceleration from rest (see
Figure 1), then for a more general acceleration (see
Figure 4). We shall demonstrate that the resulting wave form naturally gives rise to the periodic forcing needed for the Zitterbewegung reported in the previous section. Specifically, we will show that the oscillation frequencies reported in
Section 3.3 arise from the particle being repeatedly washed over by quasi-monochromatic waves of wavelength
and phase velocity
. This analysis allows for a detailed comparison of the wave forms arising in pilot-wave hydrodynamics, in HQFT, and in our new model.
We first review some elementary properties of Klein–Gordon waves before returning to our particular system. Consider the dispersion relation of Klein–Gordon waves:
or in dimensional form,
. Here,
is the local oscillation frequency and
the local wavevector. The group velocity of the wave is
which corresponds precisely to the velocity of a point-particle of mass
m and momentum
, as can be seen by inverting the above relationship:
. The system is dispersive: the group velocity depends on wavelength. Thus, following a wave disturbance from a point source, different component wavelengths
travel outwards at different speeds
. The result is a wave train centered on the original source, with each excited wavenumber
k spreading outward at its corresponding group speed
. If a particle of mass
m is traveling away from this point-source at a velocity
, the region immediately adjoining the particle must then carry a wavelength
, where
is the particle’s momentum. This region grows in size over time—as wave crests of wavenumber
and
drift apart with velocity
—giving rise to a quasi-monochromatic wave form in the vicinity of the particle.
The above behavior occurs in our system when a particle is accelerated from rest (see
Figure 1). At the particle’s point of acceleration—the origin, in the case of
Figure 1—the particle excites a range of wavevectors
defined by
where
u is the new particle speed. The different components spread out from the origin in a spherical wave train according to (
16), and the particle surfs along the outgoing, origin-centered, spherical region of local wavelength
. The particle is repeatedly washed over by these waves, which possess a phase speed
The relative speed between the wave phase and the particle is then
. We may thus deduce the forcing frequency of the wave on the particle by multiplying the relative velocity by the local wavenumber:
This result is in accord with the estimate
evident in
Figure 3 for
Compton periods, that is, before the particle encounters radiation from its periodic image sources. We note that the response to the quasi-monochromatic waves of wavelength
in the particle’s immediate vicinity is the dominant effect on the subsequent particle motion. In
Section 3.2, we have shown that the particle emits radiation at a rate proportional to its instantaneous acceleration. As such, waves generated by the particle’s resulting in-line Zitter over a single oscillation are at most of amplitude
, where
is the amplitude of particle vibration. We plot values of
in
Figure 3b, but note that
in all cases. Finally, we note that in an unbounded domain, the resulting in-line Zitter dies down over time, since the wave amplitude necessarily decays as the wave disperses. This behavior may be seen by comparing
Figure 1a–c.
Upon accelerating, the particle excites a range of wavenumbers
including the set (
17). We proceed to demonstrate that the dominant wavenumbers of the outgoing wavefront are, in fact, bounded above in norm by the value
. Consider the behavior of the particle for
Compton periods in
Figure 2, when the particle has circled the far side of its periodic domain and encounters its own waves
head-on. The relative speed between the particle and
oncoming de Broglie waves is now
, so, following a similar derivation as above, the excited vibration frequency in the particle is
In
Figure 2, this value appears as an approximate upper-bound for the particle oscillations; higher oscillation frequencies, which would correspond to waves of smaller wavelength, are not evident. Although we see in
Figure 1a that some waves of momentum
are excited—i.e., those moving faster than the particle itself—this spectrogram demonstrates that they are negligible compared to their lower-momentum counterparts. Thus, in accelerating from rest, the particle effectively excites only the wavevectors (
17).
We emphasize that when the particle starts from rest, waves are radiated outward
from the point of acceleration. This marks a point of contrast with the behavior in the walking droplet system [
18] and HQFT [
43,
44], where waves are radiated continuously along the particle’s trajectory. In our system, the free particle travels alongside a nearly-planar wavefront of the de Broglie wavelength, as shown in
Figure 1. Finally, our results support the prediction of
Section 3.2, that a nearly constant-velocity state of the free particle is also nearly non-radiating; specifically, an
-amplitude vibration about a constant velocity induces only an
rate of radiation.
We now consider a more general particle acceleration: a particle moving steadily with velocity
is accelerated quickly to a new mean velocity
. As we show rigorously in
Appendix B.2, we can apply a Lorentz transformation to reduce this case to that treated previously, and thus deduce the form of the resulting pilot wave.
Figure 4 illustrates the form of radiation arising in a Lorentz-boosted coordinate system. The wave can no longer simply radiate from the point of acceleration, which is not a well-defined notion under Lorentz symmetry. Instead, a new, continuous wave source is spawned at the point of acceleration, and travels along the
extrapolated, original trajectory of the particle at a velocity
comparable to
, to be derived shortly. This “virtual” source continues to radiate waves in all directions, their form depending on the velocity
difference between the particle and wave source.
To quantify this effect, suppose that
is the velocity of the outgoing particle in the rest frame of the incoming particle. Without loss of generality, we suppose that
lies in the
x direction. In the rest frame of the incoming particle, the outgoing wavefront takes the form described above: a spherically symmetric wavefront expanding (approximately) at the speed
. In particular, the wavefront satisfies the equations
Transforming back to the laboratory frame, the wavefront must satisfy the transformed equations
Specifically, this is the Lorentz transformation of the fastest-moving wavefront in the particle’s rest frame (wherein
), which approximately defines the outer edge of the full wave form.
One must be careful in interpreting the wave geometry in our new frame of reference. In a frame where the particle accelerates from rest, the outer edge of the wave form is a level set simultaneously of wavenumber, phase, and amplitude, but such is not the case in general. First, note that in transforming from the particle’s original rest frame to our new, laboratory frame, the wavenumber becomes higher in front of the particle and lower behind it; thus, the locus (
18) is no longer a level set of the wavenumber. It is also no longer a level set of the wave phase; indeed, the front (
18) corresponds to values of
from a finite time-interval in the particle’s original rest frame (specifically, an interval
, following the equations of a Lorentz transform), over which the high frequency
of the wave would spread the wave front over a range of phases
. However, it
is approximately a level set of the wave amplitude, which is a Lorentz scalar (avoiding the first problem) and modulates far slower than the phase (avoiding the second). We proceed by thinking of it in these terms. Other amplitude level sets, corresponding to longer wavelengths in the particle’s rest frame, can be found by replacing
v with the corresponding group speed (
16).
The relation (
18) reveals the wave form to be ellipsoidal, dilated in the direction
by a factor
. The center of the ellipsoid is traveling with the velocity
which reduces to
if either
or
. The ellipsoid is expanding at the speed
which reduces to
v in the same limits. The excited wavevectors in this wave form are necessarily given by the Lorentz-transformed version of the ball (
17) in reciprocal space.
Figure 4 sketches level sets of the wave amplitude following a general acceleration, as discussed above, but we stress again that these are
not level sets of the phase or of the wavenumber; in fact, the phase generally oscillates rapidly around the outer edge of the wavefront, and up to a rescaling, the wavevector aligns exactly with the group velocity of the immediate wave field, including drift. A particular outcome of this is that even though the particle appears to be traveling askew (not perpendicular) to the plane of the wavefront in
Figure 4, the wavevector at the particle’s position
must (by Lorentz symmetry) match the new momentum of the particle, and thus the phase increases most strongly
in the particle’s direction of motion. By Lorentz symmetry, we also see that this wave form continues to drive particle oscillations at the Compton frequency, although now, these generally have both in-line and lateral components. The wave form thus continues to surround the particle with a quasi-monochromatic region of wavelength
, or wavenumber
, even though the wavevector is
not perpendicular to the amplitude level sets depicted in
Figure 4.
Taken together with the particle oscillation of the previous section, we can understand the radiative behavior in our system as a dynamical version of de Broglie’s harmony of phases, wherein the particle’s internal clock and the underlying guiding wave remain locked in phase throughout its motion. Just as the particle’s vibration updates dynamically to remain at the characteristic frequency , the wavelength of the field (at the particle position) updates to preserve the relation . Together, these effects lock the particle and wave in phase—necessarily, as the wavelength update creates the corresponding frequency update—and preserves the harmony of phases throughout the particle motion.
3.5. Virtual Mass in the Local Wavepacket
We proceed by highlighting a critical feature of the free pilot-wave system, made evident by the local wavepacket (
15): in steady state, the particle
shares a certain amount of energy with the field around it, spread over a radius
. We refer to this as a
virtual mass. While it is not readily apparent in the Equation (
3), it impacts the particle’s inertial response through a constant augmentation
of the particle mass. Quantitatively, this virtual mass is exactly the energy carried by the wavepacket (
15); because the wavepacket is independent of particle dynamics, so too is the virtual mass
.
Now, we note that the virtual mass fraction likely differs between our two-dimensional simulations and the three-dimensional model developed previously. With that in mind, we derive this effect in a three-dimensional system, but provide numerical results only for the two-dimensional AM system.
Recall the momentum continuity Equation (
10):
with
Note that these are simply the momentum and momentum flux densities of a free particle and field, respectively. In a fixed reference frame, we define the corresponding three-momenta as the space integrals
which satisfy the conservation law
obtained by integrating (
10) over space. Now, decompose the wave field as
, where
is simply the wavepacket (
15). This gives us another natural field momentum,
which dominates the total field momentum in the case that radiation rates are small. Such is true in all of our numerical experiments, as we quantify in
Appendix B.1.
Since the particle must bring along a wavepacket of the form
, it is the
combined momentum
that determines the particle’s inertial response. The remainder
determines only higher-order couplings to the underlying wave field.
Since
, we find that
. Critically, we note that the integral for
only involves derivatives of
in the direction of
. Indeed, the coefficient
only depends on this directional derivative by construction, and likewise for the vector
by symmetry. We can then write the integrand as
using the notation
to emphasize that the latter is now a
-independent function of space. Finally, we switch coordinates of integration to those of the rest frame of the particle,
, and thus pick up a factor of
:
Here,
depends only on the mass density
m of the field. Recall from symmetry that
must point parallel to
, restricting
for some
independent of velocity. In fact, we know that
must be proportional to
m, as the only remaining mass scale in the system. In total, we find that
This
effective mass, rather than the “decoupled mass”
m, thus determines the inertial response of the particle.
We can confirm the influence of the wave-induced virtual mass by looking again at the numerical experiments of
Section 3.3. Recall that, in those experiments, we start a particle at rest, impart a momentum
to the particle, and observe its evolution. Though we focused on velocity oscillations before, we now look at how the particle approaches a steady-state speed
.
Figure 5a shows horizontal particle position after imparting a velocity
. First, note that the relaxation to the steady-state speed
occurs over the Compton timescale (roughly the period of several oscillations, as seen in the cutout). This short-time dynamics is characterized by a transfer of momentum from the particle to its adjoining wavepacket, as predicted by Corollary 1. We can think of this exchange as a reflection of the particle’s delocalised nature: since the local wavepacket requires momentum everywhere over a radius
, it requires a time
to distribute momentum appropriately. This also explains why radiation closely follows a single-source approximation, as we encountered in the preceding section; radiation occurs upon particle-to-wave energy transfer, which occurs over a Compton timescale.
Figure 6 shows the exchange of (horizontal) momentum between particle and wave for the initial time period in
Figure 5a (for
). Here, we see that the magnitude of oscillations dies down significantly after
Compton periods—i.e., as the particle approaches its steady-state velocity. After this point, the particle velocity continues oscillating at a lower amplitude, corresponding to the in-line
Zitterbewegung of
Section 3.3.
Figure 5b shows the quantity
over a range of initial velocities
. This measures the fraction of momentum retained by the particle, which we use as a rough measure of the fraction of momentum
transferred to the wavepacket. The limitation of this metric is that momentum can
either be transferred to the wavepacket
or radiated away. Here, we see both effects. First, since the momentum fraction carried by the steady-state wavepacket is velocity-independent—as quantified in (
21)—each curve in this figure is bounded above by
. To continue, we assume that at the maximum point in each curve in
Figure 5b, the particle does not radiate
any momentum away. That is, for each
b,
at the maximum point on each curve. With this approximation, the best fit to
is given by
with a maximum relative error of
. This represents a very close fit to our prediction
.
Finally, looking beyond the maximum point of each curve, note that a significant amount of momentum is lost beyond the
transferred to the wavepacket. As
b increases, the wavepacket itself grows, and more momentum is taken up by the virtual mass. Conversely, as
decreases, more is radiated away on top of the wavepacket. We can understand this through Corollary 1; the momentum transfer between particle and field is
so as
with increasing particle velocity or decreasing coupling constant, less momentum is available to radiate.
As a point of note, we see that the curve
roughly cuts the system into two states: when accelerated (from rest) above a critical momentum
, the particle settles quickly into a steady state with virtual momentum fraction
This “low-radiation regime” characterizes the experiments above the black curve in
Figure 5b. When accelerated below this critical momentum, however, the particle initially loses a substantial fraction of its momentum to radiation.
3.6. Heisenberg Uncertainty, and the Particle Cloud
Recall from
Section 3.4 that, in the case of a free particle, an in-line Zitter is excited by the phase waves washing over the particle. Now suppose the particle is not free, but confined to a finite geometry—we assume only that its waves wash over it from all directions, either reflected off of walls or generated by its periodic images. The resulting wave interference characterizes our periodic system in the long-time limit, as in
Figure 1c, but it is also expected to arise when the particle interacts with a variety of common quantum apparatuses (e.g., slits, corrals, or interferometers) or indeed during any position measurement. In such confined geometries, Zitter occurs in
all directions, owing to the complex geometry of the incoming waves. The resulting motion is characterized by a region of scale
around the mean trajectory, about which the particle vibrates at a characteristic frequency
. We call this region the
particle cloud, whose form is shown in
Figure 7b.
We investigate this vibration in 1D by returning to the numerical experiments of
Figure 2, and focusing on their long-time limit, wherein previously-radiated waves are incoming from all directions. Amplitudes of these oscillations are given in
Figure 3 over a range of velocities and coupling constants. Namely, we calculate each amplitude as the
norm (over frequency space) of the short-time Fourier transform discussed in
Figure 2, averaged over the last 1000 time-steps. Note that higher
norms are bounded by this value up to a constant multiple, as the frequency spectra have (approximately) uniformly compact support.
In
Figure 3b, we show best-fit curves of the form
for each value of
b. Here,
and
both decrease with
b for all values within the tested range. Critically, note that
for sufficiently large
b. This gives an oscillation amplitude
in any direction, where
and
are as in
Figure 3. Using the characteristic oscillation frequency
, based on our discussion in
Section 3.4, we find
To estimate the corresponding momentum amplitude
, suppose that the particle oscillates from a minimum momentum
to a maximum
in the direction of
. Here,
is the
effective mass of both the particle and its steady-state wavepacket, as discussed in
Section 3.5. Furthermore, note that
and similarly
, with equality only if the particle is confined to move in one direction. In either case, the equality
holds, where
is the
total Lorentz factor of the particle. Then, supposing transverse oscillations are small, we have
from Jensen’s inequality, using
to denote the particle’s mean velocity. In turn, our estimate (
23) gives
Modeling
x and
p as monochromatic oscillations, we have
and
, or
Finally, applying
gives
or in dimensional form,
Given (
22), this inequality reduces to the traditional uncertainty principle for sufficiently large
b. Note that a linear regression suggests that we should find
at
; in this stronger coupling regime, these oscillations would obey the stronger
relativistic uncertainty principle of Putra and Alrizal [
51]:
.
We emphasize that the uncertainty relation (
24) is somewhat different in nature to its counterpart in quantum mechanics. Instead of an underlying property of a wave-like system, as in quantum theory, our uncertainty relation characterizes a
classical uncertainty of the Compton-scale particle dynamics, brought about by its waves interacting with a wall-bounded geometry. Averaging over the Compton timescale of the particle, the particle appears to take up a Compton-scale volume in phase space, reminiscent of quantum mechanics. However, in quantum mechanics, these scales can be squeezed in either position or momentum coordinates, giving rise to (a) highly localized states of indefinite momentum and (b) delocalised states of definite momentum. This type of squeezing does not appear to have an analogue in our uncertainty relation.