You are on page 1of 28

Observation of optical de Broglie-Mackinnon wave packets

Layton A. Hall1 and Ayman F. Abouraddy1,∗


1CREOL, The College of Optics & Photonics,
University of Central Florida, Orlando, Florida 32816, USA

Abstract
arXiv:2302.13194v1 [physics.optics] 25 Feb 2023

de Broglie wave packets accompanying moving particles are dispersive and lack an intrinsic length scale
dictated solely by the particle mass and velocity. Mackinnon proposed almost 45 years ago a localized
non-dispersive wave packet constructed out of dispersive de Broglie phase waves via a Copernican
inversion of the roles of particle and observer, whereupon an intrinsic length scale emerges by accounting
for every possible observer – rather than by introducing an ad hoc uncertainty in the particle velocity. The
de Broglie-Mackinnon (dBM) wave packet has nevertheless remained to date a theoretical entity. Here, we
report the observation of optical dBM wave packets using paraxial space-time-coupled pulsed laser fields
in presence of anomalous group-velocity dispersion. Crucially, the bandwidth of dBM wave packets has
an upper limit that is compatible with the wave-packet group velocity and equivalent mass. In contrast to
previously observed linear propagation-invariant wave packets whose spatio-temporal profiles at any axial
plane are X-shaped, those for dBM wave packets are uniquely O-shaped (circularly symmetric with respect
to space and time). By sculpting their spatio-temporal spectral structure, we produce dispersion-free
dBM wave packets in the dispersive medium, observe their circularly symmetric spatio-temporal profiles,
and tune the field parameters corresponding to particle mass and velocity that uniquely determine the
wave-packet length scale.

∗ Corresponding authors: raddy@creol.ucf.edu

1
It is well-known that there are no dispersion-free wave-packet solutions to the (1 + 1)D
potential-free Schrödinger equation – with the sole exception of the Airy wave packet iden-
tified by Berry and Balasz in 1979 [1], which does not travel at a fixed group velocity, but rather
accelerates despite the absence of an external force [2]. The Airy wave packet has impacted all
areas of wave physics (e.g., optics [3], acoustics [4], water waves [5], electron beams [6], and as
a model for Dirac particles [7]). Less-known is that in the year preceding the discovery of the
Airy wave packet, Mackinnon identified a non-dispersive (1 + 1)D wave packet that travels at
a constant group velocity [8], but is constructed out of dispersive de Broglie ‘phase waves’ that
accompany the motion of a massive particle and are solutions to the Klein-Gordon equation.
de Broglie had originally demonstrated that the group velocity ve of a wave packet constructed of
phase-waves is equal to the particle velocity v [9]. However, localized de Broglie wave packets
are dispersive, as are Schrödinger wave packets [10]. Moreover, because de Broglie wave packets
necessitate introducing an ad hoc uncertainty in the particle velocity [9], and there is no upper
limit on the exploitable bandwidth, such wave packets lack an intrinsic length scale (i.e., a scale
uniquely determined by the particle mass and velocity).
Through a Copernican inversion of the roles of particle and observer, Mackinnon constructed
out of dispersive de Broglie phase waves a non-dispersive wave packet [8] – which we refer to
henceforth as the de Broglie-Mackinnon (dBM) wave packet. Instead of introducing an ad hoc
uncertainty into the particle velocity from the perspective of a privileged reference frame, Mackin-
non suggested accounting for all possible observers, who cooperatively report observations made
in their reference frames to a single agreed-upon frame in which Lorentz contraction and time
dilation are corrected for [8]. Besides retaining the salutary features of conventional de Broglie
wave packets, Mackinnon’s construction unveiled an intrinsic length scale for the dBM wave
packet determined solely by the particle mass and velocity. However, despite the clear algorithmic
process for constructing the dBM wave packet, it is not a solution to the Klein-Gordon equation
[8], and is instead constructed only epistemologically in the selected reference frame. As such,
dBM wave packets have yet to be realized in any physical wave.
Nevertheless, it has been recognized that the (1 + 1)D dBM wave packet can be mapped to
physical solutions of the optical wave equation by first enlarging the field dimensionality to
(2 + 1)D, which allows introducing angular dispersion [11, 12]. This procedure enables realizing
the dBM dispersion relationship for propagation along the optical axis in the initial reduced-
dimensionality (1 + 1)D space [13]. However, observing optical dBM wave packets in free

2
space faces insurmountable practical difficulties [14, 15]. Specifically, such wave packets are
produced by relativistic optical dipoles and are observed by stationary, coherent field detectors
that nevertheless fully encircle the moving dipole. We investigate here a different strategy that
makes use of the unique characteristics of optical-wave propagation in the anomalous group-
velocity dispersion (GVD) regime [16] to produce paraxial dBM wave packets. In this conception,
an optical dBM wave packet is a particular realization of so-called ‘space-time’ (ST) wave packets
[15, 17–20] in dispersive media [21–27]. In general, ST wave packets are pulsed optical beams
whose unique characteristics (e.g., tunable group velocity [28] and anomalous refraction [29])
stem from their spatio-temporal spectral structure rather than their particular spatial or temporal
profiles. Recent advances in the synthesis of ST wave packets make them a convenient platform
for producing a wide variety of structured pulsed fields [15], including dBM wave packets.
Here, we provide unambiguous observations of optical dBM wave packets in presence of
anomalous GVD. Starting with generic femtosecond pulses, we make use of a universal angular-
dispersion synthesizer [30] to construct spatio-temporally structured optical fields in which the
spatial and temporal degrees-of-freedom are no longer separable. Critically, the association
between the propagation angle and wavelength is two-to-one rather than one-to-one as in con-
ventional tilted pulse fronts [11]. This feature allows folding the spatio-temporal spectrum back
on itself, thereby guaranteeing the paraxiality of the synthesized dBM wave packets. Conse-
quently, these wave packets retain in the medium all the characteristic features of their free-space
counterparts while circumventing the above-mentioned difficulties. Such space-time-coupled
wave packets are dispersive in free space, but become propagation-invariant once coupled to a
medium in the anomalous-GVD regime, where they travel at a tunable group velocity ve. Although
all previously observed linear, propagation-invariant wave packets have at a fixed axial plane
been either X-shaped [15, 31–33] or separable [34] with respect to the transverse coordinate and
time, the spatio-temporal profiles of dBM wave packets are – in contrast – circularly symmetric
(O-shaped). In addition to verifying this long-theorized O-shaped spatio-temporal structure
[23–25], we confirm the impact of the two identifying parameters (equivalent to particle mass
and velocity) on the bandwidth and length scale of the non-dispersive dBM wave packets. Prop-
agation invariance in the dispersive medium constrains the maximum bandwidth (minimum
wave-packet length) according to these selected parameters. Finally, in contrast to Airy wave
packets that are the unique non-dispersive solution to the Schrödinger equation, the axial profile
of dBM wave packets can be varied almost arbitrarily, which we confirm by modulating their

3
spatio-temporal spectral phase distribution. These results may pave the way to optical tests of
the solutions of the Klein-Gordon equation for massive particles.

RESULTS

Theory of de Broglie wave packets

de Broglie posited two distinct entities accompanying massive particles: an internal ‘clock’ and
an external ‘phase wave’ [35, 36]. For a particle of rest mass mo whose energy is expressed as
Eo=moc2=h̄ωo, the internal clock and the infinite-wavelength phase wave coincide at the same
de Broglie frequency ωo in the particle’s rest frame [Fig. 1(a)]; here c is the speed of light in
vacuum, and h̄ is the modified Planck constant. When the particle moves at a velocity v, the
p
frequencies observed in the rest frame diverge: the internal frequency drops to ω =ωo 1 − β2v
p
whereas the phase-wave frequency increases to ω =ωo 1 − β2v and takes on a finite wavelength
λ, where β v= vc [Fig. 1(b)]. The wave number k = 2π
λ for the phase wave is determined by the
de Broglie dispersion relationship ω 2=ωo2 + c2k2 [Fig. 1(c)], so that it is a solution to the Klein-
Gordon equation. Because de Broglie phase waves are extended, a particle with a well-defined
velocity cannot be localized. Instead, spatially localizing the particle requires introducing an ad
hoc uncertainty in the particle velocity (a spread from v to v + ∆v) to induce a bandwidth ∆ω
(from ωc to ωc + ∆ω), or ∆k (from kc to kc + ∆k) [8, 9], thus resulting in a finite-width wave packet
that is also a solution to the Klein-Gordon equation [Fig. 1(c)]. The wave-packet group velocity
2
ve=1 dk =v is equal to the particle velocity, whereas its phase velocity is vph= ω = c (vphve=c2;

dω ωc k v
de
v
see Methods). However, de Broglie wave packets are dispersive dω 6 =0. Moreover, because there is
no upper limit on the exploitable bandwidth [Fig. 1(c)], de Broglie wave packets lack an intrinsic
length scale; that is, there is no minimum wave-packet length that is uniquely identified by the
particle parameters (mass mo and velocity v).

Non-dispersive de Broglie-Mackinnon (dBM) wave packets

Mackinnon proposed an altogether different conception for constructing localized non-dispersive


wave packets out of de Broglie phase waves that jettisons the need for introducing an ad hoc
uncertainty in particle velocity to localize it. Key to this proposal is a Copernican inversion of

4
the roles of particle and observer. Rather than a single privileged observer associated with the
rest frame in Fig. 1(c), Mackinnon considered a continuum of potential observers traveling at
physically accessible velocities (from −c to c). The wave-packet bandwidth ∆k that is established
in a particular reference frame is a result of the spread in the particle velocity as observed in all
these possible frames. Consequently, the particle can be localized, and a unique wave-packet
length scale identified, even when its velocity is well-defined.

The physical setting envisioned by Mackinnon is depicted in Fig. 1(d), where the particle moves
at a velocity v and an observer moves at u, both with respect to a common rest frame in which
the dBM wave packet is constructed. Each potential observer records a different phase-wave
frequency and wavelength. The crucial step is that all potential observers travelling at velocities u
ranging from −c to c report their observations to the selected rest frame. These phase waves are
superposed in this frame – after accounting for Lorentz contraction and time dilation (Methods) –
to yield a wave packet uniquely identified by the particle rest mass mo and velocity v.

Consider first the simple scenario where the particle is at rest with respect to the selected frame
(v=0). Each observer reports to the common rest frame a frequency ω 0=ωo/ 1 − β2u and a wave
p

number k0=−ko β u/ 1 − β2u, where β u= uc . Accounting for time dilation results in ω 0 → ω =ωo,
p

and accounting for Lorentz contraction produces k0 → k =−ko β u. Therefore, the frequency in the
rest frame based on the recordings of all the observers is ω =ωo, just as in the case of a conventional
de Broglie phase wave, but the wave number now extends over the range from −ko to ko as
the observer velocity u ranges from c to −c [Fig. 1(e)]. In other words, the observer velocity u
serves as an internal parameter that is swept to establish a new dispersion relationship whose
slope is zero, thus indicating a particle at rest ve=v=0 [8, 13]. The spectral representation of the
support domain for this wave packet is a horizontal line ω =ωo in (k, ωc )-space delimited by the
two light-lines k=± ωc [Fig. 1(e)]. In contradistinction to conventional de Broglie wave packets,
a physically motivated length scale emerges for the dBM wave packet. The maximum spatial
bandwidth is ∆k=2ko, which corresponds to a minimum wave-packet length scale of Lmin∼ λ2o ,
where λo= 2π
ko . This can be viewed as an ‘optical theorem’, whereby the dBM wave packet for
a stationary particle cannot be spatially localized below the associated de Broglie wavelength
λo. Taking an equal-weight superposition across all the wave numbers, the dBM wave packet is
ψ(z;t) ∝e−iωotsinc( ∆k sinπx
π z ), where sinc( x )= πx [8].

5
A similar procedure can be followed when v6=0, whereupon the frequency and wave num-
p p
ber in the selected reference frame are ω =ωo(1 − β v β u) 1 − β2v and k =ko( β v − β u) 1 − β2v,
respectively (Methods). Because v is fixed whereas u extends from −c to c, a linear dispersion
k2
relationship between ω and k is established, k = β1v ( ωc − ko ), where k1=ko/ 1 − β2v. The slope of
p
1

the dBM dispersion relationship indicates that ve=v as in conventional de Broglie wave packets,
de
v
but the dBM wave packet is now non-dispersive, dω =0 [Fig. 1(f)]. The limits on the spatial and
temporal bandwidths for the dBM wave packet are ∆k =2k1 and ∆ω c = β v∆k, respectively, leading
p
to a reduced characteristic length scale Lmin∼ λ2o 1 − β2v as a manifestation of Lorentz contraction;
a faster particle is more tightly localized. By assigning equal complex amplitudes to all the
phase waves associated with this moving particle, the propagation-invariant dBM wave packet is
ψ(z;t) ∝eiβ v∆k(z−vet)sinc( ∆k
π (z − v
et)). Crucially, unlike conventional de Broglie wave packets, the
dBM wave packet is not a solution to the Klein-Gordon equation, although a modified wave
equation can perhaps be constructed for it [8].

Optical de Broglie-Mackinnon wave packets in free space

Despite their intrinsic interest from a fundamental point of view, dBM wave packets have
remained to date theoretical entities. It has nevertheless been recognized that optical waves in
free space may provide a platform for their construction [13, 14]. Because (1 + 1)D optical waves
in free space are dispersion-free (k = ωc and vph=ve=c), producing optical dBM wave packets
requires first adding a transverse coordinate x to enlarge the field dimensionality to (2 + 1)D. The
dispersion relationship thus becomes k2x + k2z =( ωc )2, which represents the surface of a ‘light-cone’
[15, 37]; here k x and k z are the transverse and longitudinal components of the wave vector along
x and z, respectively. The spectral support of any optical field corresponds to some region on
the light-cone surface [Fig. 2(a)]. For a fixed value of k x=± ωco , we retrieve the axial dispersion
relationship for de Broglie phase waves ω 2=ωo2 + c2k2z. A convenient parametrization of the field
makes use of the propagation angle ϕ(ω ) with respect to the z-axis for the plane wave at a
frequency ω, whereupon k x(ω )= ωc sin ϕ(ω ) and k z(ω )= ωc cos ϕ(ω ). Angular dispersion is thus
introduced into the (2 + 1)D field [11, 12], and its spectral support on the light-cone surface is
a one-dimensional (1D) trajectory. We take optical dBM wave packets to be those whose axial
dispersion relationship ω (k z) conforms to that of a dBM wave packet. This requires that the
projection of the spectral support onto the (k z, ωc )-plane be linear and delimited by the light-lines

6
k z=± ωc . Indeed, the spectral projections onto the (k z, ωc )-plane in Fig. 2(a,b) coincide with those
in Fig. 1(e,f).
Consider first a monochromatic field ω =ωo whose spectral support is the circle at the inter-
section of the light-cone with a horizontal iso-frequency plane [Fig. 2(a)]. This monochromatic
field comprises plane waves of the same frequency ωo that travel at angles ϕ extending from
p
0 to 2π, whose axial wave numbers are k z( ϕ)=± k2o − k2x=ko cos ϕ and extend from −ko to ko.
This optical wave packet [Fig. 2(a)] corresponds to the dBM wave packet for a particle in its
rest frame [Fig. 1(e)], and ϕ serves as the new internal parameter to be swept in order to pro-
duce the targeted dBM dispersion relationship, corresponding to the observer velocity u in
Fig. 1(e). By setting the spectral amplitudes equal for all the plane-wave components, we obtain

ψ( x,z;t) ∝e−iωot sinc( ∆k
π
z
x2 + z2), where ∆k z=2ko [Fig. 2(a)]. Such a wave packet can be produced
by a stationary, monochromatic planar dipole placed at the origin of the ( x,z)-plane. Observing
this optical field requires coherent field detectors arranged around the 2π angle subtended by the
dipole, and then communicating the recorded measurements to a central station. This procedure
is therefore not dissimilar in principle from that envisioned by Mackinnon for the dBM wave
packet associated with a stationary particle, in which the measurements recorded by observers
traveling at different velocities are communicated to the common rest frame [Fig. 1(d)].
When the dipole moves at a velocity v along the z-axis with respect to stationary detectors
encircling it, each constituent plane-wave undergoes a different Doppler shift in the rest frame of
the detectors. The field still comprises plane waves travelling at angles ϕ extending from 0 to 2π,
but each plane wave now has a different frequency ω. Nevertheless, the new spectral support for
the dBM wave packet on the light-cone is related to that for the stationary monochromatic dipole.
Indeed, the Lorentz transformation associated with the relative motion between the source and
detectors tilts the horizontal iso-frequency spectral plane in Fig. 2(a) by an angle θ with respect to
the k z-axis as shown in Fig. 2(b), where tanθ = β v [13, 38–40], thus yielding a tilted ellipse whose
projection onto the (k x, ωc ) is:
k2x (ω − ck1)2
+ = 1. (1)
k2o (∆ω/2)2
k2
The spectral projection onto the (k z, ωc )-plane is now the line k z=k + + ω −veω+ = β1v ( ωc − ko ), where
q 1
ω+ 1+ β v p
ve=ctanθ =v is the wave-packet group velocity along z, k += c =ko 1− β v , and k1=ko/ 1 − β2v.
∆ω
The spatial and temporal bandwidths are related via c = β v∆k z, where ∆k z=2k1. Each plane
wave travels at a different direction in the ( x,z)-plane in such a way that their axial wave numbers

7
k z reproduce the dBM dispersion relationship [compare Fig. 1(f) to Fig. 2(b)]. By setting the
complex spectral amplitudes constant for all frequencies, we obtain the dBM wave packet (with
ve< c):
∆k z
q  
iβ v∆k z(z−vet)
ψ( x,z;t) ∝e sinc x2 +(z − vet)2 , (2)
π
Two parameters uniquely identify the optical dBM wave packet: the group velocity ve (corre-
sponding to the particle velocity) and the wave number ko (corresponding to the particle mass).
Furthermore, the signature of the dBM wave packet in Eq. 2 is its circularly symmetric spatio-
temporal profile in ( x,t)-space in any axial plane z. In contrast, all other propagation-invariant
wave packets that have been observed in free space are X-shaped [15, 22, 31–33] and are not
circularly symmetric. Indeed, truncating the spectrum of the optical dBM wave packet obstructs
the formation of the circularly symmetric profile and gives rise instead to the more familiar
X-shaped counterpart [14, 15]. The O-shaped spatio-temporal profile as indicated by Eq. 2 can be
observed only when the full bandwidth – delimited by the light-lines – is included.
The field in the ( x,z)-plane recorded by stationary detectors encircling the dipole takes the
form shown in Fig. 2(b), as pointed out recently in a thought experiment by Wilczek [41]. Despite
the conceptual simplicity of this optical scheme for producing dBM wave packets, it nevertheless
faces obvious experimental challenges. Encircling an optical dipole moving at a relativistic speed
with stationary detectors is far from practical realizability. The more realistic configuration in
which the detectors are restricted to a small angular range within the paraxial regime centered on
the z-axis truncates the recorded field and precludes observing of the O-shaped spatio-temporal
profile [14, 15]. For these reasons, it is not expected that the O-shaped dBM wave packet can be
observed using spatio-temporally structured optical fields in free space.

Optical de Broglie-Mackinnon wave packets in a dispersive medium

The necessity of including the entire bandwidth delimited by the intersection of the dBM disper-
sion relationship with the free-space light-cone [Fig. 2(a-b)] presents insurmountable experimental
obstacles. Producing paraxial dBM wave packets necessitates confining the spectrum to a narrow
range of values of k z centered at a value k z∼ko>0. Crucially, the linear spatio-temporal spectrum
projected onto the (k z, ωc )-plane must remain delimited at both ends by the light-line, to produce
the circularly symmetric spatio-temporal wave-packet profile. Clearly these requirements cannot
be met in free space. Nevertheless, this challenge can be tackled by exploiting the unique fea-

8
tures of optical-wave propagation in the presence of anomalous GVD. Specifically, the light-cone
structure is modified in presence of anomalous GVD so that the curvature of the light-line has
the same sign as that of the de Broglie dispersion relationship [Fig. 2(c)]. In this case, imposing
the characteristically linear dBM dispersion relationship produces a spectral support domain
on the dispersive light-cone surface that satisfies all the above-listed requirements: (1) k z>0 is
maintained throughout the entire span of propagation angles ϕ(ω ); (2) the field simultaneously
remains within the paraxial regime; and (3) the spectrum is delimited at both ends by the light-line
[Fig. 2(c)], thus yielding a wave packet having a circularly symmetric spatio-temporal profile.
The spectral support is in the form of an ellipse at the intersection of the dispersive light-cone
with a tilted spectral plane. The center of this ellipse is displaced to a large value of k z, and
the spectral projection onto the (k z, ωc )-plane is a line making an angle θ with the k z-axis. The
resulting wave packet is propagation-invariant in the dispersive medium and travels at a velocity
ve=ctanθ independently of the physical parameters of the dispersive medium.
In the anomalous-GVD regime, the wave number is given by k(ω )=n(ω )ω/c = k (ωo + Ω)≈
nmko + veΩm − 12 |k2m|Ω2 +··· ; where n(ω ) is the refractive index, and the following quantities are
 dk
all evaluated at ω =ωo: nm=n(ωo) is the refractive index, vem=1 dω is the group velocity for
ωo
d2k

a plane-wave pulse in the medium, and k2m= dω2 ω =−|k2m| is the negative-valued anomalous
o

GVD coefficient [16]. The dispersion relationship in the medium k2x + k2z =k2 corresponds geomet-
rically to the surface of the modified dispersive light-cone in Fig. 2(c). Similarly to the free-space
scenario, we impose a spectral constraint of the form k z=nmko + Ωve = β1v ωc − ko(1 − nm β v) in the


medium, where Ω=ω − ωo and ve=ctanθ is the group velocity of the wave packet [Fig. 2(c)]. The
wave-packet spectrum as defined by this constraint is delimited by the light-line at its two ends,
both located however in the range k z>0, in contrast to the previous scenarios depicted in Fig. 1(e,f)
and Fig. 2(a,b); see Methods.
The spectral projections onto the (k x, ωc ) and (k x,k z) planes of the spectral support on the
dispersive light-cone are ellipses (Methods):
k2x ( ω − ωc ) 2 k2x ( k z − k c)2
+ = 1, + = 1, (3)
k2x,max (∆ω/2)2 k2x,max (∆k z/2)2
∆ω ko 1− β0v
where the temporal bandwidth is c = 2 σm β2v = β v∆k z, σm= cωo| k 2m| is a dimensionless dis-

persion coefficient, β0v= vevem , k x,max= 12 ∆ω
c nmσm, ωc=ωo − ∆ω/2, kc=nmko − ∆k2 z , k x,maxnmko,
∆k znmko, and ∆ω ωo. It is crucial to recognize that the ellipse projected onto the (k x,k z)-plane
does not enclose the origin (k x,k z)=(0,0), but is rather displaced to a central value kc∆k z. There-

9
fore, the optical field comprises plane-wave components that propagate only in the forward
direction within a small angular range centered on the z-axis, and the field thus remains within
the paraxial domain. Nevertheless, because the spectrum is delineated at both ends by the curved
dispersive light-line, the resulting spatio-temporal profile is circularly symmetric in any axial
plane z. This wave packet in the dispersive medium thus satisfies all the above-listed desiderata
for an optical dBM wave packet, but can be readily synthesized and observed in contrast to its
free-space counterparts. One difficulty, however, arises from the form of ϕ(ω ) in the dispersive
medium, which differs fundamentally from that in free space. Each frequency ω in a free-space
optical dBM wave packet is associated with two propagation angles ± ϕ(ω ). However, each prop-
agation angle ϕ is associated with a single frequency, so that |φ(ω )| is one-to-one. In the optical
dBM wave packet in the dispersive medium, each ω is still associated with two propagation
angles ± ϕ(ω ); but ϕ(ω ) is now two-to-one, so that ϕ(ω ) is folded back on itself [Fig. 2(c)]. To
synthesize such a field configuration, a synthesizer capable of sculpting ϕ(ω ) almost arbitrarily
is required.

Experimental confirmation

Setup. To construct the optical dBM wave packet in free space from a generic pulsed beam
in which the spatial and temporal degrees-of-freedom are uncoupled, we introduce angular
dispersion by assigning to each wavelength λ a particular pair of angles ± ϕ(λ), thereby coupling
the spatial and temporal degrees-of-freedom. We carry out this task using a universal angular-
dispersion synthesizer [30], in which a spatial light modulator (SLM) deflects each wavelength
from a spectrally resolved laser pulse at prescribed angles, as illustrated in Fig. 3 (Methods).
Because each wavelength λ is deflected at ϕ(λ) independently of all other wavelengths, ϕ(λ) need
not be one-to-one. Indeed, it can readily be a two-to-one mapping as required for paraxial optical
dBM wave packets. The dBM wave packet is formed once all the wavelengths are recombined by
a grating to reconstitute the pulsed field. The spatio-temporal spectrum of the synthesized wave
packet is acquired by operating on the spectrally resolved field with a spatial Fourier transform
and recording the intensity with a CCD camera. This measurement yields the spatio-temporal
spectrum projected onto the (k x,λ)-plane, from which we can obtain the spectral projection onto
the (k z,λ)-plane. The spatio-temporal envelope I ( x;τ ) of the intensity profile at a fixed axial plane
z is reconstructed in the frame travelling at ve (τ =t − z/e
v) via linear interferometry exploiting

10
the procedure developed in Refs. [28, 29, 42] (Methods). The dispersive medium exploited in
our measurements is formed of a pair of chirped Bragg mirrors providing an anomalous GVD
coefficient of k2m≈−500 fs2/mm and vem≈c (Methods).
Measurement results. We first verify the unique signature of dBM wave packets in presence
of anomalous GVD; namely, the O-shaped spatio-temporal intensity profile at any axial plane
after inculcating into the field the dBM dispersion relationship. In Fig. 4 we verify three sought-
after features: (1) The closed elliptical spatio-temporal spectrum projected onto the (k x,λ)-plane;
(2) the linear spectral projection onto the (k z,λ)-plane, indicating non-dispersive propagation
in the dispersive medium; and (3) the circularly symmetric spatio-temporal intensity profile
I ( x;τ ) reconstructed at a fixed axial plane (z=30 mm). In Fig. 4(a) we plot the measurements
for an optical dBM wave packet having a group velocity ve=0.9975c. The temporal bandwidth
is constrained to a maximum value of ∆λ≈16 nm, and the associated spatial bandwidth ∆k x≈
0.03 rad/µm, thus resulting in a pulsewidth ∆T ≈200 fs at x =0, and a spatial profile width
∆x ≈38 µm at τ =0. The spectral projection onto the (k z,λ)-plane is delimited at both ends by the
curved light-line of the dispersive medium. In other words, a larger bandwidth is incompatible at
this group velocity with propagation invariance in the dispersive medium. Further increase in
the bandwidth extends the spectral projection below the dispersive light-line, which contributes to
only evanescent field components. The measured spatio-temporal profile I ( x;τ ) therefore has the
smallest dimensions in space and time for a circularly symmetric dBM wave packet compatible
with the selected group velocity in the medium.
To the best of our knowledge, this is the first observation of an O-shaped spatio-temporal
intensity profile for a dispersion-free wave packet in a linear dispersive medium. Previous
realizations of dispersion-free ST wave packets in dispersive media (whether in the normal-
or anomalous-GVD regimes) revealed X-shaped spatio-temporal profiles [27] similar to those
observed in free space [28, 31, 43] or in non-dispersive dielectrics [29]. In these experiments,
however, the wave packets were not delimited spectrally by the dispersive-medium light-line,
which is the prerequisite for the realization of O-shaped optical dBM wave packets.
As mentioned earlier, two parameters characterize a dBM wave packet: the velocity v and the
rest mass mo. The corresponding variables associated with the optical dBM wave packet are ve and
λo, which can both be readily tuned in our experimental arrangement by changing the functional
dependence of ϕ on λ. In this way we can vary the first parameter; namely, the group velocity
ve. Increasing the group velocity from ve=0.9975c [Fig. 4(a)] to ve=0.9985c [Fig. 4(b)] and then to

11
ve=0.999c [Fig. 4(c)] reduces the maximum exploitable temporal bandwidth from ∆λ≈16 nm to
∆λ≈8 nm and ∆λ≈6 nm, respectively, while retaining the closed elliptic spectral projection onto
the (k x,λ)-plane, the linear spectral projection onto the (k z,λ)-plane, and the associated O-shaped
spatio-temporal profile I ( x;τ ). The corresponding spatial bandwidths drop to ∆k x≈0.023 rad/µm
and ∆k x≈0.017 rad/µm, respectively. In all three dBM wave packets in Fig. 4, we retain a fixed
intersection with the dispersive light-line at λo≈1054 nm (corresponding to a fixed particle mass),
such that reducing ve decreases the wavelength of the second intersection point. The second
parameter, the wavelength λo corresponding to particle rest mass mo for de Broglie phase waves,
can also be readily tuned [Fig. 5]. Here, the maximum exploitable bandwidth changes as a result of
shifting the value of λo from λo=1054 nm [Fig. 5(a)] where ∆λ=16 nm, to λo=1055 nm [Fig. 5(b)]
where ∆λ=14 nm, and then to λo=1056 nm [Fig. 5(c)] where ∆λ=12 nm. Once again, both the
spatial and temporal widths of the circularly symmetric O-shaped profile in the ( x,t)-domain
change accordingly.

The Airy wave packet, as mentioned earlier, is the unique non-dispersive solution to Schrödinger’s
equation – no other waveform will do [44]. Although Mackinnon obtained a particular ‘sinc’-
function-shaped wave packet [8], this waveform is not unique. Indeed, the sinc-function results
from combining all the de Broglie phase waves with equal weights. However, dBM wave packets
can take on in principle arbitrary waveforms by associating different magnitudes or phases with
the plane-wave components constituting it. We confirm in Fig. 6 that the spatio-temporal profile
I ( x;τ ) of optical dBM wave packets can be modified while remaining propagation invariant in
the dispersive medium. First, setting the complex spectral amplitudes equal along the elliptical
spectral support, we obtain propagation-invariant circularly symmetric wave packets in the
dispersive medium [Fig. 6(a)]. Truncating the ellipse and eliminating the plane wave components
in the vicinity of k x=0 disrupts the formation of the full circular profile, but the wave packet
nevertheless propagates invariantly [Fig. 6(b)]. By introducing a π-step in the spectral phase
along k x, a spatial null is formed along x =0 in the profile of the dBM wave packet [Fig. 6(c)],
whereas introducing the π-phase-step along λ produces a temporal null along τ =0 [Fig. 6(d)].
Finally, alternating the phases between 0 and π in the four quadrants of the spatio-temporal
spectral plane (k x,λ) produces spatial and temporal nulls along x =0 and τ =0, respectively
[Fig. 6(e)]. Despite such variations in their spatio-temporal profiles, all these optical dBM wave
packets propagate invariantly in the dispersive medium.

12
DISCUSSION

The rapidly evolving versatile techniques for synthesizing optical fields [15, 45] played a
critical role in the realization of dBM wave packets as demonstrated here. This has helped
confirm the theoretical proposal made by Mackinnon almost 45 years ago for constructing a
non-dispersive wave packet from dispersive de Broglie phase waves [8]. Furthermore, the
experimental procedure implemented here points to a general synthesis strategy that extends
beyond the particular scenario of dBM wave packets. The overarching theme is that novel
dispersion relationships for the axial propagation of a wave packet can be imposed by first adding
another dimension to the space, and then exploiting the new dimension to tailor the dispersion
relationship before spectral projection back onto the original reduced-dimensionality space.
In the scenario studied here, we start with a (1 + 1)D physical wave in which an axial dis-
persion relationship ω (k z) is enforced by the dynamics of the wave equation. Increasing the
dimensionality of the space from (1 + 1)D to (2 + 1)D by including a transverse coordinate x yields
a new dispersion relationship ω (k x,k z). In free space, optical wave packets are subject to the
p
constraint ω =ck z in (1 + 1)D and ω (k x,k z)=c k2x + k2z in (2 + 1)D. Now, by judiciously associating
each transverse wave number k x with a particular axial wave number k z, a reduced-dimensional
axial dispersion relationship ωred.(k z) is obtained: ω (k x,k z)=ω (k x(k z),k z)7→ωred.(k z), which can
be engineered almost arbitrarily. In the experiment reported here, we employed this strategy to
produce a linear dispersion relationship ω (k z)=(k z − ko)ve projected onto the (k z, ωc )-plane that
deviates away from the light-line ω =ck z. In presence of anomalous GVD, such a spatio-temporal
spectrum is delimited at both ends by the curved light-line in the dispersive medium – thereby
yielding the circular symmetric spatio-temporal profile characteristic of dBM wave packets. Here,
the transverse wave number k x played the role of the observer velocity u in the physical con-
figuration envisioned by Mackinnon [Fig. 1(d)]. However, one may envision a variety of other
scenarios that can be facilitated by this general strategy. For example, besides tuning the group
velocity in free space, linear dispersive media, or nonlinear optical materials and structures, one
may produce accelerating wave packets [46–48] whose group velocity changes with propagation
in such media. These features have been recently predicted to produce a host of new phenomena
related to two photon-emission [49] and relativistic optics [50, 51].
Intriguingly, the strategy employed here is not constrained to optical waves. Indeed, our
approach to spatio-temporal structuring of the field is agnostic with respect to the physical

13
substrate, and can be implemented in principle with acoustic waves, microwaves, surface plasmon
polaritons [52], electron beams, neutron beams, or other massive particles. In all cases, an added
spatial dimension can be exploited to override the intrinsic dispersion relationship of the particular
wave phenomenon, thus producing novel propagation dynamics.

The dimensionality of the (2 + 1)D dBM wave packets synthesized here can be further extended
to the full-dimensional (3 + 1)D space of ( x,y,z;t) by including the second transverse coordinate
y. This can now be achieved in light of very recent progress in producing so-called 3D ST
wave packets that are localized in all dimensions of (3 + 1)D space [53–55]. Combining this new
synthesis methodology with the procedure outlined here for producing dBM wave packets in
the anomalous-GVD regime will yield spherically symmetric propagation-invariant pulsed field
structures. Such field configurations provide a platform for exploring proposed topological
structures associated with polarization (spin texture) [53] without resorting to stereo-projection
onto a 2D plane. Moreover, such spherically symmetric optical dBM wave packets are compatible
with coupling to optical fibers and waveguides, thus enabling new opportunities in optical
communications, optical signal processing, and nonlinear and quantum optics.

Finally, the ideal spectral constraint underlying optical dBM wave packets implies an exact
association between the spatial and temporal frequencies. Such idealized wave packets con-
sequently have infinite energy [56]. In any realistic system, however, a spectral uncertainty is
inevitably introduced into this association, resulting in a finite-energy wave packet traveling for
a finite distance over which it is approximately invariant [57]. In our experiments, this spectral
uncertainty arises from the finite spectral resolution of the diffraction grating employed (Fig. 3),
which is estimated to be ≈16 pm, corresponding to a propagation distance of ≈32 m at a spectral
tilt angle θ = 44.99◦ [42].

[1] M. V. Berry and N. L. Balazs, “Nonspreading wave packets,” Am. J. Phys. 47, 264–267 (1979).
[2] D. M. Greenberger, “Comment on “Nonspreading wave packets”,” Am. J. Phys. 48, 256 (1980).
[3] G. A. Siviloglou and D. N. Christodoulides, “Accelerating finite energy Airy beams,” Opt. Lett. 32,
979–981 (2007).
[4] P. Zhang, T. Li, J. Zhu, X. Zhu, S. Yang, Y. Wang, X. Yin, and X. Zhang, “Generation of acoustic
self-bending and bottle beams by phase engineering,” Nat. Commun. 5, 4316 (2014).

14
[5] S. Fu, Y. Tsur, J. Zhou, L. Shemer, and A. Arie, “Propagation dynamics of Airy water-wave pulses,”
Phys. Rev. Lett. 115, 034501 (2015).
[6] N. Voloch-Bloch, Y. Lereah, Y. Lilach, A. Gover, and A. Arie, “Generation of electron Airy beams,”
Nature 494, 331–335 (2013).
[7] I. Kaminer, J. Nemirovsky, M. Rechtsman, R. Bekenstein, and M. Segev, “Self-accelerating Dirac
particles and prolonging the lifetime of relativistic fermions,” Nat. Phys. 11, 261–267 (2015).
[8] L. Mackinnon, “A nondispersive de Broglie wave packet,” Found. Phys. 8, 157–176 (1978).
[9] L. de Broglie, “Recherches sur la théorie des quanta,” Ann. de Phys., 3, 22 (1925).
[10] C. Cohen-Tannoudji, B. Diu, and F. Laloe, Quantum Mechanics (Wiley, 1991).
[11] J. P. Torres, M. Hendrych, and A. Valencia, “Angular dispersion: an enabling tool in nonlinear and
quantum optics,” Adv. Opt. Photon. 2, 319–369 (2010).
[12] J. A. Fülöp and J. Hebling, “Applications of tilted-pulse-front excitation,” Recent Optical and Photonic
Technologies , 207–230 (2010).
[13] P. Saari and K. Reivelt, “Generation and classification of localized waves by Lorentz transformations
in Fourier space,” Phys. Rev. E 69, 036612 (2004).
[14] M. Zamboni-Rached and E. Recami, “Subluminal wave bullets: Exact localized subluminal solutions
to the wave equations,” Phys. Rev. A 77, 033824 (2008).
[15] M. Yessenov, L. A. Hall, K. L. Schepler, and A. F. Abouraddy, “Space-time wave packets,”
arXiv:2201.08297 (2022).
[16] B. E. A. Saleh and M. C. Teich, Principles of Photonics (Wiley, 2007).
[17] H. E. Kondakci and A. F. Abouraddy, “Diffraction-free pulsed optical beams via space-time correla-
tions,” Opt. Express 24, 28659–28668 (2016).
[18] K. J. Parker and M. A. Alonso, “The longitudinal iso-phase condition and needle pulses,” Opt. Express
24, 28669–28677 (2016).
[19] L. J. Wong and I. Kaminer, “Ultrashort tilted-pulsefront pulses and nonparaxial tilted-phase-front
beams,” ACS Photon. 4, 2257–2264 (2017).
[20] M. Yessenov, B. Bhaduri, H. E. Kondakci, and A. F. Abouraddy, “Weaving the rainbow: Space-time
optical wave packets,” Opt. Photon. News 30, 34–41 (2019).
[21] M. A. Porras, S. Trillo, C. Conti, and P. Di Trapani, “Paraxial envelope X waves,” Opt. Lett. 28,
1090–1092 (2003).

15
[22] M. A. Porras, G. Valiulis, and P. Di Trapani, “Unified description of Bessel X waves with cone
dispersion and tilted pulses,” Phys. Rev. E 68, 016613 (2003).
[23] S. Longhi, “Localized subluminal envelope pulses in dispersive media,” Opt. Lett. 29, 147–149 (2004).
[24] M. A. Porras and P. Di Trapani, “Localized and stationary light wave modes in dispersive media,”
Phys. Rev. E 69, 066606 (2004).
[25] S. Malaguti, G. Bellanca, and S. Trillo, “Two-dimensional envelope localized waves in the anomalous
dispersion regime,” Opt. Lett. 33, 1117–1119 (2008).
[26] S. Malaguti and S. Trillo, “Envelope localized waves of the conical type in linear normally dispersive
media,” Phys. Rev. A 79, 063803 (2009).
[27] L. A. Hall and A. F. Abouraddy, “Canceling and inverting normal and anomalous group-velocity
dispersion using space-time wave packets,” arXiv:2202.01148 (2022).
[28] H. E. Kondakci and A. F. Abouraddy, “Optical space-time wave packets of arbitrary group velocity in
free space,” Nat. Commun. 10, 929 (2019).
[29] B. Bhaduri, M. Yessenov, and A. F Abouraddy, “Anomalous refraction of optical spacetime wave
packets,” Nat. Photon. 14, 416–421 (2020).
[30] L. A. Hall and A. F. Abouraddy, “Universal angular-dispersion synthesizer,” arXiv:2109.13987 (2021).
[31] P. Saari and K. Reivelt, “Evidence of X-shaped propagation-invariant localized light waves,” Phys.
Rev. Lett. 79, 4135–4138 (1997).
[32] J. Turunen and A. T. Friberg, “Propagation-invariant optical fields,” Prog. Opt. 54, 1–88 (2010).
[33] H. E. Hernández-Figueroa, E. Recami, and M. Zamboni-Rached, eds., Non-diffracting Waves (Wiley-
VCH, 2014).
[34] A. Chong, W. H. Renninger, D. N. Christodoulides, and F. W. Wise, “Airy–Bessel wave packets as
versatile linear light bullets,” Nat. Photon. 4, 103–106 (2010).
[35] E. MacKinnon, “De Broglie’s thesis: A critical retrospective,” Am. J. Phys. 44, 1047–1055 (1976).
[36] J. M. Espinosa, “Physical properties of de Broglie’s phase waves,” Am. J. Phys. 50, 357–362 (1982).
[37] R. Donnelly and R. W. Ziolkowski, “Designing localized waves,” Proc. R. Soc. Lond. A 440, 541–565
(1993).
[38] P. A. Bélanger, “Lorentz transformation of packetlike solutions of the homogeneous-wave equation,”
J. Opt. Soc. Am. A 3, 541–542 (1986).
[39] S. Longhi, “Gaussian pulsed beams with arbitrary speed,” Opt. Express 12, 935–940 (2004).

16
[40] H. E. Kondakci and A. F. Abouraddy, “Airy wavepackets accelerating in space-time,” Phys. Rev. Lett.
120, 163901 (2018).
[41] F. Wilczek, A Beautiful Question: Finding Nature’s Deep Design (Penguin Press, 2015).
[42] M. Yessenov, L. Mach, B. Bhaduri, D. Mardani, H. E. Kondakci, M. A. Alonso, G. A. Atia, and A. F.
Abouraddy, “What is the maximum differential group delay achievable by a space-time wave packet
in free space?” Opt. Express 27, 12443–12457 (2019).
[43] H. E. Kondakci and A. F. Abouraddy, “Diffraction-free space-time beams,” Nat. Photon. 11, 733–740
(2017).
[44] K. Unnikrishnan and A. R. P. Rau, “Uniqueness of the Airy packet in quantum mechanics,” Am. J.
Phys. 64, 1034–1035 (1996).
[45] A. Forbes, M. de Oliveira, and M. R. Dennis, “Structured light,” Nat. Photon. 15, 253–262 (2021).
[46] M. Yessenov and A. F Abouraddy, “Accelerating and decelerating space-time optical wave packets in
free space,” Phys. Rev. Let. 125, 233901 (2020).
[47] L. A. Hall, M. Yessenov, and A. F. Abouraddy, “Arbitrarily accelerating space-time wave packets,”
Opt. Lett. 47, 694–697 (2022).
[48] Z. Li and J. Kawanaka, “Velocity and acceleration freely tunable straight-line propagation light bullet,”
Sci. Rep. 10, 11481 (2020).
[49] J. Sloan, N. Rivera, J. D. Joannopoulos, and M. Soljačić, “Controlling two-photon emission from
superluminal and accelerating index perturbations,” Nat. Phys. 18, 67–73 (2022).
[50] K. Y. Bliokh and F. Nori, “Spatiotemporal vortex beams and angular momentum,” Phys. Rev. A 86,
033824 (2012).
[51] C. Caloz and Z.-L. Deck-Léger, “Spacetime metamaterials–Part I: General concepts,” IEEE Trans.
Antennas Propag. 68, 1569–1582 (2020).
[52] K. L Schepler, M. Yessenov, Y. Zhiyenbayev, and A. F Abouraddy, “Space–time surface plasmon
polaritons: A new propagation-invariant surface wave packet,” ACS Photon. 7, 2966–2977 (2020).
[53] C. Guo, M. Xiao, M. Orenstein, and S. Fan, “Structured 3D linear space-time light bullets by nonlocal
nanophotonics,” Light Sci. Appl. 10, 160 (2021).
[54] K. Pang, K. Zou, H. Song, M. Karpov, M. Yessenov, Z. Zhao, A. Minoofar, R. Zhang, H. Song, H. Zhou,
X. Su, N. Hu, T. J. Kippenberg, A. F. Abouraddy, M. Tur, and A. E. Willner, “Synthesis of near-
diffraction-free orbital-angular-momentum space-time wave packets having a controllable group

17
velocity using a frequency comb,” Opt. Express 30, 16712–16724 (2022).
[55] M. Yessenov, J. Free, Z. Chen, E. G. Johnson, M. P. J. Lavery, M. A. Alonso, and A. F. Abouraddy,
“Space-time wave packets localized in all dimensions,” arXiv:2111.03095 (2021).
[56] A. Sezginer, “A general formulation of focus wave modes,” J. Appl. Phys. 57, 678–683 (1985).
[57] H. E. Kondakci, M. A. Alonso, and A. F. Abouraddy, “Classical entanglement underpins the invariant
propagation of space–time wave packets,” Opt. Lett. 44, 2645–2648 (2019).

Data availability statement


The data that support the plots within this paper and other findings of this study are available
from the corresponding author upon reasonable request.

Acknowledgments
We thank M. Yessenov, K. L. Schepler, D. N. Christidoulides and A. Dogariu for helpful discus-
sions. This work was supported by the U.S. Office of Naval Research (ONR) under contracts
N00014-17-1-2458 and N00014-20-1-2789.

Author contributions

Correspondence and requests for materials should be addressed to A.F.A.


(email: raddy@creol.ucf.edu)

Competing interests: The authors declare no competing interests.

18
METHODS

de Broglie phase waves. At rest, the frequency of the internal ‘clock’ and that of the infinite-
wavelength phase wave are the de Broglie frequency ωo. When the particle moves at a velocity
p
v, the observed frequency of the internal ‘clock’ in the rest frame is reduced to ω =ωo 1 − β2v
p
whereas that of the phase wave is increased to ω =ωo/ 1 − β2v, where β v=v/c []. To obtain
The phase velocity vphh of the phase wave, de Broglie proposed a ‘theory of phase harmony’,
which requires that the internal clock and the phase wave remain in phase for all t and at any
p
v []. The phase of the moving clock in the rest frame is φ=ωot 1 − β2v, and that of the phase
wave is φ=ωo(t − vz )/ 1 − β2v. At time t, the particle has covered a distance z=vt, and equating
p
ph
2
the phases yields vph= cv >c. Alternatively, the Lorentz transformation of the proper time is
2
t0=(t − cv2 z)/ 1 − β2v, ωot0=ωt − kz=ωt − vω z, from which we again have vph= cv .
p
ph
p
Conventional de Broglie wave packets. The equations ω =ωo/ 1 − β2v and k = vω = ωc β for the
ph

phase wave yield the dispersion relationship k= 1c ω 2 − ωo2. A de Broglie wave packet of finite
p

bandwidth ∆ω centered at ω =ωc requires including an uncertainty ∆v in the particle velocity


centered on the speed v (which corresponds to ωc). The group velocity of the de Broglie wave
 dk q
packet is thus ve=1 dω =c 1 −( ωo )2=cβ v=v.
ω c ωc
Formulation of de Broglie-Mackinnon (dBM) wave packets. Consider the configuration de-
picted in Fig. 1(d), where the particle moves at a velocity v and an observer moves at a velocity u,
both with respect to a selected rest frame. Here the relative velocity of the particle with respect
β −β
to the observer is ξ, where β ξ = ξc = 1−v β v βuu , β v= vc , and β u= uc . According to this observer, the
q q
frequency and wave number are ω 0=ωo 1 − β2ξ and k0=ko β ξ 1 − β2ξ , respectively. The crucial
step proposed by Mackinnon is that all potential observers with velocities u ranging from −c to c
report their observations of ω 0 and k0 to the common rest frame, where the wave packet is con-
structed epistemologically after accounting for Lorentz contraction and time dilation. Following
this prescription, it is straightforward to show that

1 − βvβu βv − βu
ω = ωo p , k = ko p . (4)
1 − β2v 1 − β2v

Because v is a fixed velocity whereas u extends from −c to c, a linear dispersion relationship


between ω and k is established [Fig. 1(f)],
 
1
q
k= ω
c − ko 1 − β2v . (5)
βv

19
q
1+ β
This line intersects with the light-line k = ωc at k =k +=ko 1− β vv (when u=−c), and with the light-
q
1− β
line k =− ωc at k =−k −=−ko 1+ β vv (when u=c), where k += ωc+ >ko and k −= ωc− <ko [Fig. 1(f)].
When u=0, the associated de Broglie phase wave has ω
c =k1 and k= β vk1, where k1= √ ko > k o.
1− β2v
Another phase wave of interest is the one that retains the stationary frequency ω =ωo, which
occurs when β u= β1v (1 − 1 − β2v)= β1v (1 − kko ), and is associated with k =k2=ko β u. Finally, the
p
1
p k2
linear dispersion relationship has k =0 when u=v and thus ωc =ko 1 − β2v= ko <ko. Throughout,
1

setting v=0 (β v=0) in these relationships yields the result in Fig. 1(e) for a dBM wave packet
associated with a particle in its rest frame.
Optical dBM wave packets in free space. For a monochromatic optical field at ω =ωo, we have
the dispersion relationship k2x + k2z =k2o, which is the circle at the intersection of the free-space
light-cone k2x + k2z =( ωc )2 with the horizontal iso-frequency plane ω =ωo. As described in the main
text, the full circle is the spectral support for the optical field produced by a stationary planar
dipole. When the source and detector are in relative motion along the z-axis at a velocity v, the
initially horizontal iso-frequency-plane in (k x,k z, ωc )-space is tilted by an angle θ with respect
to the k z-axis,where tanθ = β v [40]. The resulting spectral constraint at the intersection with the
light-cone (that conforms to a dBM wave packet) is:

ω − ω+ ω k2o
   
1 1
q
ω
kz = k+ + = − ko 1 − β2v = − , (6)
ve βv c βv c k1

where k += ωc+ is the point on the light-line k z= ωc intersecting with the tilted spectral plane, and
we make use of the same definitions of ko, k +, and k1 as above for dBM wave packets. Eliminating
k z from these two relationships (the light-cone and the titled spectral plane) yields the spectral
projection onto the (k x, ωc )-plane in the form of an ellipse (Eq. 1) [Fig. 2(b)]. We can also obtain
p
the spectral projection onto the (k x,k z)-plane by eliminating ω: substituting ωc = β vk z + ko 1 − β2v
k2x
from the spectral constraint into k2x + k2z =( ωc )2 yields the ellipse k2o
+ (∆k 1/2)2 (k z − β vk1)2=1, where
z

∆k z=2k1, and β vk1 is the central axial wave number. Substituting k x= ωc


sin ϕ and k z= ωc cos ϕ, the
p
propagation angle is cos ϕ(ω )={1 − ωωo 1 − β2}/β v, ω 6=0, which is not differentiable at ϕ=0 or
ϕ=π – corresponding to the maximum and minimum points on the ellipses in the (k x, ωc ) or
(k x,k z) planes.
Optical dBM wave packets in presence of anomalous GVD. In presence of anomalous GVD,
vm − 21 |k2m|Ω2;
the wave number in the medium expanded around ωo is k=n(ω )ω/c=nmko + Ω/e
here Ω=ω − ωo and n(ω ) is the frequency-dependent refractive index. The quantities nm, vem, and

20
 dk
k2m are all evaluated in the medium at ω =ωo: nm=n(ωo) is the refractive index; ve=1 dω is the
ωo
d2k

group velocity, and k2m= dω2 ω =−|k2m| is the negative-valued GVD coefficient in the anomalous
o

dispersion regime. In the small-angle (paraxial) approximation,

Ω 1 k2x
q
kz = k2 − k2x ≈ nmko + − |k2m|Ω2 − , (7)
vem 2 2nmko

and the light-line k x=0) is now curved [Fig. 2(c)]. To produce a dBM wave packet, we impose the
spectral constraint

k z = nmk o + , (8)
ve
1− β v 0
which intersects with the light-line at two points: Ω=0 (ω =ωo) and ω
c =−2k a β0v ; here β0v=
ve
ve<vem, ka=(c2|k2m| β m)−1= ωca . The maximum temporal bandwidth compatible with
βv
β m = vem ,
0
∆ω 1− β v
dispersion-free propagation in this dispersive medium at a group velocity ve is thus c =2k a β0v ,
1− β 0
and the corresponding maximum bandwidth of the axial wave number is ∆k z=2 βkma β02 v , so that
v
∆ω/c 0
∆k z = β m β v= β v.
We can now obtain the spectral projections onto the (k x, ωc ) and (k z, ωc ) planes for the optical
dBM wave packet in presence of anomalous GVD just as we did for their counterparts in free
space. By equating Eq. 7 and Eq. 8 we eliminate k z and obtain in the (k x, ωc )-plane the ellipse
k2x 2
∆ω/2 2
k2x,max
− ωc )
+ ((ω∆ω/2 )2
=1, where the central frequency is ωc=ωo − ∆ω 2
2 , k x,max= σm( c ) , and σm=
nmcωo|k2m| is a dimensionless GVD parameter. Similarly, we can obtain the spectral projection
onto the (k x,k z)-plane by substituting Ω=ve(k z − nmko) from the spectral constraint in Eq. 8 into
k2x 2
k z − k c)
the light-cone in Eq. 7 to obtain the ellipse: 2
k x,max
+ ((∆k /2)2
=1, where kc is the center of the k z-span.
z

Details of the experimental setup and spectral measurements. The field configuration shown
in Fig. 2(c) is produced via the angular-dispersion synthesizer depicted in Fig. 3 []. Starting
with femtosecond laser pulses (central wavelength λ=1064 nm, bandwidth ∆λ=20 nm, and
pulsewidth ∆T ≈100 fs; Spark Lasers, Alcor), a diffraction grating (1200 lines/mm) resolves the
spectrum spatially, and a cylindrical lens (focal length f =500 mm) collimates the wave front
before incidence on a reflective, phase-only SLM (Meadowlark, E19X12). The SLM deflects each
wavelength λ at angles ± ϕ(λ) according to the elliptical spatio-temporal spectrum given in
Eq. 3. The wave front retro-reflected from the SLM returns to the grating, and the optical dBM
wave packet is formed. The spatio-temporal spectrum is acquired by directing a portion of the
spectrally resolved wave front reflecting back from the SLM to a spatial Fourier transform (not
shown in Fig. 3 for simplicity), which thus yields the spatio-temporal spectral projection onto the

21
(k x,λ)-plane. We then obtain the spectral projection onto the (k z,λ)-plane in the medium via the
q
relationship k z(ω )= {(n(ω ) ωc }2 − k2x(ω ).
Dispersive medium. The dispersive sample that we exploit comprises a pair of chirped Bragg
mirrors (Edmund 12-335) that provide anomalous group-delay dispersion (GDD). By adjusting
the separation between the two mirrors and the incident angle of the wave packet, we can
control the number of bounces off the mirrors, thereby producing an anomalous-GVD coefficient
of k2m≈−500 fs2/mm. Because the thickness of the mirrors is negligible with respect to the
free-space gap separating them, we can thus have vem≈c (e
nm≈1).
Reconstruction of the spatio-temporal profiles of dBM wave packets. To reconstruct the spatio-
temporal intensity profile of a dBM wave packet I ( x,z;τ ) at a fixed axial plane z, we make use of
the interferometric arrangement depicted schematically in Fig. 3. We bring together two wave
packets: the synthesized optical dBM wave packet, and a reference plane-wave pulse taken from
the initial laser pulse [28]. An optical delay τ is placed in the path of the reference pulse. When the
dBM wave packet and the reference pulse overlap in space and time, we observe high-visibility
spatially resolved interference fringes at the CCD camera placed in their common path. As we
sweep the optical delay τ (thus reducing the overlap between the dBM wave packet and the
reference pulse), the interference visibility drops. We make use of the recorded visibility along x
and tau to reconstruct the wave packet intensity profile I ( x;τ ).

22
FIG. 1. de Broglie phase waves and wave packets, and de Broglie-Mackinnon (dBM) wave packets. (a)
In the rest frame of a particle, the internal ‘clock’ and the external ‘phase wave’ theorized by de Broglie
both have the same frequency ωo. (b) When the particle moves at a velocity v along z, the frequency of
p
the internal clock in the rest frame decreases to ω =ωo 1 − β2v, whereas that of the phase wave increases
p
to ω =ωo/ 1 − β2v. (c) The dispersion relationship for de Broglie phase waves ω 2=ωo2 + c2k2 plotted in
(k, ωc )-space. The group velocity evaluated at ω =ωc is ve=v. Constructing a localized de Broglie wave
packet necessitates introducing an ad hoc uncertainty in the particle velocity. (d) The physical setting for a
dBM wave packet. The particle travels at v and the observer at u (both along the z-axis) with respect to a
common rest frame. (e) The dispersion relationship for a dBM wave packet in (k, ωc )-space (lower panel)
corresponding to a stationary particle v=0, delimited by the light-lines |k |= ωc . The observer velocity u
(upper panel) is an internal parameter swept from −c to c to produce the dBM dispersion relationship. (g)
q q
1+ β v 1− β
Same as (f) for v6=0; here k +=ko 1− β v , k −=ko 1+ β vv , k1= √ ko 2 , and k2= βkov (1 − kko1 ); see Methods
1− β v

23
FIG. 2. Optical de Broglie-Mackinnon (dBM) wave packets. From left to right we depict the following:
the light-cone in (k x,k z, ωc )-space intersecting with a spectral constraint in the form of a plane; the spectral
projection onto the (k z, ωc )-plane; the spectral projection onto the (k x, ωc )-plane; the propagation angle
ϕ(ω ); and the real part of the spatio-temporal field profile ψ( x,z;t) at a fixed axial plane z. (a) A stationary
monochromatic planar dipole in free space resulting from the constraint ω =ωo. (b) A moving planar
2
dipole in free space corresponding to the constraint k z= β1v ( ωc − kko1 ). (c) The field in a dispersive medium in
the anomalous regime after imposing the constraint k z= β1v { ωc − ko(1 − nm β v)}.

24
FIG. 3. Synthesizing and characterizing optical dBM wave packets. Starting with femtosecond laser
pulses, the space-time (ST) synthesis arrangement associates each wavelength λ with prescribed propaga-
tion angles ± ϕ(λ) before traversing the dispersive medium [30]. Interfering the synthesized optical dBM
wave packets with reference plane-wave pulses from the initial laser helps reconstruct the spatio-temporal
intensity profile of the dBM wave packets.

25
FIG. 4. Observation of optical dBM wave packets in presence of anomalous GVD and tuning their
group velocity. In the first column we plot the measured spatio-temporal spectrum projected onto the
(k x,λ)-plane; the dotted curve is the theoretical expectation based on Eq. 3. In the second column we plot
the spectral projection onto the (k z,λ)-plane using the data from the first column; the dotted curve is the
dispersive light-line. In the third column we plot the spatio-temporal intensity profile I ( x;τ ) acquired
in a moving reference frame traveling at the group velocity of the wave packets; the dotted circles are
guides for the eye. (a) Measurements for an optical dBM wave packet having a group velocity ve=0.9975c;
(b) ve=0.9985c; and (c) ve=0.999c. The dimensionaless dispersion coefficient is σm=cωok2m = 0.3, and the
measurements are carried out at z=30 mm.

26
FIG. 5. Tuning the equivalent rest-mass of an optical dBM wave packets. The columns correspond
to those in Fig. 4. The dimensionless dispersion coefficient is σm=cωok2m = 0.3, the measurements are
all acquired at z=30 mm, and the group velocity is held fixed at ve= 0.9975c. (a) The short-wavelength
intersection with the dispersive light-line is λo=1054 nm; (b) λo=1055 nm; and (c) λo=1056 nm.

27
FIG. 6. Changing the spatio-temporal structure of optical dBM wave packets. The columns from left
to right are: the measured spatio-temporal spectrum projected onto the (k x,λ)-plane, along with the
spectral phase; the measured spatio-temporal intensity profile I ( x,z;τ ) at z=15 mm; the measured spatio-
temporal intensity profile I ( x,z;τ ) at z=30 mm; the calculated spatio-temporal intensity profile I ( x,z;τ )
at fixed z; I ( x;τ = 0) at z=30 mm; and I ( x = 0;τ ) at z=30 mm. The dimensionless dispersion coefficient is
Measurement of an O shaped spectrum with v a = 0.9975c and σm=cωok2m = 0.3, and the group velocity is
held fixed at ve=0.9975c, corresponding to Fig. 4(a). (a) The spectrum has uniform phase; (b) the spectrum
has uniform phase but its amplitude is truncated along λ; (c) a π-phase step is introduced along k x; (d) a
π-phase step is introduced along λ; and (e) the spectral phase is alternated between 0 and π in the four
quadrants of the (k x,λ)-plane.

28

You might also like