文献1
文献1
文献1
Topological photonics in
synthetic dimensions
Eran Lustig1,2 AND Mordechai Segev1,2,3, *
1
Solid State Institute, Technion, Haifa 32000, Israel
2
Physics Department, Technion, Haifa 32000, Israel
3
Electrical Engineering Department, Technion, Haifa 32000, Israel
* Corresponding author: [email protected]
Received December 21, 2020; revised March 30, 2021; accepted March 30, 2021;
published June 1, 2021 (Doc. ID 418074)
Topological photonics is a new and rapidly growing field that deals with topological
phases and topological insulators for light. Recently, the scope of these systems was
expanded dramatically by incorporating non-spatial degrees of freedom. These syn-
thetic dimensions can range from a discrete ladder of cavity modes or Bloch modes of
an array of waveguides to a time-bin division (discrete time steps) in a pulsed system
or even to parameters such as lattice constants. Combining spatial and synthetic dimen-
sions offers the possibility to observe fundamental and exotic phenomena such as
dynamics in four dimensions or higher, long-range interaction with disorder, high-
dimensional nonlinear effects, and more. Here, we review the latest developments in
using non-spatial dimensions as a means to enhance fundamental features of photonic
topological systems, and we attempt to identify the next challenges. c 2021 Optical
Society of America
https://doi.org/10.1364/AOP.418074
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
1.1 Topological Insulators for Electrons . . . . . . . . . . . . . . . . . . 427
1.2 Photonic Topological Insulators . . . . . . . . . . . . . . . . . . . . 429
1.3 Role of Dimensionality and Connectivity in Topological Physics . . . 429
1.4 Historic Development of Synthetic Dimensions in Photonics . . . . . 430
2. Basic Physics of Topology in Synthetic Dimensions . . . . . . . . . . . . . 433
2.1 Topological Phases and Topological Invariants . . . . . . . . . . . . 433
2.2 Richness of Topological Realm . . . . . . . . . . . . . . . . . . . . 435
3. Description and Definition of Non-Spatial Dimensions . . . . . . . . . . . . 436
3.1 Topological Photonics in Parameter Space . . . . . . . . . . . . . . . 437
3.2 Synthetic Space Relying on Modal Ladders . . . . . . . . . . . . . . 441
3.3 Using Time-Bins as Synthetic Dimension . . . . . . . . . . . . . . . 446
4. Future Directions and Vision . . . . . . . . . . . . . . . . . . . . . . . . . 450
Funding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
Disclosures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
References and Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 427
Topological photonics in
synthetic dimensions
Eran Lustig AND Mordechai Segev
1. INTRODUCTION
1.1 Topological Insulators for Electrons
In the past 15 years, topological insulators have been attracting much research interest
in many areas of physics [1–5]. Unlike ordinary materials—conductors, semiconduc-
tors, and insulators—topological insulators can conduct on their edges and insulate
in their bulk, but what makes them so special is that the current on their edge is com-
pletely indifferent to the shape of the material: it is unidirectional (chiral) and robust
against most types of disorder. In fact, these materials are considered to be a phase
of matter of their own, with features dictated by the topological properties of the
conducting and insulating bands that are usually defined on the reciprocal space
[6]. Consequently, the bands of topological insulators are classified according to
their topology—which is an abstract way to count the number of twists, holes, and
other types of geometric features that cannot continuously deform from one to the
other. That is, the allowed states for electrons can be classified through topological
invariants, determined by externally induced fields or internally through interactions
between the particles, and by the structural symmetries in the medium. The first con-
nection between conductance and topology was in the integer quantum Hall effect
(IQH) [7]. In the IQH, a magnetic field applied to a 2D electron gas caused its Hall
conductance to have integer values. It was then realized that the integer values of the
conductance are related to the topology of the allowed states for electrons [6,8,9], and
that the classification of the different bands of states is related to robust edge states
and a topological invariant called the Chern number [10]. In the following years, it
was found that topological insulators can occur even without a net magnetic flux [11]
and, even more surprisingly, without any external fields and without even breaking
time-reversal symmetry. This happens due to the spin of the electrons themselves (the
quantum spin-Hall effect) or via the spin–orbit coupling [4,12,13]. This notion was
considered a major breakthrough some 15 years ago since it meant that the properties
of topological insulators that were thought to be extrinsic can actually be intrinsic in
certain materials. In these intrinsic topological insulators, pairs of edge states related
to different spins propagate in opposite directions, but still each one of them is chiral.
Moreover, these topological edge states do not couple to each other as long as the
time-reversal symmetry is maintained and the gap remains open. In such cases, the
topological protection is dependent, in principle, not only on the gap remaining open
but also on the existence of a symmetry of the system, in this case: time-reversal
symmetry.
These breakthroughs have opened the door for finding many materials with different
topological phases of matter with exotic properties. Due to their unique properties,
topological materials can potentially be utilized in quantum computing [14], spin-
tronics [15], and more. The relation between topological materials and applications
in quantum computing involves additional notions such as fractional Hall states or
Majorana fermions that usually require fermionic interactions and, hence, are beyond
the scope of this review—which deals mostly with photonics. We are now almost four
decades after the discovery of the first topological insulator (the IQH). To understand
the underlying principles, it is instructive to briefly review the basics of topological
428 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
Figure 1
three dimensions [2–4,12,17] or more than three dimensions [45–49], these phases
are very hard to realize in photonic systems. This means that, in all our efforts so far
with traditional photonic lattices and structures, we were always only examining the
tip of the iceberg, in terms of the diversity of topological phases with different proper-
ties. The reason is twofold: First, photons propagate at the speed of light; thus, in the
direction of their propagation, it is hard to induce the complex dynamics necessary
for topological phases in high dimensions. The second reason is the complexity of
fabricating 3D topological structures, such as the 3D structures that were recently
used for demonstrating Weyl points [50–52] and the 3D topological insulator in the
microwave regime using magneto-optic effects [53]. 3D fabrication becomes much
more complicated for higher frequencies in the optical regime, where the length
scales are much smaller (implying that higher resolution in fabrication is crucial) and
magneto-optic effects are very weak. Of course, systems of dimensions higher than
three cannot be fabricated directly and require special approaches.
Another aspect in which most photonic topological structures are limited is their
connectivity. Designing a photonic structure that corresponds to a topological phase
usually requires the construction of a lattice, which is a periodic arrangement of
optical elements in space [54–57]. Generally, lattices play an important role in many
topological insulators, as the symmetries and the parameters of the lattice dictate the
topological behavior directly. Since our world has only three spatial dimensions, and
the coupling between sites is mostly by proximity, this limits the number of possible
topological geometries and consequently possible topological phases [11,58], as
much as it does for the topological insulators found in condensed matter. That is, the
possibility of coupling two sites (or elements) that are far apart from one another is
limited in traditional photonic structures, which typically rely on evanescent coupling.
We stress that this problem of connectivity is not limited to high-dimensional systems
nor to photonics; it is a limitation common to all physical systems in real space and
in any dimension. Thus, a possible way to conclude the discussion so far could have
been that the major limiting factor in experimentally realizing lattice models in high
dimensions and with long-range connectivity is the traditional approach in which
lattice sites are identified by their location in space. Luckily, it turns out that this does
not have to be the case: it is possible to utilize non-spatial degrees of freedom such
as spin, momentum, energy, and time as “lattice degrees of freedom” [59–63]. These
non-spatial degrees of freedom are called “synthetic dimensions,” and they are not
restricted by spatial proximity; therefore, they offer new possibilities in experimen-
tally realizing topological phases that have thus far never been observed. Indeed,
recent years have witnessed the use of synthetic dimensions as a means to realize
novel topological phenomena in photonics, and even promising avenues to utilize this
concept for future applications. In this paper, we review the latest advancements on
topological photonics utilizing synthetic dimensions. We explain the fundamental
science underlying the ideas of synthetic dimensions in photonics, and we describe
the current three leading techniques for realizing synthetic dimensions and the recent
experimental efforts. Finally, based on the current state of the research, we map future
directions and possible applications for synthetic dimensions in photonics, and we
suggest some visionary ideas that may open new avenues in untraditional paths.
The downside of all of these is that the number of spin states and momentum states
is small, and generally excited states have a short lifetime, which altogether limit the
actual size of the synthetic dimension. But a more recent suggestion—of using modes
of a periodically shaken harmonic trap [81]—could, in principle, overcome those
limitations.
Slightly after the suggestion for implementing topological physics by utilizing syn-
thetic dimensions with cold atom systems [51], several groups proposed using this
concept in photonics. These groups suggested the use of a ladder of resonator modes
[82–84], coupled by external modulation, for inducing photonic artificial gauge
fields. And, more recently, photonic waveguide modes coupled by spatial modula-
tion with a spatially varying phase were employed to demonstrate the first photonic
topological insulator utilizing synthetic dimensions [85]. In a similar vein, magnetic
fields in synthetic dimensions were experimentally demonstrated in microwaves [86]
and on a single resonator with two independent mode ladders in optical wavelengths
[87,88]. These ideas, together with suggestions to exploit synthetic dimensions for
generating higher dimensional topological phases [83,89,90] and high-order topo-
logical phases [91], have recently become promising avenues for future research,
together with novel applications employing topologically protected edge transport for
robust unidirectional frequency conversion [92] and mode-locking an array of laser
emitters [93]. We discuss this approach in Section 2 below. Another advantage in
coupling ladders of modes as synthetic dimensions is the ability to couple modes that
are not close to one another. In this way, using synthetic dimensions can be designed
to introduce long-range coupling, which offers a means not only to enrich the topo-
logical phases [94] but also to increase their fundamental dimensionality [95]. In fact,
using long-range connectivity can be a type of synthetic dimension of its own, as was
proposed already in 2013 [63] in the context of demonstrating high-dimensional opti-
cal solitons, and it was recently used to demonstrate experimentally 4D topological
phases in a circuit implementation [96]. Furthermore, coupling synthetic degrees of
freedom does not necessarily require active modulation. A static geometry can also
induce a topological model utilizing synthetic dimensions by transforming not to an
eigenstate basis but to a non-diagonal basis [97–99].
Finally, the third approach for introducing a synthetic dimension has to do with dis-
crete dynamics in a synthetic space of “time-bins.” This approach employs laser
pulses propagating in a periodic cycle. At each period, each pulse splits into sev-
eral consecutive pulses, such that they arrive to the measurement device at different
times—due to different delay for each pulse. Each time slot of arrival (“time-bin”) is
analogous to a location in space. In this approach, the light changes its location (which
is the time-bin) constantly; therefore, this approach is endowed with a kinetic term
and displays transport in the synthetic dimension. Unlike the first two approaches,
here the propagation is only sampled at discrete times, associated with the steps of a
discrete time quantum walk (DTQW). Quantum walks have been known for almost
three decades to be a useful tool for studying a plethora of quantum and wave phe-
nomena [100–104]. In photonics, DTQWs were experimentally realized with short
pulses in fiber loops [105] and allowed the experimental realization of phenom-
ena such as parity–time (PT)-symmetric lattice [59], Kapitza light guiding [106],
time-reversed light [107], optical diametric drive [108], and more [109–112].
A decade ago, along with the growing interest in topological phases of matter, it
was realized that DTQW can exhibit topological phases, not only in 1D but also
in higher dimensions [113]. At first, 1D topological phases and topological edge
states in DTQW were experimentally demonstrated in photonics in several platforms
[114–118]. The 1D topological phases in DTQW are now being extended to include
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 433
exotic phenomena such as non-Hermitian topological edge states [119] and obser-
vation of the topological funneling of light in a non-Hermitian topological DTQW
[120]. In parallel, experimental efforts to realize 2D (or higher dimensional) synthetic
lattice models that can exhibit topological phases are ongoing. Since the basic time-
multiplexing system is zero-dimensional, the basic addition of time-bins converts it
to 1D only. Topological systems with higher dimensions can take several forms: they
can either be further extension of time-bins [121,122], or in combination with other
degrees of freedom such as orbital angular momentum [123].
More recently, time-bins were employed to demonstrate unidirectional topological
edge states in two dimensions. This was done both by extending the dimensionality
of the time-bins to the synthetic dimension [123,124] and by using an additional
synthetic dimension of angular momentum modes [125]. This platform is a promising
tool to combine synthetic gauge fields [124], non-Hermitian physics [59], and other
effects for photons to demonstrate exotic photonic topological phases. We focus on
this method and its topological aspects in Section 2 of this review.
deform the Hamiltonian H(k) from one to the other, without closing the gap—which
manifests a discontinuous change in the topological structure. For example, if our
Hamiltonian describes a 1D dimer (diatomic) lattice, i.e., a 1D lattice that has bonds
with strengths alternating between two values (say, u and v, where u is the bond inside
the unit cell and v is the bond between unit cells), then there can be two possible topol-
ogies (Z2 ) [126], where Z2 symbolizes a set of two integers{0,1}. These two possible
topologies are either a trivial topology (u > v), which maps to the topological struc-
ture in Fig. 2(a), or a non-trivial topology (u < v), which is isomorphic to Fig. 2(b). In
Figs. 2(a) and 2(b), the loop maps to the 1D Brillouin zone of the 1D dimer, and the
phase of the eigenstates maps to the arrow that winds around the loop. In Fig. 2(a), the
arrow does not wind around the loop (trivial), and, in Fig. 2(b), it winds once (non-
trivial). The two structures cannot continuously deform to one another and, therefore,
are topologically distinct (a structure that winds twice can continuously deform to
a structure that does not wind). Mathematically, what classifies the two topological
regimes is a topological invariant calculated on the 1D band. In a 1D dimer lattice, the
invariant is called the Zak phase, given by
Z
d
γ =i ψ(k) ψ(k) dk, (1)
k dk
where the integration is carried out on a 1D band in the Brillouin zone and the Zak
phase is equal to 0 or π (indicating the two possible topological phases). As can
be seen in Eq. (1), the Zak phase is more general than the dimer case and can be
calculated for any 1D lattice. The physical meaning is that, if two 1D lattices with dif-
ferent integer Zak phases border each other, this interface will host an exponentially
localized edge state in the gap. From Figs. 2(a) and 2(b) and Eq. (1), we can actually
understand that the Zak phase calculates the geometric phase acquired by the wave
function when it continuously slides along the Brillouin zone.
Interestingly, the picture looks completely different if we look at different dimensions
or different symmetries. For example, for a 2D square lattice with a perpendicular
magnetic field (the IQH) [10] or other lattices such as the extended Haldane model
[11,58], the topology can potentially be classified not to two subclasses (as in the
1D dimer case) but to infinitely many subclasses identified by integers Z. The topo-
logical structures map to structures of 2D vectors on a torus (band in the Brillouin
Figure 2
(a) and (b) Trivial and topological structures of a dimer lattice, respectively, com-
posed of a 1D loop with a vector on each point. The structures are mappings of the
topology of a band in a 1D dimer lattice. There are only two distinct topological
structures (adapted from [127]). (c) Topological structure of a 2D torus with vector
on each point (the projections of the vectors on the plane is presented). This structure
is a mapping of the topology of a band in a 2D Chern insulator. (d) “Periodic table”
of possible topological phases for quadratic Hamiltonians (see description of the
different symmetry classes in [45]).
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 435
zone) [Fig. 1(c)]. If we integrate along a loop that encircles the center, we can see that
the vector can wind and accumulate a geometric phase when completing the loop.
The geometric phase accumulated can only come in integer multiples of 2π, and it
is clear from the image why these cannot deform to one another without a disconti-
nuity. Mathematically and more generally, this is calculated by the Chern number.
The Chern number calculation is also an integral on the band, but—since this is a 2D
band—the integral is carried out over a closed path encircling the Brillouin zone,
I
C = hψ(k)|∇k |ψ(k)i dk, (2)
L
where L is a closed path in the 2D Brillouin zone. The possible topologies can be any
integer Z depending on the Hamiltonian, and this maps to different discontinuities
with different windings around them in Fig. 2(c). We note that the Chern number is
not the only topological invariant used to characterize a 2D topological structure;
rather, we give it here only because it is a commonly used one. The choice of topo-
logical invariants depends on the symmetries and dimensionality of the system, and
the Chern number commonly describes the topological structure in models such as
the IQH or the Haldane model, where the non-trivial topology arises from broken
time-reversal symmetry (driven by the magnetic field, in these cases). Of course, one
can go on and describe topological phases in higher dimensions that are described by
other structures.
extended Brillouin zone. For example, the Chern number [Eq. (2)] in a system with
one real spatial dimension k x and a synthetic space dimension kα is calculated in the
following way:
I
C= ψ( k̃) ∇k̃ ψ( k̃) d k̃), (3)
L̃
constant magnetic flux b, where one of its dimensions m has been Fourier transformed
to the momentum coordinate km . Substituting instead of φ, km gives
X
H(n, km ) = λc n† c n cos(2πbn + km ) + tc n† c n+1 + h.c .
(5)
n
Next, we Fourier transform km in our synthetic space Hamiltonian of Eq. (5) and
obtain the 2D Harper model, describing a square lattice with perpendicular magnetic
flux b,
X
H(n, m) = λe i2πbn c n,m
† †
c n,m+1 + tc n,m c n+1,m . (6)
n
The Brillouin zone of our fictitious 2D lattice is a surface in the coordinates (kn , km ),
and due to the magnetic flux b it has a topological invariant—the Chern number
[Eq. (2)].
As mentioned earlier, it is clear that 2D photonic systems are easier to fabricate and
to control experimentally than 3D ones, but this is not the only reason why using
synthetic dimensions increases the experimental possibilities. For example, in the AA
model just described, the magnetic field is now included in the parameter b, which
represents part of the prescribed spatial variation in the coupling constant. Thus, while
effective magnetic gauge fields are very hard to implement for photons in the physical
world because photons are non-interacting particles (and, thus, are not affected by
free-space EM fields), this model allows us to effectively do this simply by changing
the distances between lattice sites.
In the AA model, it is possible to calculate the Chern number in synthetic dimensions
with the following expression:
I
C= ψ( k̃ )|∇k̃ |ψ( k̃) d k̃, (7)
L̃
where k̃ = (kn , km ) = (νn , φ), and L̃ is a closed path encircling the extended Brillouin
zone in synthetic space.
Importantly, the reason to use synthetic dimensions is to observe physical phenomena
that are hard or impossible to realize in a real-space lattice. One such phenomenon,
relevant to the Harper model, is called quantum pumping, and it is deeply related to
the Chern number and edge states. Quantum pumps are the quantized transformation
of wave packets (quantum particles) through an adiabatic process [151,152]. For
photons, the process results in the transformation of all the light from one edge to the
other if the process is adiabatic, but disorder can cause some backscattering [153].
Due to the fact that this approach allows for demonstrating 2D physics with a mag-
netic field on a 1D lattice with just changing distances between sites, it was possible to
experimentally observe, for the first time in photonics, topological pumping of edge
states [55,61]. We will now explain how this was demonstrated experimentally in
more detail. The experiment was done in a transparent dielectric medium with index
of refraction n 0 + 1n(E r ), where 1n(E
r ) represents small variations in the refractive
index: |1n(E r )| n 0 . In this system, the light is propagating in the medium along
the optical axis z, with the linearly polarized field E (E r ) = ψ(E r )e −ik0 z e iωt (where
ω
k0 = c n 0 ). We assume the paraxial approximation, which means that light only
deviates by small angles from the optical axis: | ∂ψ | 2k0 | ∂ψ
2
∂z 2 ∂z
|. Substituting these
expressions into the Helmholtz equation yields the following propagation equation,
known as the paraxial wave equation:
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 439
∂ψ 1 2 k0
i = ∇⊥ ψ + 1n(E
r ) · ψ. (8)
∂z 2k0 n0
This equation is equivalent to the Schrodinger equation with two spatial dimensions,
where the optical axis z plays the role of time, nk00 1n(E r ) acts as the potential V (E
r ),
1 2
and 2k0 ∇⊥ is the kinetic term. To observe topological effects in this optical system,
one can use a lattice potential 1n(E r ), which is formed by an array of evanescently
coupled waveguides. Thus, by fabricating a lattice with distances that correspond
to the coupling in Eq. (4) [Fig. 3(a)], the wave dynamics is the same as the temporal
evolution of a single electron in an AA lattice. Scanning φ (which is equivalent to the
momentum in the synthetic space) is translated to scanning the 2D Brillouin zone with
a 1D line. This scanning reveals the unidirectional evolution of edge states associated
with the 2D Harper model system [Fig. 3(b)]. Thus, by adiabatically changing (in
z) the coupling along the propagation axis (equivalent to adiabatically varying the
hopping parameter in time), the photons are transferred (“pumped”) between different
regions of the topological band structure [Fig. 3(b)] and, consequently, move from
one edge of the lattice to the other edge, through the bulk of the lattice [Fig. 3(c)].
The topological pumping (also known as Thouless pumping) is a phenomena that is
not restricted to one type of topological lattice and also not to a single dimension. In
fact, by extending this idea to a 2D physical lattice, which is the vector product of two
1D AA spatial lattices, one can obtain a representation of 4D quantum Hall physics,
described by the Hamiltonian:
X
† †
H= tx c n,m c n+1,m + t y c n,m c n,m+1 + h.c, (9)
n,m
Figure 3
(a) (b) (c)
multiple edge states in Fig. 4, but also a second-order Chern number in the 4D space,
which corresponds to corner states. The explicit calculation of the 4D Chern number
in the synthetic dimension is
∂ Pj ∂ Pj ∂ Pj ∂ Pj
1
Z
Cj =− 2 εµξρσ Tr P j d 4 k, (10)
8π ∂kµ ∂kξ ∂kρ ∂kσ
where P j (k) is the projector onto the subspace below gap j spanned by the extended
Brillouin zone k̃ = {kn , km , kk , kl }. The second Chern number (as the first Chern num-
ber) is related to quantized pumping and edge states.
In this fashion, even though the system describes a plane in the 4D hyper-volume
at each moment, 4D topological pumping can be observed experimentally with a
2D paraxial photonic waveguide array [72,73] at each “instant” of the propagation
along the propagation axis. The light has a 2D band structure, which is a 2D plane in
the 4D band structure in parameter space. Consequently, by changing the coupling
distances, it is possible to topologically pump light along a line in this 4D parameter
space (Fig. 4).
This example is just one out of several experiments relying on synthetic dimensions
in parameter space. In this vein, treating a static parameter as the quantum momentum
was also used to experimentally demonstrate type-III Weyl semimetal [154], and
other parameters (such as magnetic field) were used to show photonic Weyl points
due to broken time-reversal symmetry [51] and exceptional surfaces [155]. While
the experiment described above is based on a space that is a hybrid real-synthetic
space, it is also possible to have a full synthetic parameter space. For example, it was
demonstrated that specific inner layer parameters, p and q of a 1D photonic crystal,
can form a 3D parameter space together with the wavenumber k. This parameter space
has experimentally measurable Weyl points and edge states [75]. A similar approach
was also used to observe exceptional cones in 4D parameter space [156].
To summarize, parametric synthetic dimensions represent a useful tool for demon-
strating topological phases in high dimensions, with the main advantage being that,
unlike other methods, here the complexity does not scale up with the dimensionality
Figure 4
2D lattice exhibiting the physics of 4D integer quantum Hall effect [72]: eigen-
energies of the system for φx = φ y . The orange and red wedges in the dispersion curve
are edge states (there are several due to different momentum in k x − k y ). The black
curves are corner states, which are not wedges in the dispersion curve.
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 441
will travel along the λ coordinate as it travels in space, i.e., we would like Htot to have
a kinetic term. As in every basis change, the transition to the basis λ can be done by
a unitary transformation. When transforming q Htot with Û = exp(itω D λ̂) and using
λ
the rotating wave approximation, ω D κ 8Mω , we obtain the following effective
Hamiltonian:
X X
Heff = 1 λ|λihλ| + J λ (|λ − 1ihλ|e iφ + h.c), (13)
λ λ
q
λ
where J λ is the effective coupling between modes and is equal to J λ = κ 8Mω . The
effective Hamiltonian obtained is that of a 1D lattice of modes [Fig. 5(a)] with an
effective electric field 1 in the direction of the synthetic axis, and a phase term in the
coupling: φ. At this stage, the effective electric field, which can be tuned by changing
the size of 1, affects the dynamics along the “modal dimension,” but the phase φ does
not since it can be gauged out. According to the gauge theory, a local phase degree
of freedom of the wave function is equivalent to gauge fields (electric and magnetic
fields) acting on the particles represented by the wave function—in our case, photons.
At this point, the phase φ represents a global phase of the wave function and, thus, has
no impact on the dynamics.
However, this will become completely different upon introducing more harmonic
traps to the system, with a phase difference between one another. Having a phase
φ(Er ) that depends on the location in space (of the harmonic trap) introduces a local
gauge degree of freedom, which translates to gauge fields according the gauge theory.
Indeed, by choosing different φ(E r ), it is possible to induce effective EM fields in the
hybrid modal-spatial space. To see this in a simple form in our derivation, we couple
such resonators with each other via proximity, in a spatial lattice with n dimensions.
The coupling is resonant (hence, efficient) when phase matching is satisfied; that is,
each mode is coupled to the same mode in its neighboring sites.
The idea is then to have an n + 1 dimensional lattice, where the extra dimension is
the ladder of modes associated with each site in the spatial lattice. Formally, this is
achieved by a basis transformation of the Hamiltonian of the n-dimensional lattice
of harmonic oscillators Hn D,tot with the unitary transformation, V = ⊗s U , which
Figure 5
is the tensor product of U with itself over the spatial dimension s . The resulting
Hamiltonian is H(n+1)D,eff = V † Hn D,tot V .
This recipe, which can be applied to any kind of arrangement of sites with inner
modes, results in the following Hamiltonian for an n + 1 dimensional lattice:
X
H(n+1)D,eff = 1s λ|λ, s ihλ, s | + J λ (|λ − 1, s ihλ, s |e iφs + h.c)
λ,s
where J s ,λ is the spatial coupling between mode λ in sites s and s − 1, and is depen-
dent on the geometry of the lattice, and φs , 1s are the phase term φ and frequency bias
term 1, which are dependent on the modulation phase and frequency at each site s .
Importantly φs , 1s form effective gauge fields (magnetic and electric) in the synthetic
dimension, which can be controlled by adjusting a different modulation in each site.
For example, a 1D system s = {1, . . . m}, with φs = s π2 φ and 1s = 0, corresponds to
a 2D lattice with a perpendicular magnetic field with flux π/2 in each unit cell, and 0
electric field since 1s = 0 [Fig. 5(b)]. With these parameters, Eq. (14) becomes the
Hamiltonian of the IQH, which is a 2D model that is now realized on a 1D spatial
lattice. This idea was demonstrated experimentally in two different platforms in pho-
tonics (spatial modes of an array of waveguides and frequency modes of a temporally
modulated ring resonator [85,88]), in microwaves [86] and also, before that, in cold
atoms with spin modes [76,77] and momentum modes [79].
We will now explain how this was realized in photonics and what the implications are
of such realizations on physical photonic systems. We go back to the physical system
of light propagating in a paraxial waveguide array (as discussed in Section 3.1). In
order to have synthetic dimensions in modal space, the first step is to have a confined
structure with a ladder of modes [examples in Figs. 5(c) and 5(d)]. This confined
structure plays the role of the harmonic trap as in the example above. In direct anal-
ogy to the harmonic trap example, it is possible to create a harmonic potential by
changing the index refraction 1ε(E r ) in a dielectric material. Conveniently, instead of
encoding a continuous harmonic potential in the permeability 1ε(E r ), it is possible to
have a potential made of discrete (single-mode) waveguides at varying distances, so
as to obtain the equally spaced modal structure of a harmonic potential. This lattice
is called a J x lattice, and in this lattice the distances between adjacent waveguides
increase as a function of their distance from the lattice center [Fig. 6(a)] [160]. As
a harmonic potential, the J x lattice has equally spaced modes that function as the
synthetic space. In order to couple the modes, an external perturbation is needed
[Eq. (12)]. The perturbation itself can take many forms. One option is to oscillate the
waveguide spatially at exactly the longitudinal spatial frequency that separates the
different modes [Fig. 6(b)]. Thus, this J x structure of a waveguide lattice functions
as a harmonic trap, and it can be placed near other identical J x lattices and obtain a
Hamiltonian as in Eq. (14). For example, placing the J x lattices near each other in
a row such that they are evanescently coupled [Fig. 6(c)] corresponds to a 2D syn-
thetic space with two dimensions: one spatial dimension (the location of the J x lattice
along x axis) and one synthetic dimension (the mode of each J x lattice). In the 1D
row of different J x lattices, each J x lattice oscillates at a different phase according to
φs = s π2 φ, where s counts the J x lattice along the x axis. A phase shift in φs that breaks
the z-reversal symmetry in the paraxial equation is equivalent to an effective magnetic
field [see Eq. (14)] and induces the formation of edge states in the synthetic space
lattice. Importantly, just as for the Harmonic traps, it is possible also to place the J x
photonic topological insulators in a geometry of a 2D lattice (for example, a diamond
lattice) and obtain lattices in three dimensions [85,161] with a 2D lattice geometry.
444 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
To understand the dynamics of the propagation of light on the edge of a space with
one spatial dimension and one synthetic dimension, we describe a wave packet that
propagates along the edge state in synthetic dimension of the waveguide structure
described.
In practice, the direct meaning of this in photonics is that we can design pho-
tonic systems with a topologically protected transport that is not on the spatial
edge of the lattice. Rather, the protected transport can be in the bulk of the lattice
instead of the edge [85], in the frequency dimension [84,88], or in both dimensions
simultaneously—for example, in a spiral motion [89]. To understand what it means,
let us analyze the propagation of light in an edge state in the modal dimension, as
illustrated by Fig. 6(c).
A single edge state in the synthetic space of Fig. 6 encircles the perimeter of the 2D
synthetic structure shown in Fig. 6(c). Thus, a wave packet propagating along the
edge in this synthetic space comprises a superposition of edge states and takes the
following course. In Stage A in Fig. 6(c), the wave packet propagates to the left in
the lowest mode of each J x lattice; a snapshot of the propagating wave packet in real
space is given in Fig. 6(d). Note that the lowest mode is the edge in synthetic space,
but not the edge in real space, as seen in Fig. 6(d). In fact, at this point the synthetic
space edge state is extended on the entire real-space lattice. This, however, does not
change the fact that its transport is topologically protected [162], since the edge state
is still a state crossing the gap exactly as an edge state in the IQH [see Fig. 1(c)]. When
this state reaches the leftmost J x lattice of the array, it does not scatter backward but
Figure 6
(a) J x waveguide array with equally spaced spatial modes. The modes are separated
by a longitudinal spatial frequency . (b) The same J x lattice as in (a), but here it
oscillates in z, at a frequency equal to . The oscillation couples the modes, thereby
creating a 1D lattice in synthetic space. A wavepacket initially populating the lowest
mode is efficiently transferred to the second mode by the modulation, then to the third
mode, etc. (c) A row of J x lattices (columns) and above it the 2D picture in synthetic
dimensions, where the horizontal axis is the location of every J x lattice (column) and
the vertical axis is the mode in each J x lattice. The arrows indicate the directionality
of the topological edge state in the synthetic dimension. (d) and (e) The real-space
intensity of the light when propagating in the structure: (d) is on the bottom edge [see
A in (c)], so the wavepacket populates a single lower mode (extended) on many J x
lattices. (e) The wavepacket at instant B in the propagation—on the leftmost edge.
In this case, the light is in a coherent superposition of modes on the leftmost J x lat-
tice (concentrated due to interference between the modes). The frame of the lattice
has a wave shape since the columns oscillate at a different phase along the vertical
real-space dimension (adapted from [85]).
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 445
rather concentrates entirely on the leftmost J x and starts ascending in the modes unidi-
rectionally (Stage B). At this point, the wave packet is in a superposition of modes that
appears as a concentrated wave packet [Fig. 6(e)]. After ascending to the top mode,
the light propagates to the right, while remaining strictly in the top mode at Stage C.
Finally, the light descends to the lowest mode on the rightmost J x lattice. In doing
that, the wave packet has completed a full cycle of transport around the synthetic
space edge.
As mentioned earlier, robust topological propagation in non-spatial dimensions is
by itself interesting even before considering the dimensionality increase. In these
systems, the light performs a robust unidirectional ascent (or descent) on the modal
ladder, which can potentially be used in future applications. One example of a poten-
tial application of synthetic space systems relying on modal ladders is the recent
proposal for a mode-locked topological insulator laser [93]. In short, a topological
insulator laser is an array of emitters (typically semiconductor lasers) judiciously
positioned on a topological photonic platform, so as to force locking of all the emit-
ters associated with the topological mode, making them act as a single high-power
coherent laser source [41,93,139,140,163–167].
The topological photonic platform can be any of the 2D platforms described above,
where each site is replaced by a laser emitter (e.g., a microring laser), with pumping
provided preferentially to the edge mode. What makes the emitters lock together
is two features of topological edge modes: (1) the non-zero velocity, implying that
the light in the edge mode must travel between all emitters associated with the edge
mode and (2) the topological robustness, which facilitates coupling between emitters
despite some inevitable level of imperfections or disorder. Thus far, however, all the
settings in which topological insulator lasers were demonstrated were real-space topo-
logical platforms, and they were geared toward coherent locking of many emitters to
a single frequency. But a recent proposal has targeted a different goal: locking many
emitters in a train of equally spaced frequencies for the purpose of generating mode-
locked laser pulses from a large array of semiconductor lasers [93]. The necessity of
locking a train of frequencies naturally suggested the use of synthetic dimensions,
with the frequency eigenmodes of the laser resonators acting as the modal dimension.
Let us describe that system in some detail. Considering a row of coupled ring res-
onators [83,84], we can treat each ring resonator in the same way we did with the
harmonic trap. The ring resonator also has a ladder of equally spaced resonance
frequencies that can be coupled by electro-optic phase modulation [Fig. 7(a)]. This
turns each ring resonator into a 1D lattice of modes. Then, as before, a row of ring
resonators represents a 2D lattice with one spatial dimension—the location of each
ring—and one synthetic dimension—the frequency mode of each ring. This princi-
ple was recently demonstrated on a single ring with two ladders (each frequency is
actually degenerate to two modes—clockwise and anticlockwise), and it was shown
experimentally that this structure can have synthetic magnetic fields [88] and even
exhibit non-Hermitian physics [168]. Each ring can be made to lase by adding satu-
rable gain, and together with the electro-optical modulation a ring laser can be made
to mode-lock and emit a train of mode-locked pulses. However, if we would like
several such rings to mode-lock together, it is virtually impossible to synchronize all
the ring resonators to lock on the same train of equally spaced frequencies. This is
simply because coupling just two rings causes the splitting of each mode, resulting
in two sets of equally spaced frequencies. In this vein, having N rings creates N sets
of frequencies, and it becomes virtually impossible to get all rings to synchronize on
the same set of frequencies, as required for coherent emission of mode-locked pulses.
In fact, the largest number of semiconductor emitters that was ever made to mutually
446 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
Figure 7
(a) Edge state in the synthetic modal space of a row of phase modulated ring res-
onators. (b) A chain of modulated ring resonators with selective saturable gain that
forms an edge state. This causes a topological edge state to lase given robust proper-
ties to the lasing (adapted from [93]). (c) Temporally modulated ring resonator with
χ 3 nonlinearity and spatially modulated GVD can support Hamiltonians with local
interactions in synthetic space (adapted from [170]).
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 447
U n = e −i Heff nT , (15)
where Heff is a static effective Hamiltonian. Notice that the full temporal dynamics
of Heff coincide with the dynamics ψnT at the beginning of each period. The Floquet
Hamiltonian is known to exhibit topological phases in a variety of lattice systems
[25,29,177], but here we will not assume any underlying spatial lattice whatsoever.
Since analyzing the dynamics of this system requires sampling it every period T,
the remaining time on the time axis remains unused. The idea then is to utilize the
“unused time” as a spatial degree of freedom. If each time-bin represents a different
location in space, adding a delay 1t (−1t) to the information expected to arrive at
time nT means that the information has moved one step to the right (or left).
On the basis of the time-bin lattice with an additional degree of freedom of the
Hamiltonian (such as polarization), it is possible to construct a system with a topologi-
cal band structure, where the Brillouin zone is the reciprocal space to the time-bins,
and the additional degree of freedom allows for the complexity required for the
non-trivial topology. To understand this, consider a typical system composed of two
coupled fiber loops with different length x ± 1x , as shown in Fig. 8(a). A light pulse
is launched into the system and completes encircling each fiber in times T ± 1T,
depending on which fiber it travels in. After completing n periods, the pulse can be
in one of the following times: t = nT + p1T, p = {−n, −n + 1, . . . , n}. Thus, each
value of p is a time-bin, and it is a synthetic dimension that represents a location
in space. Each time 1T is regarded as one step in the motion in which the photon
can move left or right in this synthetic space [Fig. 8(b)]. In principle, additional
synthetic space dimensions can be obtained by using several time scales. That is,
adding another bin splitting to obtain two different paths with temporal difference
1T̃, which is larger than 2n1T, makes it possible to have another synthetic spatial
dimension. The pulses in the system have a spatial dimensionality of the number of
time scales added, but in order to have a sense of a geometric phase involved in topo-
logical phases, another degree of freedom is added to the spatial one. In the system
described in Fig. 8, it is the polarization of the pulse. The polarization is described
by a 2D vector and can accumulate a phase and, thus, a geometric phase. In this set-
ting, topological edge states can be obtained by coupling the polarization degree of
freedom to the translation degree of freedom. This operation can be done simply by
using a polarization-dependent beam splitter, which can be described by the following
operator:
X
Tx = |x + 1ihx |⊗| ↑ih↑ |+| x − 1x ih| ⊗ |↓ih↓| , (16)
x
where |x i is the synthetic position degree of freedom and | ↑i (| ↓i) is the spin up
(down). The polarization here is analogous to the geometric phase of the wave func-
tion, which is responsible for the topological effects described in previous sections.
448 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
To obtain a topological phase in 1D, we split Tx to two operators, T↑ (T↓ ), which move
the light one site right (left) if the spin is up (down). Then, by adding two rotations
between the two splitted parts of Tx we obtain
where R(θ) is polarization rotation by angle θ . The DTQW related to this operator
[Fig. 8(c)] has a topological phase diagram with two possible topological phases
[Fig. 8(d)]. The topological phases are characterized by the Zak phase from Eq. (1)
of the effective Hamiltonian [178]. The implication of a boundary in the DTQW
between different regions with different θ2 [Fig. 8(e)], and with different topologi-
cal phases, can be seen in the left panel [Fig. 8(f)]. Exciting the edge state causes
localization, where the wave remains concentrated on the edge without coupling to
neighboring sites. This is contrary to the right panel in which the two regions have the
same topological phase (but different θ2 ), and in this case no localization occurs. As
for the system described in previous sections, the studies on these synthetic dimen-
sions can be extended both in dimensionality and in fundamental properties such as
hermiticity and nonlinearity. Moreover, by considering an additional time scale—it
is possible to add the same spin dependence also in y : Ty and to consider the cyclic
process:
Figure 8
(a) Illustration of two coupled fiber loops of different lengths. Different output arrival
times correspond to different locations on a 1D lattice in space. (b) Steps of the
discrete time quantum walks (DTQWs) corresponding to the two-fiber setting in
(a) (adapted from [179]). (c) DTQW coupling of the polarization degree of freedom to
the spatial (time-bins) degree of freedom in a DTQW process [113]. (d) Topological
phase diagram corresponding to different rotation angles θ1 and θ2 of the process in
(c). Here, gray regions correspond to non-trivial topology with Rudner number 1
[113]. (e) A boundary with edge states forming by changing θ2 [113]. (f) Edge propa-
gation of two regions in the DTQW. In the left panel, the two regions are in a different
topological phase inducing a localized state on the edge. In the right panel, the two
regions have distinct θ2 but same topological phase, so no localization occurs [113].
Review Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics 449
The process described by this equation also has topological phases, and, since it is
a 2D model, the topological properties are different from those of the 1D model. In
this Floquet system, the topological invariant is the Rudner number [180]. Just like
the Chern number, the Rudner number predicts the existence of unidirectional robust
edge states in the gap. As in the 1D case, one can have a phase diagram as of a function
of θ1 and θ2 [Fig. 9(a)]. The edge state at the boundary between different values of θ1
is illustrated in Fig. 9(b), and the band diagram. The reason that the Chern number is
not used here is because the Chern number gives for each band the difference between
the number of edge states bellow the band and above the band. Since in this type of
Floquet Hamiltonians the difference is zero (but edge states still exist) [see Figs. 9(c)
and 9(d)], the Rudner number, which takes into account the temporal dependence of
the Hamiltonian, is the proper quantity to use.
The propagation of light in this system is simulated in Fig. 9(e). The light bypasses
two corners along the edge, demonstrating the topological robustness of the
propagation [125].
Figure 9
Figure 10
(a) Non-Hermitian fiber loop setup. (b) PT-symmetric fiber loop setup. The purple
color on the fiber loops indicates that the gain and loss are alternating on every other
round trip [111].
450 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
FUNDING
Air Force Office of Scientific Research; Israel Science Foundation.
ACKNOWLEDGMENT
Eran Lustig is supported by the Adams Fellowship Program of the Israel Academy of
Sciences and Humanities.
DISCLOSURES
The authors declare no conflicts of interest.
80. H. Cai, J. Liu, J. Wu, Y. He, S.-Y. Zhu, J.-X. Zhang, and D.-W. Wang,
“Experimental observation of momentum-space chiral edge currents in
room-temperature atoms,” Phys. Rev. Lett. 122, 023601 (2019).
81. H. M. Price, T. Ozawa, and N. Goldman, “Synthetic dimensions for cold atoms
from shaking a harmonic trap,” Phys. Rev. A 95, 023607 (2017).
82. X.-W. Luo, X. Zhou, C.-F. Li, J.-S. Xu, G.-C. Guo, and Z.-W. Zhou, “Quantum
simulation of 2D topological physics in a 1D array of optical cavities,” Nat.
Commun. 6, 7704 (2015).
83. T. Ozawa, H. M. Price, N. Goldman, O. Zilberberg, and I. Carusotto, “Synthetic
dimensions in integrated photonics: from optical isolation to four-dimensional
quantum Hall physics,” Phys. Rev. A 93, 043827 (2016).
84. L. Yuan, Y. Shi, and S. Fan, “Photonic gauge potential in a system with a
synthetic frequency dimension,” Opt. Lett. 41, 741–744 (2016).
85. E. Lustig, S. Weimann, Y. Plotnik, Y. Lumer, M. A. Bandres, A. Szameit, and
M. Segev, “Photonic topological insulator in synthetic dimensions,” Nature 567,
356–360 (2019).
86. C. W. Peterson, W. A. Benalcazar, M. Lin, T. L. Hughes, and G. Bahl, “Strong
nonreciprocity in modulated resonator chains through synthetic electric and
magnetic fields,” Phys. Rev. Lett. 123, 063901 (2019).
87. A. Dutt, M. Minkov, Q. Lin, L. Yuan, D. A. B. Miller, and S. Fan, “Experimental
band structure spectroscopy along a synthetic dimension,” Nat. Commun. 10,
3122 (2019).
88. A. Dutt, Q. Lin, L. Yuan, M. Minkov, M. Xiao, and S. Fan, “A single photonic
cavity with two independent physical synthetic dimensions,” Science 367, 59–64
(2020).
89. Q. Lin, X.-Q. Sun, M. Xiao, S.-C. Zhang, and S. Fan, “A three-dimensional pho-
tonic topological insulator using a two-dimensional ring resonator lattice with a
synthetic frequency dimension,” Sci. Adv. 4, eaat2774 (2018).
90. Q. Lin, M. Xiao, L. Yuan, and S. Fan, “Photonic Weyl point in a two-
dimensional resonator lattice with a synthetic frequency dimension,” Nat.
Commun. 7, 13731 (2016).
91. A. Dutt, M. Minkov, I. A. D. Williamson, and S. Fan, “Higher-order topological
insulators in synthetic dimensions,” Light Sci. Appl. 9, 131 (2020).
92. L. Yuan, Y. Shi, and S. Fan, “Achieving the gauge potential for the photon in a
synthetic space,” in Conference on Lasers and Electro-Optics (Optical Society
of America, 2016), paper SF2E.5.
93. Z. Yang, E. Lustig, G. Harari, Y. Plotnik, Y. Lumer, M. A. Bandres, and M.
Segev, “Mode-locked topological insulator laser utilizing synthetic dimensions,”
Phys. Rev. X 10, 011059 (2020).
94. B. A. Bell, K. Wang, A. S. Solntsev, D. N. Neshev, A. A. Sukhorukov, and B. J.
Eggleton, “Spectral photonic lattices with complex long-range coupling,” Optica
4, 1433–1436 (2017).
95. L. Yuan, Q. Lin, M. Xiao, and S. Fan, “Synthetic dimension in photonics,”
Optica 5, 1396–1405 (2018).
96. Y. Wang, H. M. Price, B. Zhang, and Y. D. Chong, “Circuit implementation of a
four-dimensional topological insulator,” Nat. Commun. 11, 2356 (2020).
97. L. Nemirovsky, M. Cohen, E. Lustig, and M. Segev, “Magnetic gauge field for
photons in synthetic dimensions by a propagation-invariant photonic structure,”
in Conference on Lasers and Electro-Optics, OSA Technical Digest (Optical
Society of America, 2009), paper FW3D.7.
98. L. Nemirovsky, M. Cohen, Y. Lumer, E. Lustig, and M. Segev, “Topological
evolution-invariant photonic structures in synthetic dimensions,” in Conference
456 Vol. 13, No. 2 / June 2021 / Advances in Optics and Photonics Review
Science (NAS) of the United States of America. In 2014, Moti Segev won the Israel
Prize in Physics and Chemistry (highest honor in Israel), and in 2019 he won the
EMET Prize (Israel). However, above all his personal achievements, he takes pride
in the success of his graduate students and postdocs; among them are currently 23
professors in the USA, Germany, Taiwan, Croatia, Italy, India, China, and Israel, and
many hold senior R&D positions in the industry.