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

Artificial magnetic field for synthetic quantum matter without dynamical modulation

Tomoki Ozawa Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan RIKEN Interdisciplinary Theoretical and Mathematical Sciences Program, Wako 351-0198, Japan
Abstract

We propose an all-static method to realize an artificial magnetic field for charge neutral particles without introducing any time modulation. Our proposal consists of one-dimensional tubes subject to harmonic trapping potentials with shifted centers. We show that this setup realizes an artificial magnetic field in a hybrid real-momentum space. We discuss how characteristic features of particles in a magnetic field, such as chiral edge states and the quantized Hall response, can be observed in this setup. We find that the mean-field ground state of bosons in this setup in the presence of long-range interactions in physical real space can have quantized vortices in the hybrid real-momentum space; such a state with vortices exhibits a supersolid structure in the physical real space. Our method can be applied to a variety of synthetic quantum matter, including ultracold atomic gases, coupled photonic cavities, coupled waveguides, and exciton-polariton lattices.

Introduction Quantum simulation has been a central theme in the research of synthetic quantum matter based on atomic, molecular, and optical systems, such as ultracold atomic gases, trapped ions, polaritons, and photonics Bloch:2012 ; Blatt:2012 ; Guzik:2012 ; Carusotto:2013 ; Georgescu:2014 ; Bloch:2018 ; Boulier:2020 . In particular, simulation of topological phenomena using synthetic quantum matter has attracted a considerable attention in the recent years Hasan:2010 ; Qi:2011 ; Chiu:2018 ; Armitage:2018 ; Cooper:2019 ; Ozawa:2019 . A prototypical example of topological phenomena is the quantum Hall effect, which was first found in a two-dimensional electron gas under a magnetic field Klitzing:1980 ; Tsui:1982 ; Yoshioka:Book . To simulate the quantum Hall effect and other topological phenomena, one often needs to simulate physics of charged particles in a magnetic field, which is not straightforward because atoms in ultracold gases and photons in cavities and waveguides are charge-neutral.

Various methods have been employed to simulate effects of a magnetic field, namely to realize an artificial magnetic field Aidelsburger:2018 . In ultracold atomic gases, the earliest example is to rotate the system, exploiting the similarity between the Coriolis force and the Lorentz force Madison:2000 ; Cooper:2008 ; Fetter:2009 . Other examples include Floquet engineering GoldmanDalibard:2014 , light-induced gauge potential Goldman:2014 , and synthetic dimensions Ozawa:2019SD . Most of the existing proposals involve adding fast time modulation and/or exploiting different energy and time scales present in the system to explore adiabatic physics in a low energy subspace. These methods often suffer from instability and heating as a result of time modulation and spontaneous emissions. A notable exception is the method using spinor condensates in a quadrupole magnetic field where spatially dependent spin texture results in an artificial magnetic field Ho:1996 ; Pietila:2009 ; Choi:2013 ; Ray:2014 . The situation is similar in photonics and optics. For light close to optical frequencies, magneto-optical effects are generally weak and thus, similar to ultracold atomics gases, one needs to realize an artificial magnetic field to explore the physics of quantum Hall effects Rechtsman:2013 ; Yuan:2018 ; Lustig:2019 . Static realizations of topological models in photons are in the class of quantum spin-Hall insulators where two degrees of freedom feel opposite magnetic fields and thus topological protection relies upon decoupling of different degrees of freedom Hafezi:2013 ; Schine:2016 .

Refer to caption
Figure 1: Illustration of the scheme. (a) One-dimensional tubes elongated in the xx direction are aligned along the yy direction. Harmonic trapping potentials, depicted by parabolas, are applied to the tubes. The trapping centers are on the dashed line, which is tilted with respect to the xx axis by an angle θ\theta. (b) Setup viewed from top. Tubes are located at y=ny0y=ny_{0}, where nn is an integer. The trapping centers are on the dashed line y=xtanθy=x\tan\theta. (c) In the xx-pyp_{y} plane, an artificial magnetic field of strength B=tanθB=-\tan\theta is realized. We have the geometry of a cylinder, where the pyp_{y} direction is periodic with periodicity 2π/y02\pi/y_{0}.

In this paper, we propose a completely static way to realize an artificial magnetic field regardless of the spin of the particles, which can be applied to a variety of synthetic quantum matter. Our proposal is based on a well-known fact that the Hamiltonian of a charged quantum particle in two dimensions under a uniform magnetic field is equivalent to a set of harmonic oscillators with shifted centers LL:QM . This fact implies, conversely, that a set of harmonic oscillators with shifted centers can be viewed as a charged particle in a uniform magnetic field. We pursue this analogy further and consider a set of one-dimensional systems elongated along the xx direction and align these one-dimensional tubes in the yy-direction and place them under harmonic trapping potentials with shifted centers (Fig. 1). We find that the Hamiltonian of a charged particle in a magnetic field is realized in the xx-pyp_{y} plane where pyp_{y} is the momentum along the yy direction. Even though the proposal looks deceptively simple, as we discuss in detail, we can still observe most of the phenomena characteristic of charged particles in a magnetic field, such as the chiral edge states and quantized Hall response. We note that there is no dynamical component in the proposal, and, in particular, our proposal does not break the time-reversal symmetry in the physical xx-yy plane. Our proposal is thus expected to be more stable and free from heating compared with existing proposals which rely on dynamical modulations.

We also explore the effect of interparticle interactions in the artificial magnetic field. Weakly interacting bosons under a magnetic field form vortices Madison:2000 ; Cooper:2008 ; Fetter:2009 ; Spielman:2009 ; Lin:2009 . However, a contact interaction, common in ultracold atomic gases and photons with optical nonlinearity, translates into extremely long-range interaction in the pyp_{y} direction, and we find that vortices do not form under ordinary contact interactions. On the other hand, by including nearest-neighbor interaction in yy direction we find that vortices can form in the xx-pyp_{y} plane. Such a vortex structure corresponds to supersolid structure in the physical xx-yy plane, where the phase coherence of a condensate is maintained and the discrete translational symmetry is broken. Finally, we discuss possible experimental platforms where the proposal can be implemented.

I Setup

We consider a set of uncoupled one-dimensional tubes elongated in the xx direction and aligned in the yy direction, where each tube is subject to a harmonic trapping potential. The center of the trapping potential shifts between tubes by a constant amount x0x_{0}. We assume that tubes are separated by a constant amount y0y_{0} in the yy direction. See Fig. 1 for an illustration. We use the second-quantized formalism, where ψ^n(x)\hat{\psi}^{\dagger}_{n}(x) and ψ^n(x)\hat{\psi}_{n}(x) are the creation and annihilation operators, respectively, of a particle at coordinate (x,ny0)(x,ny_{0}), where nn indicates that the particle is in the nnth tube. The single-particle Hamiltonian is

H^=\displaystyle\hat{H}= ndx[12m(xψ^n(x))(xψ^n(x))\displaystyle\sum_{n}\int dx\left[\frac{1}{2m}\left(\partial_{x}\hat{\psi}_{n}^{\dagger}(x)\right)\left(\partial_{x}\hat{\psi}_{n}(x)\right)\right.
+12mω2(xnx0)2ψ^n(x)ψ^n(x)],\displaystyle+\left.\frac{1}{2}m\omega^{2}(x-nx_{0})^{2}\hat{\psi}_{n}^{\dagger}(x)\hat{\psi}_{n}(x)\right], (1)

Note that there is no motion (coupling) between different tubes in the yy direction. By performing Fourier transformation in the yy direction,

ψ^n(x)=y02ππ/y0π/y0𝑑pyeiny0pyψ^(x,py),\displaystyle\hat{\psi}_{n}(x)=\sqrt{\frac{y_{0}}{2\pi}}\int_{-\pi/y_{0}}^{\pi/y_{0}}dp_{y}e^{iny_{0}p_{y}}\hat{\psi}(x,p_{y}), (2)

the Hamiltonian can be rewritten as

H^=\displaystyle\hat{H}= dxπ/y0π/y0dpy[12m(xψ^(x,py))(xψ^(x,py))\displaystyle\int dx\int_{-\pi/y_{0}}^{\pi/y_{0}}dp_{y}\left[\frac{1}{2m}\left(\partial_{x}\hat{\psi}^{\dagger}(x,p_{y})\right)\left(\partial_{x}\hat{\psi}(x,p_{y})\right)\right.
+12mω2tan2θ{(ipy+xtanθ)ψ^(x,py)}\displaystyle+\left.\frac{1}{2}\frac{m\omega^{2}}{\tan^{2}\theta}\left\{(i\partial_{p_{y}}+x\tan\theta)\hat{\psi}^{\dagger}(x,p_{y})\right\}\right.
{(ipy+xtanθ)ψ^(x,py)}].\displaystyle\hskip 56.9055pt\left.\left\{(-i\partial_{p_{y}}+x\tan\theta)\hat{\psi}(x,p_{y})\right\}\right]. (3)

In the first-quantized form, this Hamiltonian is

H=12mx2+mω22tan2θ(ipy+xtanθ)2\displaystyle H=-\frac{1}{2m}\partial_{x}^{2}+\frac{m\omega^{2}}{2\tan^{2}\theta}\left(-i\partial_{p_{y}}+x\tan\theta\right)^{2} (4)

This is nothing but the Hamiltonian of a charged particle in the xx-pyp_{y} plane under a magnetic vector potential 𝐀(x,py)=(0,xtanθ)\mathbf{A}(x,p_{y})=(0,-x\tan\theta) in the Landau gauge, corresponding to the artificial magnetic field B=tanθB=-\tan\theta penetrating the plane. (We have set the charge ee, speed of light cc, and Plank constant \hbar to be unity.) We have chosen to use ipy=y-i\partial_{p_{y}}=-y as the “momentum” operator along the pyp_{y} direction, regarding pyp_{y} as the synthetic “position.” The mass is anisotropic; the mass in the xx direction is the original mass mx=mm_{x}=m, whereas the mass in the pyp_{y} direction is mpy=tan2θ/mω2m_{p_{y}}=\tan^{2}\theta/m\omega^{2}. This anisotropy can be absorbed by rescaling units of length in two directions and does not pose any problem in the rest of the discussion. Another noticeable feature here is that the xx-pyp_{y} plane is a cylinder, open along the xx direction and periodic along the pyp_{y} direction.

Let us comment on the time-reversal symmetry in our system. Our original Hamiltonian (1) does not break the physical time-reversal symmetry. For a single component (spinless) system, the time-reversal operation flips the sign of the momentum 𝐩𝐩\mathbf{p}\to-\mathbf{p}. Since we want to look at the physics related to the artificial magnetic field in the xx-pyp_{y} plane, the relevant effective time-reversal symmetry is now to flip their reciprocal variables 𝒯:(px,y)(px,y)\mathcal{T}:(p_{x},-y)\to(-p_{x},y); the effective time-reversal symmetry is the simultaneous operation of flipping of pxp_{x} and the parity in yy. Our Hamiltonian does break the parity in the yy direction due to the shifted harmonic trapping potentials, which enables us to realize the artificial magnetic field in the xx-pyp_{y} plane without breaking the physical time-reversal symmetry.

Our method is related to the idea of the topological charge pumping, where a momentum is replaced by an external parameter; by sweeping the parameter one obtains topological properties of the original Hamiltonian Thouless:1983 ; Nakajima:2016 ; Lohse:2016 . Our method, however, is crucially different from charge pumping in that we replace a momentum pyp_{y} by a physical real coordinate y-y, and in this way we succeed in recovering the full dynamics in all dimensions, whereas no dynamics occurs in charge pumping in the direction replaced by an external parameter.

Refer to caption
Figure 2: Energy level and the Hall response. (a) The energy level of each tube at position nn, where length of each tube is 20aosc20a_{\mathrm{osc}} and we employ an open boundary condition. We choose tanθ=1\tan\theta=1, and the energy is in units of the cyclotron frequency ω\omega. Solid lines are to guide the eye. (b) The center-of-mass motion of wave packets when a synthetic electric field Ex=0.1ω/aoscE_{x}=0.1\omega/a_{\mathrm{osc}} is applied. Three solid lines show the time evolution of the center of mass in the pyp_{y} direction for tanθ=2/3\tan\theta=2/3 (top, green), tanθ=1\tan\theta=1 (middle, blue), and tanθ=1/2\tan\theta=1/2 (bottom, red). The dashed lines going through the solid lines are theoretical predictions from Eq. (5) with 𝒞=1\mathcal{C}=1. The units of time are 1/ω1/\omega, and the units of pyp_{y} are 1/y01/y_{0}.

II Chiral edge states

We now analyze how characteristic features of charged particles in a magnetic field can be observed in our setup in the xx-pyp_{y} plane. We first point out that the cyclotron frequency of the artificial magnetic field is ω\omega, as expected. This can be derived from the definition of the cyclotron frequency B/m~B/\tilde{m}; as the mass m~\tilde{m}, we need to use a geometric mean of anisotropic effective mass in two directions m~=mxmpy\tilde{m}=\sqrt{m_{x}m_{p_{y}}}. Using B=tanθB=-\tan\theta, we then obtain ω=|B|/m~\omega=|B|/\tilde{m}. This observation implies the well-known fact that the energy level of harmonic oscillators in the xx-yy plane, which is ω(l+1/2)\omega(l+1/2), where l=0,1,2,l=0,1,2,\cdots is a nonnegative integer, is nothing but the Landau-level energy spectrum. High degeneracy of Landau levels come from the many uncoupled tubes present in the system, which all have the same energy spectrum.

In the presence of edges, charged particles in two dimensions under a magnetic field should have chiral edge states. In Fig. 2(a), we plot the energy level of the system in the presence of sharp edges in the xx direction as a function of nn, which is the reciprocal variable of pyp_{y}. We consider 21 tubes and the length of each tube in the xx direction is 20aosc20a_{\mathrm{osc}}, where aosc1/mωa_{\mathrm{osc}}\equiv 1/\sqrt{m\omega} is the oscillator length. We can see that at large values of |n||n|, the dispersion goes upward; these states are localized at the edges and they propagate in one direction. The existence of chiral edge states can be probed from the dynamics of wave packets localized at the edges as we show below. Since there is no coupling between different tubes, the population of particles in each tube does not change; in real space, the propagation of chiral edge states in the pyp_{y} direction appears as an appropriate evolution of the relative phases in each tube, which can be probed, for example, in ultracold atomic gases through time-of-flight imaging.

We now numerically demonstrate how the chiral edge mode in the xx-pyp_{y} plane can be observed through wave packet dynamics. We choose the strength of the artificial magnetic field to be B=tanθ=1B=-\tan\theta=-1 for the numerical simulation. As an initial state, we consider three wave packets, located at the left edge, center, and the right edge of the system. To consider physics relevant to a single band or band gap, we project the wave packet onto the lowest level of each tube. Experimentally, in ultracold gases, such an initial state can be prepared by creating a wave packet out of single-particle ground states of each tube by starting from Bose-Einstein condensates in each tube. In photonics, where light can be injected from the outside, we can insert light with frequency corresponding to the lowest Landau level to prepare a similar initial condition. The wave function of the initial state is plotted in Fig. 3(a) and Fig. 3(b), in both the xx-yy and xx-pyp_{y} planes. We then let the wave packet evolve freely. The wave functions in the xx-pyp_{y} plane after time t=1/ωt=1/\omega, t=2/ωt=2/\omega, and t=3/ωt=3/\omega are plotted in Figs. 3(c)-3(e). We observe that the wave packet at the left edge goes upward, whereas the wave packet at the right edge goes downward, showing that the edge state goes clockwise in a chiral manner. This motion of the wave packet at the edge is consistent with the artificial magnetic field in the xx-pyp_{y} plane; the strength of the magnetic field is B=tanθ<0B=-\tan\theta<0, which implies that the cyclotron motion inside the bulk of the system is counterclockwise, and the edge state goes clockwise. We note that the absolute value of the wave function plotted in the xx-yy plane looks identical during time evolution. This is because our initial state is projected onto the lowest Landau level, and thus the wave function in each tube is an eigenstate of the Hamiltonian for each tube, and thus the only time evolution appears in the phase of the wave function. The evolution of the phase in the xx-yy plane appears as chiral edge motion in the xx-pyp_{y} plane.

Refer to caption
Figure 3: Dynamics of a wave packet and chiral edge modes. The units of length in the xx direction are aosca_{\mathrm{osc}}, and the units of pyp_{y} are 1/y01/y_{0}. (a) The absolute value of the initial wave function is plotted in the xx-yy plane, where each nn represents a tube located at the position y=ny0y=ny_{0}. (b) The initial state in the xx-pyp_{y} plane. The lower panel shows the wave function in the xx-pyp_{y} plane after time evolution for duration (c) t=1/ωt=1/\omega, (d) t=2ωt=2\omega, and (e) t=3/ωt=3/\omega. Chiral edge motion is clearly visible.

We should also comment on the direction of the edge mode with respect to the energy spectrum, see Fig. 2(a). One sees that, in the energy spectrum, the energy goes up at large values of nn. These states at large values of n>0n>0 are localized at the right edge. Since the energy goes up, one might want to conclude that the group velocity is positive at the right edge. However, as one sees from Fig. 3, this is not the case; the group velocity is negative at the right edge. This apparent discrepancy is because one needs to consider n-n as the conjugate variable of pyp_{y}. Namely, the horizontal axis of Fig. 2(a) should be flipped when one wants to find the correct group velocity. By flipping the horizontal axis of Fig. 2(a), one obtains the correct group velocity, which is negative at the right edge.

III Quantized Hall response

Since Landau levels are topological, the Hall response should be quantized, which can be a signature of nontrivial topology obtained from the bulk property of the system. The topological Chern number 𝒞\mathcal{C} of Landau levels with B<0B<0 is one, 𝒞=1\mathcal{C}=1; we now investigate if this unit Chern number can be observed in our setup.

One way to see the topological nature of the system is through Laughlin charge pumping Laughlin:1981 ; on a two-dimensional cylinder with a perpendicular magnetic field, when a magnetic flux quantum is inserted through the cylinder, the state on the cylinder moves along the cylinder by one unit. As we naturally have a cylindrical structure in the xx-pyp_{y} plane, it is natural to ask how Laughlin charge pumping appears in our setup. As we see below, quantized pumping in our system occurs naturally due to construction. A magnetic flux can be inserted through the cylinder in the xx-pyp_{y} plane by shifting the center of the harmonic trapping potential in the xx-yy plane. If we shift the central position of the harmonic trapping potentials by δx\delta x, the harmonic trapping potential acts as 12mω2(xnx0δx)2\frac{1}{2}m\omega^{2}(x-nx_{0}-\delta x)^{2} for a given tube labeled by an integer nn. In the xx-pyp_{y} plane, this shift corresponds to the shift of the magnetic vector potential to 𝐀(x,py)=(0,xtanθ+δxtanθ)\mathbf{A}(x,p_{y})=(0,-x\tan\theta+\delta x\tan\theta). The additional contribution of δxtanθ\delta x\tan\theta is a constant term in the magnetic vector potential, and thus if we integrate it over the pyp_{y} direction, we see that the cylinder in the xx-pyp_{y} plane is penetrated by a magnetic flux 2πδxtanθ/y0=2πδx/x02\pi\delta x\tan\theta/y_{0}=2\pi\delta x/x_{0}. Therefore, inserting a magnetic flux quantum of 2π2\pi corresponds to taking δx=x0\delta x=x_{0}. This shift of δx=x0\delta x=x_{0} corresponds to no change in the overall Hamiltonian except that a tube previously labeled by an integer nn now acts as a tube with a label n+1n+1. Thus, as in the Laughlin pumping, the overall Hamiltonian comes back to the original one after the shift of δx=x0\delta x=x_{0}. In the xx-yy plane, the shift of δx=x0\delta x=x_{0} corresponds to moving the entire system by an amount δx=x0\delta x=x_{0} to the right, thus any state should also move horizontally by x0x_{0}. This implies that any state on the cylinder of the xx-pyp_{y} plane should also move along the cylinder by x0x_{0}, which is the manifestation of the Laughlin charge pumping in our setup. Another way to see the topological nature of the system is through dynamical response of the system. In ultracold gases, quantized Hall response can be observed through the center-of-mass response upon application of a force (synthetic electric field) 𝐄=(Ex,Epy)\mathbf{E}=(E_{x},E_{p_{y}}) Dauphin:2013 ; Aidelsburger:2015 ; Price:2016 . When a synthetic electric field EiE_{i} is applied along the ii direction, the center-of-mass velocity is Price:2016 viC.M.=ϵij2πABZEj𝒞v^{\mathrm{C.M.}}_{i}=-\epsilon_{ij}\frac{2\pi}{A_{\mathrm{BZ}}}E_{j}\mathcal{C}, where ABZA_{\mathrm{BZ}} is the area of the Brillouin zone and ϵxpy=ϵpyx=1\epsilon_{xp_{y}}=-\epsilon_{p_{y}x}=1. For our setup of a uniform magnetic field in the xx-pyp_{y} plane, there is no translational symmetry in the xx direction, and therefore we need to consider a magnetic unit cell and the magnetic Brillouin zone. A magnetic unit cell is an arbitrary rectangle in the xx-pyp_{y} plane which encloses one magnetic flux quantum. Since the strength of the artificial magnetic field is |B|=tanθ|B|=\tan\theta, the area of a magnetic unit cell is 2π/tanθ2\pi/\tan\theta. Then the area of the magnetic Brillouin zone is (2π)2(2\pi)^{2} divided by the area of a magnetic unit cell and thus ABZ=2πtanθA_{\mathrm{BZ}}=2\pi\tan\theta. Therefore, the center-of-mass velocity should follow

viC.M.=ϵijEj𝒞/tanθ=ϵijEj𝒞/B.\displaystyle v^{\mathrm{C.M.}}_{i}=-\epsilon_{ij}E_{j}\mathcal{C}/\tan\theta=\epsilon_{ij}E_{j}\mathcal{C}/B. (5)

We should then be able to determine the Chern number of the Landau levels in the xx-pyp_{y} plane by monitoring the center-of-mass motion. As we see below, the quantized response upon adding EpyE_{p_{y}} is nothing but the Laughlin charge pumping we saw above and turns out to be trivially satisfied, whereas the response upon ExE_{x} can provide observable signatures of the quantized Hall response.

We first consider the Hall response upon adding EpyE_{p_{y}}, which can be applied by making the artificial magnetic vector potential time dependent 𝐀=(0,xtanθEpyt)\mathbf{A}=(0,-x\tan\theta-E_{p_{y}}t). Such a time dependence can again be introduced by changing the position of the minima of the harmonic trapping potentials. Changing the location of the trap minima from nx0nx_{0} to nx0+δxnx_{0}+\delta x, in a yy-independent manner, the magnetic vector potential in the xx-pyp_{y} plane becomes 𝐀=(0,xtanθ+δxtanθ)\mathbf{A}=(0,-x\tan\theta+\delta x\tan\theta) as before. This implies that the desired magnetic vector potential can be created by moving δx\delta x as δx=Epyt/tanθ\delta x=-E_{p_{y}}t/\tan\theta. Now, the quantized Hall response through the center-of-mass velocity has a clear physical meaning. When the motion of the trapping potential is slow enough, the center of mass should follow the motion of the trap minima δx(t)\delta x(t). Therefore, the center-of-mass velocity in the xx direction is vxC.M.=tδx(t)=Epy/tanθv^{\mathrm{C.M.}}_{x}=\partial_{t}\delta x(t)=-E_{p_{y}}/\tan\theta; from Eq. (5), this implies the correct relation 𝒞=1\mathcal{C}=1.

Adding ExE_{x}, the center-of-mass velocity follows vpyC.M.=Ex𝒞/tanθv^{\mathrm{C.M.}}_{p_{y}}=E_{x}\mathcal{C}/\tan\theta, which can provide a more nontrivial observable signature of topology. A synthetic electric field ExE_{x} can be applied by adding a linear potential gradient Exx-E_{x}x to the Hamiltonian. We numerically simulate the time evolution of a wave packet located at the center, projected onto the lowest Landau level, under the field ExE_{x}. In Fig. 2(b), we plot the center-of-mass position of the wave packet in the pyp_{y} direction under time evolution for different values of B=tanθB=-\tan\theta. We also plot the theoretical prediction from Eq. (5) assuming 𝒞=1\mathcal{C}=1, which agrees well with the numerical simulation of py\langle p_{y}\rangle. The numerical simulation also exhibits oscillation around the theoretical line with oscillation period 2π/ω2\pi/\omega, which is a detailed structure of the cyclotron motion. Thus, the quantized Hall response can be observed from the center-of-mass motion in the pyp_{y} direction under the application of ExE_{x}, and the Chern number can be estimated from its slope.

IV Interaction and vortices

We now turn to interacting many-body cases. An important feature is that a typical short-range interaction in the xx-yy plane translates into long-ranged and unconventional interaction in the pyp_{y} direction. In particular, the formation of vortices, which is a characteristic feature of weakly interacting bosons in a magnetic field Madison:2000 ; Cooper:2008 ; Fetter:2009 ; Spielman:2009 ; Lin:2009 , cannot be observed in the xx-pyp_{y} plane only with a contact interaction in the xx-yy plane. Nonetheless, we find that, by adopting a long-range interaction in the yy direction, vortices can still form in our setup as we discuss below. We now focus on bosons with the mean-field interaction and ask if vortices can form in xx-pyp_{y} plane. We use description in terms of the Gross-Pitaevskii equation with the condensate wave function Ψn(x,t)\Psi_{n}(x,t):

itΨn(x,t)=Hyhop+Hint+\displaystyle i\partial_{t}\Psi_{n}(x,t)=H_{\mathrm{yhop}}+H_{\mathrm{int}}+
{x22m+12mω2(xnx0)2+12mω~2x2}Ψn(x,t),\displaystyle\left\{-\frac{\partial_{x}^{2}}{2m}+\frac{1}{2}m\omega^{2}(x-nx_{0})^{2}+\frac{1}{2}m\tilde{\omega}^{2}x^{2}\right\}\Psi_{n}(x,t), (6)
Hyhop=Jy(Ψn1(x,t)+Ψn+1(x,t)),\displaystyle H_{\mathrm{yhop}}=J_{y}(\Psi_{n-1}(x,t)+\Psi_{n+1}(x,t)), (7)
Hint=gs|Ψn(x,t)|2Ψn(x,t)\displaystyle H_{\mathrm{int}}=g_{s}|\Psi_{n}(x,t)|^{2}\Psi_{n}(x,t)
+g1(|Ψn1(x,t)|2+|Ψn+1(x,t)|2)Ψn(x,t).\displaystyle\hskip 28.45274pt+g_{1}(|\Psi_{n-1}(x,t)|^{2}+|\Psi_{n+1}(x,t)|^{2})\Psi_{n}(x,t). (8)

We additionally included an overall weak harmonic potential with the oscillator frequency ω~\tilde{\omega} which confines the particles. We take ω~\tilde{\omega} to be much smaller than ω\omega so that its only effect is to prevent particles from flying away. We also added a small intertube hopping term, HyhopH_{\mathrm{yhop}}, which allows an exchange of particles between different tubes. This hopping term allows the system to establish an overall phase coherence. In the xx-pyp_{y} plane, such an intertube hopping term enters as a weak periodic potential of the form 2Jycos(y0py)2J_{y}\cos(y_{0}p_{y}). In the numerical calculations below, we take a~osc1/mω~=10aosc\tilde{a}_{\mathrm{osc}}\equiv 1/\sqrt{m\tilde{\omega}}=10a_{\mathrm{osc}} and Jy=0.01ωJ_{y}=-0.01\omega. The interaction term HintH_{\mathrm{int}} contains the contact interaction with strength gsg_{s} and also a nearest-neighbor (long-range) interaction between tubes with strength g1g_{1}.

Although the interaction HintH_{\mathrm{int}} looks innocuous in the xx-yy plane, it translates into rather unusual infinite and long-ranged interactions in the xx-pyp_{y} plane. Writing the Fourier transform of Ψn(x)\Psi_{n}(x) along the yy direction as Ψ(x,py)\Psi(x,p_{y}), and considering the corresponding second-quantized creation and annihilation operators Ψ^(x,py)\hat{\Psi}^{\dagger}(x,p_{y}) and Ψ^(x,py)\hat{\Psi}(x,p_{y}), the interaction term written in xx-pyp_{y} operators, up to an overall normalization factor, reads

p1+p2=p3+p4\displaystyle\sum_{p_{1}+p_{2}=p_{3}+p_{4}} {gs2+g1cos(y0(p1p2)2)cos(y0(p3p4)2)}\displaystyle\left\{\frac{g_{s}}{2}+g_{1}\cos\left(\frac{y_{0}(p_{1}-p_{2})}{2}\right)\cos\left(\frac{y_{0}(p_{3}-p_{4})}{2}\right)\right\}
𝑑x\displaystyle\int dx\, Ψ^(x,p1)Ψ^(x,p2)Ψ^(x,p3)Ψ^(x,p4).\displaystyle\hat{\Psi}^{\dagger}(x,p_{1})\hat{\Psi}^{\dagger}(x,p_{2})\hat{\Psi}(x,p_{3})\hat{\Psi}(x,p_{4}). (9)

This is a position conserving interaction considering pyp_{y} as a synthetic position coordinate. The contact interaction gsg_{s} is a uniform interaction in the xx-pyp_{y} plane as long as the position is conserved, and thus infinite-ranged. On the other hand, the intertube interaction g1g_{1} has a factor of cos(y0p1p22)cos(y0p3p42)\cos(y_{0}\frac{p_{1}-p_{2}}{2})\cos(y_{0}\frac{p_{3}-p_{4}}{2}); the interaction is strongest when p1p2p_{1}\approx p_{2} and p3p4p_{3}\approx p_{4}, which corresponds to the situation where the interacting particles are close to each other. Therefore, the intertube interaction introduces a shorter-ranged interaction along the pyp_{y} direction compared with the contact interaction. We note, however, that, considering pyp_{y} as a synthetic position, this intertube interaction is still an unconventional form which cannot be derived from a two-body potential depending only on the relative positions between two interacting particles.

Refer to caption
Figure 4: The interacting ground states. The ground-state wave function |Ψn(x)||\Psi_{n}(x)| when gs=aoscωg_{s}=a_{\mathrm{osc}}\omega and g1=0g_{1}=0, plotted in (a) the xx-yy plane, and (b) the corresponding wave function in the xx-pyp_{y} plane, |Ψ(x,py)||\Psi(x,p_{y})|. The ground-state wave function when gs=aoscωg_{s}=a_{\mathrm{osc}}\omega and g1=2aoscωg_{1}=2a_{\mathrm{osc}}\omega, plotted in (c) the xx-yy plane, and (d) the corresponding wave function in the xx-pyp_{y} plane, where dips of the wave function can be observed. (e) The integrated density for each tube, 𝑑x|Ψn(x)|2\int dx|\Psi_{n}(x)|^{2}, plotted as a function of nn; a density modulation is visible. (f) The phase profile of the wave function Ψ(x,py)\Psi(x,p_{y}). The phase singularities (vortices) at py>0p_{y}>0 are marked by blue circles. The units of length in the xx direction are aosca_{\mathrm{osc}}, and the units of pyp_{y} are 1/y01/y_{0}.

We first consider the ground state when only the contact interaction gsg_{s} is present. In Figs. 4(a) and 4(b), we plot the ground state, obtained by imaginary-time propagation of the Gross-Pitaevskii equation (6), when gs=aoscωg_{s}=a_{\mathrm{osc}}\omega and g1=0g_{1}=0, assuming the normalization n𝑑x|Ψn(x,t)|2=1\sum_{n}\int dx|\Psi_{n}(x,t)|^{2}=1. Figure 4(a) shows |Ψn(x)||\Psi_{n}(x)| in the xx-yy plane, whereas Fig. 4(b) shows |Ψ(x,py)||\Psi(x,p_{y})|. We see that the particles are clumped at the center of the xx-pyp_{y} plane, and no vortex-like structure is seen. This is because, in order to obtain vortices in the xx-pyp_{y} plane, we need a short-range repulsive interaction in xx-pyp_{y}, but the interaction due to gsg_{s} is infinite-ranged in the pyp_{y} direction.

To obtain vortices in the xx-pyp_{y} plane, we need to have a long-range interaction along the yy direction, which introduces a shorter-range interaction in the xx-pyp_{y} plane. Technology to realize long-range interactions in synthetic quantum matter has been rapidly developing; representative setups include atoms with dipole-dipole interactions Lahaye:2009 , systems involving Rydberg states Saffman:2010 ; Gorshkov:2011 ; Sevincli:2011 , and atoms in a cavity with light-mediated interactions Ritsch:2013 . Now we include a long-range interaction in the simplest form through the nearest-neighbor interaction g1g_{1}, which can be realized in ultracold atomic gases using dipole-dipole interactions with proper orientation of the dipole field Lahaye:2009 . In Figs. 4(c) and 4(d), we plot the ground state when gs=aoscωg_{s}=a_{\mathrm{osc}}\omega and g1=2aoscωg_{1}=2a_{\mathrm{osc}}\omega. We see that |Ψ(x,py)||\Psi(x,p_{y})| has six dips. By looking at its phase profile in Fig. 4(f), we confirm that these dips are the vortices formed in the xx-pyp_{y} plane. In the xx-yy plane, the state has density peaks at every other tube, as seen in the integrated density for each tube plotted in Fig. 4(e). Vortices on a cylinder are known to have dynamical properties different from those on a plane Ho:2015 ; Guenther:2017 ; our proposal provides a unique setup where such phenomena can be studied. Since the Hamiltonian has a discrete translational symmetry along the angle θ\theta from the xx axis, the state breaks this translational symmetry, thus forming a supersolid state in the xx-yy plane. Whether this supersolid structure persists in the absence of the confining potential ω~\tilde{\omega} in a thermodynamic limit is a question to be addressed in a future work.

We note a crucial difference between previous studies relating long-range interactions, such as a dipole-dipole interaction, and the formation of vortices in the presence of an artificial magnetic field Lahaye:2009 ; Zhang:2016 . In these earlier works, the dipole-dipole interaction, an artificial magnetic field, and vortices all appear in the same (physical) two-dimensional plane. Our result, on the other hand, considers a long-range (intertube) interaction in the xx-yy plane, but the vortices appear in the reciprocal xx-pyp_{y} plane where the artificial magnetic field is realized.

Below, we give a more detailed analysis of the vortices. We first show the wave function profile when a hypothetical contact interaction in the xx-pyp_{y} plane is present. We see that the vortices corresponds to a supersolid structure in the xx-yy plane, confirming that the vortices we observed above are indeed created by the same mechanism as the ordinary vortices in the presence of a magnetic field. We then discuss how vortices form as one increases the nearest-neighbor interaction g1g_{1}.

IV.1 Vortices under a contact interaction in the xx-pyp_{y} plane

We first analyze the bosonic ground state if the interaction is contact in the xx-pyp_{y} plane. Note that this is the situation one often encounters in the xx-yy plane in ultracold atomic gases in the presence of an artificial magnetic field Madison:2000 ; Cooper:2008 ; Fetter:2009 , and one expects the formation of vortices. We include the contact interaction in the xx-pyp_{y} plane by adding a term 𝑑x𝑑pyg~s|Ψ(x,py)|4\int dx\int dp_{y}\tilde{g}_{s}|\Psi(x,p_{y})|^{4} in the energy functional, where Ψ(x,py)\Psi(x,p_{y}) is the normalized wave function in the xx-pyp_{y} plane. We choose g~s=10aoscy0ω\tilde{g}_{s}=10a_{\mathrm{osc}}y_{0}\omega, gs=g1=0g_{s}=g_{1}=0 and we take all the other parameters to be the same as before; in particular, we take tanθ=1\tan\theta=1. In Fig. 5, we plot the ground state obtained by imaginary time evolution of the Gross-Pitaevskii equation. The absolute values of the wave function in the xx-yy as well as the xx-pyp_{y} planes are plotted in Figs. 5(a) and 5(b), respectively. We can see two density peaks in the xx-yy plane, similar to the structure seen in Fig. 4 when the nearest-neighbor interaction is included. The total density of each tube is plotted in Fig. 5(c), where the vertical axis denotes the integral of |Ψn(x)|2|\Psi_{n}(x)|^{2} over xx. The corresponding wave function in the xx-pyp_{y} plane shows the vortex structure, as confirmed by the phase profile in Fig. 5(d). These vortices, formed by an artificial magnetic field and a contact interaction in the xx-pyp_{y} plane, are the ordinary vortices which have been observed in ultracold gases in the presence of an artificial magnetic field in real (xx-yy) space. Comparing these results with the wave function found in Fig. 4, we conclude that the vortex structure found in Fig. 4 has the same origin as the conventional vortices in the presence of an artificial magnetic field.

Refer to caption
Figure 5: The mean-field ground state when we consider a hypothetical contact interaction in the xx-pyp_{y} plane. The units of length in the xx direction are aosca_{\mathrm{osc}}, and the units of pyp_{y} are 1/y01/y_{0}. We take the strength of the contact interaction in the xx-pyp_{y} plane to be g~s=10aoscy0ω\tilde{g}_{s}=10a_{\mathrm{osc}}y_{0}\omega, and no other interparticle interaction exists. Other parameters are tanθ=1\tan\theta=1, ω~=ω/100\tilde{\omega}=\omega/100, and Jy=0.01ωJ_{y}=-0.01\omega. (a) The absolute value of the wave function in the xx-yy plane, where two density peaks are seen. (b) The wave function in the xx-pyp_{y} plane, where two dips of the wave function are seen. (c) The integrated wave function of each tube, 𝑑x|Ψn(x)|2\int dx|\Psi_{n}(x)|^{2}, from which one can see the density peaks more clearly. (d) The phase profile of the wave function in the xx-pyp_{y} plane; the two blue circles indicate the position of the phase singularities, namely the vortices.

IV.2 Vortex formation

Now we come back to the interaction we originally considered: the contact interaction gsg_{s} and the nearest-neighbor interaction g1g_{1}. We analyze how vortices form by varying the value of g1g_{1} and looking at the wave function profile. In Fig. 6, we plot, for three different values of g1g_{1}, the wave function in the xx-yy plane and the xx-pyp_{y} plane, as well as the phase profile of the wave function in the xx-pyp_{y} plane and the integrated density for each tube. We use tanθ=1\tan\theta=1, gs=aoscωg_{s}=a_{\mathrm{osc}}\omega, ω~=ω/100\tilde{\omega}=\omega/100, and Jy=0.01ωJ_{y}=-0.01\omega. We can observe that vortices form as one increases g1g_{1}. We see that, when g1=0.8aoscωg_{1}=0.8a_{\mathrm{osc}}\omega, there is no phase singularity in the xx-pyp_{y} plane, and the integrated density for each tube has a peak at n=0n=0 and monotonically decreases as one goes away from n=0n=0. The situation changes for g1=1.0aoscωg_{1}=1.0a_{\mathrm{osc}}\omega. Although it is still difficult to see vortex formation in |Ψ(x,py)||\Psi(x,p_{y})|, by looking at the phase profile of Ψ(x,py)\Psi(x,p_{y}), we clearly see six phase singularities, indicating the formation of vortices. Correspondingly, density modulation starts to be seen in the integrated density for each tube. When g1=1.2aoscωg_{1}=1.2a_{\mathrm{osc}}\omega, the vortices become easily visible in |Ψ(x,py)||\Psi(x,p_{y})|. The phase singularity and the density modulation also become more pronounced. The vortices are fragile against increase of the intertube hopping amplitude JyJ_{y}. If we use Jy=0.02ωJ_{y}=-0.02\omega, the density dip in the xx-yy plane, and hence the vortex structure in the xx-pyp_{y} plane, similar to the one found for the middle column of Fig. 6, will not appear until we increase g1g_{1} until around g11.3aoscωg_{1}\approx 1.3a_{\mathrm{osc}}\omega. Thus, a stronger interaction is necessary to form vortices. This tendency can be understood in the following way: Intertube hopping JyJ_{y} introduces a sinusoidal potential energy in the xx-pyp_{y} plane along the pyp_{y} direction. To create vortices in the xx-pyp_{y} plane, the interaction effect should be strong enough to overcome the sinusoidal potential created by JyJ_{y}. We note that the vortices can in principle form also when there is no intertube hopping, Jy=0J_{y}=0. In such a case, however, there is no reason for the entire system to develop a phase coherence, and thus the resulting structure in the xx-yy plane is not a supersolid anymore; it would rather be a phase incoherent density wave.

Refer to caption
Figure 6: Formation of vortices in the xx-pyp_{y} plane. The three columns represent the ground states for g1=0.8aoscωg_{1}=0.8a_{\mathrm{osc}}\omega, g1=1.0aoscωg_{1}=1.0a_{\mathrm{osc}}\omega, and g1=1.2aoscωg_{1}=1.2a_{\mathrm{osc}}\omega, respectively, from left to right. In each column, the top row shows |Ψn(x)||\Psi_{n}(x)| of the ground state, the second row shows |Ψ(x,py)||\Psi(x,p_{y})|, the third row shows the phase profile of Ψ(x,py)\Psi(x,p_{y}), and the bottom row shows 𝑑x|Ψn(x)|2\int dx|\Psi_{n}(x)|^{2} as a function of nn. In the phase profile panels, we marked phase singularities at py>0p_{y}>0 by blue circles. (There is also the same number of phase singularities at py<0p_{y}<0, which are not encircled.) The units of length in the xx direction are aosca_{\mathrm{osc}}, and the units of pyp_{y} are 1/y01/y_{0}.

V Experimental platforms

Our proposal is relevant to various platforms of synthetic quantum matter. In ultracold atomic gases, decoupled arrays of one-dimensional tubes can be realized with a two-dimensional optical lattice Bloch:NP . Our proposal can then be realized by superposing another harmonic trapping potential with an appropriate angle. The quantized Hall response and many-body physics are directly relevant in ultracold atomic gases Dauphin:2013 ; Aidelsburger:2015 ; Price:2016 . Various photonic platforms are also relevant to our proposal. For example, exciton-polariton microcavity can be made into an array of one-dimensional wires Boulier:2020 ; Amo:rv1 , and an additional harmonic potential can be applied by varying the width of each wire. Since it is extremely challenging to break the physical time-reversal symmetry in exciton-polaritons to open a topological gap Klembt:2018 , our proposal provides a valuable alternative to explore Chern insulator physics in exciton-polariton microcavities. Photonic cavity arrays and propagating waveguide geometry are also suited to realize our proposal, where harmonic potentials can be applied by varying the cavity or waveguide size along the xx direction Hafezi:2013 ; Rechtsman:2013 ; for these systems it is more natural to consider a lattice structure in the xx direction, which results in a coupled wire setup in the xx-pyp_{y} plane Ozawa:2017 . Edge-related physics are more easily accessible in such photon-related platforms Hafezi:2013 ; Rechtsman:2013 ; Milicevic:2017 .

We note that our proposal depends crucially on the fact that properties in momentum space (pyp_{y} direction) are observable in this synthetic quantum matter. In ultracold gases, the momentum-space wave function is accessible through time-of-flight imaging and, in photon-related systems, far-field imaging provides information on momentum space.

VI Conclusion

We have shown that a set of tubes with harmonic trapping potentials whose centers are shifted is equivalent to charged particles in a magnetic field in the xx-pyp_{y} plane. The underlying idea of our proposal is to replace a momentum in one direction by position to realize a simpler Hamiltonian, which simulates the original Hamiltonian in a hybrid real-momentum space. This method is quite versatile; for example, with a similar mechanism, a collection of one-dimensional Aubry-André model Aubry:1980 (the Harper model) with shifted modulation phases simulates a two-dimensional lattice with a magnetic field, the Harper-Hofstadter model Harper:1955 ; Hofstadter:1976 , in real-momentum space, without breaking the physical time-reversal symmetry. Our proposal provides a simple and powerful method applicable to realize various Hamiltonians, opening a possibility for quantum simulation in synthetic quantum matter.

Acknowledgements.
I am grateful to Bryce Gadway, the discussion with whom led to the initial idea of this paper. I also thank Iacopo Carusotto, Jean Dalibard, Nathan Goldman, Maciej Lewenstein, and Hannah M. Price for useful discussions. I am supported by JSPS KAKENHI Grant Number JP20H01845, JST PRESTO Grant Number JPMJPR19L2, JST CREST Grant Number JPMJCR19T1, RIKEN Incentive Research Project, and the Interdisciplinary Theoretical and Mathematical Sciences Program (iTHEMS) at RIKEN.

References

  • (1) I. Bloch, J. Dalibard, and W. Zwerger, Many-body physics with ultracold gases, Rev. Mod. Phys. 80, 885 (2008).
  • (2) I. Bloch, J. Dalibard, S. Nascimbène , Quantum simulations with ultracold quantum gases, Nature Physics 8, 267-276 (2012).
  • (3) R. Blatt, C. F. Roos, Quantum simulations with trapped ions, Nature Physics 8, 277-284 (2012).
  • (4) A. Aspuru-Guzik, P. Walther, Photonic quantum simulators, Nature Physics 8, 285-291 (2012).
  • (5) I. Carusotto and C. Ciuti, Quantum fluids of light, Rev. Mod. Phys. 85, 299 (2013).
  • (6) I. M. Georgescu, S. Ashhab, and F. Nori, Quantum simulation, Rev. Mod. Phys. 86, 153 (2014).
  • (7) T. Boulier, M. J. Jacquet, A. Maître, G. Lerario, F. Claude, S. Pigeon, Q. Glorieux, A. Bramati, E. Giacobino, A. Amo, J. Bloch, Microcavity Polaritons for Quantum simulation, arXiv:2005.12569 (2020).
  • (8) M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).
  • (9) X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • (10) C.-K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, Classification of topological quantum matter with symmetries, Rev. Mod. Phys. 88, 035005 (2016).
  • (11) N.P. Armitage, E. J. Mele, and A. Vishwanath, Weyl and Dirac semimetals in three-dimensional solids, Rev. Mod. Phys. 90, 015001 (2018).
  • (12) N. R. Cooper, J. Dalibard, and I. B. Spielman, Topological bands for ultracold atoms, Rev. Mod. Phys. 91, 015005 (2019).
  • (13) T. Ozawa, H. M. Price, A. Amo, N. Goldman, M. Hafezi, L. Lu, M. C. Rechtsman, D. Schuster, J. Simon, O. Zilberberg, and I. Carusotto, Topological photonics, Rev. Mod. Phys. 91, 015006 (2019).
  • (14) K. v. Klitzing, G. Dorda, and M. Pepper, New Method for High- Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance, Phys. Rev. Lett. 45, 494 (1980).
  • (15) D. C. Tsui, H. L. Stormer, and A. C. Gossard, Two-Dimensional Magnetotransport in the Extreme Quantum Limit, Phys. Rev. Lett. 48, 1559 (1982).
  • (16) D. Yoshioka, The Quantum Hall Effect (Springer, Berlin, 2002).
  • (17) M. Aidelsburger, S. Nascimbene, and N. Goldman, Artificial gauge fields in materials and engineered systems, Comptes Rendus Physique 19, 394 (2018).
  • (18) K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Vortex Formation in a Stirred Bose-Einstein Condensate, Phys. Rev. Lett. 84, 806 (2000).
  • (19) N. R. Cooper, Rapidly rotating atomic gases, Advances in Physics 57, 539 (2008).
  • (20) A. L. Fetter, Rotating trapped Bose-Einstein condensates, Rev. Mod. Phys. 81, 647 (2009).
  • (21) N. Goldman and J. Dalibard, Periodically Driven Quantum Systems: Effective Hamiltonians and Engineered Gauge Fields, Phys. Rev. X 4, 031027 (2014).
  • (22) N. Goldman, G. Juzeliūnas, P. Öhberg, and I. B. Spielman, Light-induced gauge fields for ultracold atoms, Rep. Prog. Phys. 77, 126401 (2014).
  • (23) T. Ozawa and H. M. Price, Topological quantum matter in synthetic dimensions, Nature Reviews Physics 1, 349-357 (2019).
  • (24) T.-L. Ho and V. B. Shenoy, Local Spin-Gauge Symmetry of the Bose-Einstein Condensates in Atomic Gases, Phys. Rev. Lett. 77, 2595 (1996).
  • (25) V. Pietilä and Mikko Möttönen, Creation of Dirac Monopoles in Spinor Bose-Einstein Condensates, Phys. Rev. Lett. 103, 030401 (2009).
  • (26) J.-y. Choi, S. Kang, S. W. Seo, W. J. Kwon, and Y.-i. Shin, Observation of a Geometric Hall Effect in a Spinor Bose-Einstein Condensate with a Skyrmion Spin Texture, Phys. Rev. Lett. 111, 245301 (2013).
  • (27) M. W. Ray, E. Ruokokoski, S. Kandel, M. Möttönen, and D. S. Hall, Observation of Dirac monopoles in a synthetic magnetic field, Nature 505, 657 (2014).
  • (28) M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit, Photonic Floquet topological insulators, Nature 496, 196 (2013).
  • (29) L. Yuan, Q. Lin, M. Xiao, and S. Fan, Synthetic dimension in photonics, Optica 5, 1396-1405 (2018).
  • (30) 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).
  • (31) M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. M. Taylor, Imaging topological edge states in silicon photonics, Nature Photonics 7, 1001 (2013).
  • (32) N. Schine, A. Ryou, A. Gromov, A. Sommer, and J. Simon, Synthetic Landau levels for photons, Nature 534, 671-675 (2016).
  • (33) L. D. Landau and E. M. Lifshitz, Quantum Mechanics, Third edition, (Butterworth-Heinemann, Oxford, 1977).
  • (34) I. Spielman, Raman processes and effective gauge potentials, Phys. Rev. A 79, 063613 (2009).
  • (35) Y. J. Lin, R. L. Compton, K. Jiménez-García, J. V. Porto, and I. B. Spielman, Synthetic magnetic fields for ultracold neutral atoms, Nature 462, 628 (2009).
  • (36) D. J. Thouless, Quantization of particle transport, Phys. Rev. B 27, 6083 (1983).
  • (37) S. Nakajima, T. Tomita, S. Taie, T. Ichinose, H. Ozawa, L. Wang, M. Troyer, Y. Takahashi, Topological Thouless pumping of ultracold fermions, Nature Physics 12, 296-300 (2016).
  • (38) M. Lohse, C. Schweizer, O. Zilberberg, M. Aidelsburger, and I. Bloch, A Thouless quantum pump with ultracold bosonic atoms in an optical superlattice, Nature Physics 12, 350-354 (2016).
  • (39) R. Laughlin, Quantized Hall conductivity in two dimensions, Phys. Rev. B 23, 5632 (1981).
  • (40) A. Dauphin and N. Goldman, Extracting the Chern Number from the Dynamics of a Fermi Gas: Implementing a Quantum Hall Bar for Cold Atoms, Phys. Rev. Lett. 111, 135302 (2013).
  • (41) M. Aidelsburger, M. Lohse, C. Schweizer, M. Atala, J. T. Barreiro, S. Nascimbène, N. R. Cooper, I. Bloch, and N. Goldman, Measuring the Chern number of Hofstadter bands with ultracold bosonic atoms, Nature Physics 11, 162-166 (2015).
  • (42) H. M. Price, O. Zilberberg, T. Ozawa, I. Carusotto, and N. Goldman, Measurement of Chern numbers through center-of-mass responses, Phys. Rev. B 93, 245113 (2016).
  • (43) T. Lahaye, C. Menotti, L. Santos, M. Lewenstein, and T Pfau, The physics of dipolar bosonic quantum gases, Rep. Prog. Phys. 72, 126401 (2009).
  • (44) M. Saffman, T. G. Walker, and K. Mølmer, Quantum information with Rydberg atoms, Rev. Mod. Phys. 82, 2313 (2010).
  • (45) A. V. Gorshkov, J. Otterbach, M. Fleischhauer, T. Pohl, and M. D. Lukin, Photon-Photon Interactions via Rydberg Blockade, Phys. Rev. Lett. 107, 133602 (2011).
  • (46) S. Sevinçli, N. Henkel, C. Ates, and T. Pohl, Nonlocal Nonlinear Optics in Cold Rydberg Gases, Phys. Rev. Lett. 107, 153001 (2011).
  • (47) H. Ritsch, P. Domokos, F. Brennecke, and T. Esslinger, Cold atoms in cavity-generated dynamical optical potentials, Rev. Mod. Phys. 85, 553 (2013).
  • (48) T.-L. Ho and B. Huang, Spinor Condensates on a Cylindrical Surface in Synthetic Gauge Fields, Phys. Rev. Lett. 115, 155304 (2015).
  • (49) N.-E. Guenther, P. Massignan, and A. L. Fetter, Quantized superfluid vortex dynamics on cylindrical surfaces and planar annuli, Phys. Rev. A 96, 063608 (2017).
  • (50) X.-F. Zhang, L. Wen, C.-Q. Dai, R.-F. Dong, H.-F. Jiang, H. Chang, and S.-G. Zhang, Exotic vortex lattices in a rotating binary dipolar Bose-Einstein condensate, Sci. Rep. 6, 19380 (2016).
  • (51) I. Bloch, Ultracold quantum gases in optical lattices, Nat. Phys. 1, 23 (2005).
  • (52) A. Amo and J. Bloch, Exciton-polaritons in lattices: A non-linear photonic simulator, Comp. Rend. Phys. 17, 805 (2016).
  • (53) S. Klembt, T. H. Harder, O. A. Egorov, K. Winkler, R. Ge, M. A. Bandres, M. Emmerling, L. Worschech, T. C. H. Liew, M. Segev, C. Schneider, and S. Höfling, Exciton-polariton topological insulator, Nature 562, 552 (2018).
  • (54) T. Ozawa and I. Carusotto, Synthetic Dimensions with Magnetic Fields and Local Interactions in Photonic Lattices, Phys. Rev. Lett. 118, 013601 (2017).
  • (55) M. Milićević, T. Ozawa, G. Montambaux, I. Carusotto, E. Galopin, A. Lemaître, L. Le Gratiet, I. Sagnes, J. Bloch, and A. Amo, Orbital edge states in a photonic honeycomb lattice, Phys. Rev. Lett. 118, 107403 (2017).
  • (56) S Aubry, G André, Analyticity breaking and Anderson localization in incommensurate lattices, Ann. Israel Phys. Soc 3, 133 (1980).
  • (57) P. G. Harper, Single Band Motion of Conduction Electrons in a Uniform Magnetic Field, Proc. Phys. Soc. Lond. A 68, 874 (1955).
  • (58) D. R. Hofstadter, Energy levels and wave functions of Bloch electrons in rational and irrational magnetic fields, Phys. Rev. B 14, 2239 (1976).