Thouless pumping in a driven-dissipative Kerr resonator array

S. Ravets sylvain.ravets@c2n.upsaclay.fr Université Paris-Saclay, CNRS, Centre de Nanosciences et de Nanotechnologies (C2N), 91120 Palaiseau, France    N. Pernet Université Paris-Saclay, CNRS, Centre de Nanosciences et de Nanotechnologies (C2N), 91120 Palaiseau, France    N. Mostaan Department of Physics and Arnold Sommerfeld Center for Theoretical Physics (ASC), Ludwig-Maximilians-Universität München, Theresienstr. 37, D-80333 München, Germany Munich Center for Quantum Science and Technology (MCQST), Schellingstr. 4, D-80799 München, Germany CENOLI, Université Libre de Bruxelles, CP 231, Campus Plaine, B-1050 Brussels, Belgium    N. Goldman CENOLI, Université Libre de Bruxelles, CP 231, Campus Plaine, B-1050 Brussels, Belgium Laboratoire Kastler Brossel, Collège de France, CNRS, ENS-PSL University, Sorbonne Université, 11 Place Marcelin Berthelot, 75005 Paris, France    J. Bloch Université Paris-Saclay, CNRS, Centre de Nanosciences et de Nanotechnologies (C2N), 91120 Palaiseau, France
Abstract

Thouless pumping is an emblematic manifestation of topology in physics, referring to the ability to induce a quantized transport of charge across a system by simply varying one of its parameters periodically in time. The original concept of Thouless pumping involves a non-interacting system, and has been implemented in several platforms. One current challenge in the field is to extend this concept to interacting systems. In this article, we propose a Thouless pump that solely relies on nonlinear physics, within a chain of coupled Kerr resonators. Leveraging the driven-dissipative nature of the system, we modulate in space and time the onsite Kerr interaction energies, and generate 1+1-dimensional topological bands in the Bogoliubov spectrum of excitations. These bands present the same topology as the ones obtained within the Harper-Hofstadter framework, and the Wannier states associated to each band experience a net displacement and show quantized transport according to the bands Chern numbers. Remarkably, we find driving configurations leading to band inversion, revealing an interaction-induced topological transition. Our numerical simulations are performed using realistic parameters inspired from exciton polaritons, which form a platform of choice for investigating driven topological phases of matter.

A one-dimensional system of non interacting electrons in a time-modulated periodic potential can show transport in absence of any external bias, a phenomenon known as geometric pumping. Strikingly, there exists modulation protocols where the electron motion is quantized in direct connection to the topology of the 1+1-dimensional (1+1D) bands that the system outlines during the protocol, in analogy to the quantum Hall effet [1]. In the case of filled bands, electrons move during a pump period by an integer number of unit cells equal to the sum of the Chern numbers of the filled bands [2]. Noteworthily, as quantized transport is intrinsically linked to the band structure, it can be accessed by monitoring the motion of the Wannier states. This way, Thouless pumping can also be accessed in bosonic systems.

The development of synthetic experimental platforms, providing exquisite control over the system parameters, has enabled experimental demonstrations of Thouless pumping. Examples include cold atoms confined in tunable optical lattices [3, 4, 5] or coupled to a cavity field [6], photons in coupled waveguides [7, 8, 9], plasmons in plasmonic waveguides [10], or mechanical excitations in metamaterials [11] or waveguides [12].

The picture gets even richer when considering interacting particles [13]. Interactions can induce topological pumping [14, 15], or lead to its breakdown [16, 17] depending on the interaction strength. Nonlinear solitonic waves on top of a Thouless pump have shown to exhibit quantized transport [18, 19, 20, 21, 22]. Thouless pumping has been proposed as a way to channel multiphoton quantum states [23, 24]. Remarkably, interactions in bosonic systems are also known to renormalize the band structure for the elementary Bogoliubov excitations [25], and have been shown to induce non trivial topology in the Bogoliubov bands of an otherwise trivial system [26, 27]. To the best of our knowledge, the question whether one can organize a Thouless pump for the Bogoliubov modes of an interacting bosonic system remains unexplored so far.

In this letter, we propose an optically driven topological pump that entirely relies on the optical nonlinearity in a 1D array of driven-dissipative Kerr resonators. We consider a protocol where an optical drive is periodically distributed across the lattice sites with adiabatic time-periodic intensity modulation. Because of the Kerr nonlinearity, this protocol induces a modulation of the lattice onsite energies, proportionally to the local photon number. As a result, topological bands akin to Harper-Hofstadter bands emerge in the 1+1D Bogoliubov spectrum of excitations and Thouless pumping of the Wannier states is demonstrated. Moreover, we show that anaharmonicities in the modulation protocol may induce topological transitions. We anticipate that this protocol could be implemented using exciton polariton lattices [28, 29], as this platform combines a giant Kerr nonlinearity with the ability to image Bogoliubov excitations both in real and k-space [30, 31, 32, 33]. We therefore base our numerical simulations on realistic parameters that are relevant for the exciton polariton platform.

Refer to caption
Figure 1: a. Schematic of the lattice of coupled Kerr resonators. Each lattice site is individually driven (zigzag arrows). b. Energy dispersion of the lattice modes in the linear regime. The horizontal dashed line shows the energy ωPPlanck-constant-over-2-pisubscript𝜔P\hbar\omega_{\rm P}roman_ℏ italic_ω start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT of the drives.

We consider an infinite periodic chain of identical Kerr resonators with constant couplings along the chain (see Fig. 1a). The linear part of the Hamiltonian writes:

H^=nE0|ϕnϕn|J(|ϕnϕn+1|+|ϕnϕn1|),^𝐻subscript𝑛subscript𝐸0subscriptitalic-ϕ𝑛subscriptitalic-ϕ𝑛𝐽subscriptitalic-ϕ𝑛subscriptitalic-ϕ𝑛1subscriptitalic-ϕ𝑛subscriptitalic-ϕ𝑛1\displaystyle\hat{H}=\sum_{n}E_{0}\outerproduct{\phi_{n}}{\phi_{n}}-J\left(% \outerproduct{\phi_{n}}{\phi_{n+1}}+\outerproduct{\phi_{n}}{\phi_{n-1}}\right)\ ,over^ start_ARG italic_H end_ARG = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG | - italic_J ( | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT end_ARG | + | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT end_ARG | ) , (1)

where |ϕnketsubscriptitalic-ϕ𝑛\left|\phi_{n}\right>| italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ is the mode of resonator n𝑛nitalic_n with on-site energy E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and J𝐽Jitalic_J is the coupling between neighboring sites. The eigenenergies of the Bloch eigenstates form a single band of dispersion E(k)=E02Jcos((ka))𝐸𝑘subscript𝐸02𝐽𝑘𝑎E(k)=E_{0}-2J\cos{(ka)}italic_E ( italic_k ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 2 italic_J roman_cos ( start_ARG ( italic_k italic_a ) end_ARG ), where a𝑎aitalic_a is the lattice period and k]π/a,π/a]k\in]-\pi/a,\pi/a]italic_k ∈ ] - italic_π / italic_a , italic_π / italic_a ] (see Fig. 1b). The nonlinear resonators are coherently driven by a monochromatic field oscillating at angular frequency ωpsubscript𝜔p\omega_{\rm p}italic_ω start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. The complex field amplitude at site n𝑛nitalic_n is Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, and all sites experience a uniform loss rate γ𝛾\gammaitalic_γ. In the mean field approximation, the system dynamics is governed by the following driven-dissipative Gross Pitaevskii equation:

it𝝍=𝑖Planck-constant-over-2-pisubscript𝑡𝝍absent\displaystyle i\hbar\partial_{t}\boldsymbol{\psi}=italic_i roman_ℏ ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT bold_italic_ψ = (H^iγ2𝟙^+gN^)𝝍+i𝑭eiωpt,^𝐻𝑖𝛾2^1𝑔^𝑁𝝍𝑖𝑭superscript𝑒𝑖subscript𝜔p𝑡\displaystyle\left(\hat{H}-i\frac{\gamma}{2}\hat{\mathbb{1}}+g\hat{N}\right)% \boldsymbol{\psi}+i\boldsymbol{F}e^{-i\omega_{\rm p}t}\,,( over^ start_ARG italic_H end_ARG - italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG over^ start_ARG blackboard_1 end_ARG + italic_g over^ start_ARG italic_N end_ARG ) bold_italic_ψ + italic_i bold_italic_F italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT , (2)

where g𝑔gitalic_g is the Kerr nonlinearity, 𝝍𝝍\boldsymbol{\psi}bold_italic_ψ is a vector of components ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, N^^𝑁\hat{N}over^ start_ARG italic_N end_ARG is a diagonal operator with non-zero elements equal to N^n,n=|ψn|2subscript^𝑁𝑛𝑛superscriptsubscript𝜓𝑛2\hat{N}_{n,n}=|\psi_{n}|^{2}over^ start_ARG italic_N end_ARG start_POSTSUBSCRIPT italic_n , italic_n end_POSTSUBSCRIPT = | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, 𝟙^^1\hat{\mathbb{1}}over^ start_ARG blackboard_1 end_ARG is the identity matrix, and 𝑭𝑭\boldsymbol{F}bold_italic_F is a vector of components Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. In the rest of this letter, we define the energy detuning between the drive and the top of the band ΔE=ωPE02JΔ𝐸Planck-constant-over-2-pisubscript𝜔Psubscript𝐸02𝐽\Delta E=\hbar\omega_{\rm P}-E_{0}-2Jroman_Δ italic_E = roman_ℏ italic_ω start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 2 italic_J (see Fig. 1b), and the drive power P=n|Fn|2𝑃subscript𝑛superscriptsubscript𝐹𝑛2P=\sum_{n}\,\left|F_{n}\right|^{2}italic_P = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

We propose a drive protocol where we spatially modulate the pump amplitudes as follows:

Fn(φ)=F|cos(φ/2+παn)|,subscript𝐹𝑛𝜑𝐹𝜑2𝜋𝛼𝑛F_{n}(\varphi)=F\left|\cos(\varphi/2+\pi\alpha\,n)\right|\,,italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_φ ) = italic_F | roman_cos ( start_ARG italic_φ / 2 + italic_π italic_α italic_n end_ARG ) | , (3)

where α=1/5𝛼15\alpha=1/5italic_α = 1 / 5 and the parameter φ𝜑\varphiitalic_φ is a phason that can take values between 00 and 2π2𝜋2\pi2 italic_π. Note that the spatial periodicity of this protocol is 5a5𝑎5a5 italic_a, meaning that each unit cell of this drive pattern contains five sites.

We first set φ=0𝜑0\varphi=0italic_φ = 0 and compute the system nonlinear steady-state versus P𝑃Pitalic_P, solving Eq. 2. We plot, in Fig. 2a, the stationary field intensity pattern |ψn(s)|2superscriptsuperscriptsubscript𝜓𝑛𝑠2\left|\psi_{n}^{(s)}\right|^{2}| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT calculated for increasing values of P𝑃Pitalic_P. The intensity pattern follows the periodicity of the driving field, and evidences a nonlinear threshold at P=Pth𝑃subscript𝑃thP=P_{\rm th}italic_P = italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT, characteristic of a multi-stable nonlinear system. In the following, unless stated otherwise, we set P/Pth=0.79𝑃subscript𝑃th0.79P/P_{\rm th}=0.79italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 0.79 (red dashed line in Fig. 2a).

Refer to caption
Figure 2: a. Stationary intensity map versus unit cell number and power for the lattice driven according to φ=0𝜑0\varphi=0italic_φ = 0 in Eq. 3. The dotted horizontal line marks the drive power P=0.79Pth𝑃0.79subscript𝑃thP=0.79P_{\rm th}italic_P = 0.79 italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT used in most part of this letter. b. Nonlinear onsite energies 2g|ψn(s)(Fn(φ))|22𝑔superscriptsuperscriptsubscript𝜓𝑛𝑠subscript𝐹𝑛𝜑22g\left|\psi_{n}^{(s)}(F_{n}(\varphi))\right|^{2}2 italic_g | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT ( italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_φ ) ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as a function of φ𝜑\varphiitalic_φ. The red arrow is a guide to the eye to highlight the trajectory for one energy minimum. The vertical white dashed lines in a. and b. delimit the five sites forming one unit cell. c) Top panel: the population |ψ3(s)|2superscriptsuperscriptsubscript𝜓3𝑠2\left|\psi_{3}^{(s)}\right|^{2}| italic_ψ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT on site 3 of the unit cell is shown as a function of φ𝜑\varphiitalic_φ for P/Pth=0.79𝑃subscript𝑃th0.79P/P_{\rm th}=0.79italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 0.79 (solid blue line) and for P/Pth=0.996𝑃subscript𝑃th0.996P/P_{\rm th}=0.996italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 0.996 (dashed red line). The associated power spectra (normalized to the power in the fundamental harmonic |1|2superscriptsubscript12\left|\mathcal{F}_{1}\right|^{2}| caligraphic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) are plotted in the bottom panel. The numerical parameters are γ=44.7μeV,J/γ=2.2,g/γ=103,ΔE=5.14Jformulae-sequence𝛾44.7𝜇eVformulae-sequence𝐽𝛾2.2formulae-sequence𝑔𝛾superscript103Δ𝐸5.14𝐽\gamma=44.7~{}{\rm\mu eV},\,J/\gamma=2.2,\,g/\gamma=10^{-3},\,\Delta E=5.14Jitalic_γ = 44.7 italic_μ roman_eV , italic_J / italic_γ = 2.2 , italic_g / italic_γ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , roman_Δ italic_E = 5.14 italic_J.

We now compute the excitation spectrum on top of this nonlinear steady state. To do so, we use a Bogoliubov approach writing: 𝝍(t)=(𝝍(s)+δ𝝍(t))eiωPt𝝍𝑡superscript𝝍𝑠𝛿𝝍𝑡superscript𝑒𝑖subscript𝜔P𝑡\boldsymbol{\psi}(t)=(\boldsymbol{\psi}^{(s)}+\delta\boldsymbol{\psi}(t))e^{-i% \omega_{\rm P}t}bold_italic_ψ ( italic_t ) = ( bold_italic_ψ start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT + italic_δ bold_italic_ψ ( italic_t ) ) italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT roman_P end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT. Introducing this ansatz into the Gross-Pitaevskii equation, we obtain the following linearized equations for δ𝝍(t)𝛿𝝍𝑡\delta\boldsymbol{\psi}(t)italic_δ bold_italic_ψ ( italic_t ) and δ𝝍(t)𝛿superscript𝝍𝑡\delta\boldsymbol{\psi}^{*}(t)italic_δ bold_italic_ψ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t ):

it(δ𝝍(t)δ𝝍(t))=(^𝒩^𝒩^^)(δ𝝍(t)δ𝝍(t)),𝑖Planck-constant-over-2-pisubscript𝑡matrix𝛿𝝍𝑡𝛿superscript𝝍𝑡matrix^^𝒩superscript^𝒩superscript^matrix𝛿𝝍𝑡𝛿superscript𝝍𝑡\displaystyle i\hbar\partial_{t}\begin{pmatrix}\delta\boldsymbol{\psi}(t)\\ \delta\boldsymbol{\psi}^{*}(t)\end{pmatrix}=\begin{pmatrix}\hat{\mathcal{M}}&% \hat{\mathcal{N}}\\ -\hat{\mathcal{N}}^{*}&-\hat{\mathcal{M}}^{*}\end{pmatrix}\begin{pmatrix}% \delta\boldsymbol{\psi}(t)\\ \delta\boldsymbol{\psi}^{*}(t)\end{pmatrix}\,,italic_i roman_ℏ ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_δ bold_italic_ψ ( italic_t ) end_CELL end_ROW start_ROW start_CELL italic_δ bold_italic_ψ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t ) end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL over^ start_ARG caligraphic_M end_ARG end_CELL start_CELL over^ start_ARG caligraphic_N end_ARG end_CELL end_ROW start_ROW start_CELL - over^ start_ARG caligraphic_N end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL - over^ start_ARG caligraphic_M end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_δ bold_italic_ψ ( italic_t ) end_CELL end_ROW start_ROW start_CELL italic_δ bold_italic_ψ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t ) end_CELL end_ROW end_ARG ) , (4)

where ^=H^(ωP+iγ2)𝟙^+2gN^(s)^^𝐻Planck-constant-over-2-pisubscript𝜔𝑃𝑖𝛾2^12𝑔superscript^𝑁𝑠\hat{\mathcal{M}}=\hat{H}-\big{(}\hbar\omega_{P}+i\frac{\gamma}{2}\big{)}\hat{% \mathbb{1}}+2g\hat{N}^{(s)}over^ start_ARG caligraphic_M end_ARG = over^ start_ARG italic_H end_ARG - ( roman_ℏ italic_ω start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ) over^ start_ARG blackboard_1 end_ARG + 2 italic_g over^ start_ARG italic_N end_ARG start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT, and 𝒩^^𝒩\hat{\mathcal{N}}over^ start_ARG caligraphic_N end_ARG is a diagonal operator with non-zero elements equal to 𝒩^n,n=g(ψn(s))2subscript^𝒩𝑛𝑛𝑔superscriptsuperscriptsubscript𝜓𝑛𝑠2\hat{\mathcal{N}}_{n,n}=g\left(\psi_{n}^{(s)}\right)^{2}over^ start_ARG caligraphic_N end_ARG start_POSTSUBSCRIPT italic_n , italic_n end_POSTSUBSCRIPT = italic_g ( italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

The stationary solutions of Eq. 4 have the form: δ𝝍(t)=𝒖eit/+𝒗eit/𝛿𝝍𝑡𝒖superscript𝑒𝑖𝑡Planck-constant-over-2-pisuperscript𝒗superscript𝑒𝑖superscript𝑡Planck-constant-over-2-pi\delta\boldsymbol{\psi}(t)=\boldsymbol{u}e^{-i\mathcal{E}t/\hbar}+\boldsymbol{% v}^{*}e^{i\mathcal{E}^{*}t/\hbar}italic_δ bold_italic_ψ ( italic_t ) = bold_italic_u italic_e start_POSTSUPERSCRIPT - italic_i caligraphic_E italic_t / roman_ℏ end_POSTSUPERSCRIPT + bold_italic_v start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i caligraphic_E start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_t / roman_ℏ end_POSTSUPERSCRIPT, where (𝒖,𝒗)Tsuperscript𝒖𝒗T(\boldsymbol{u},\boldsymbol{v})^{\mathrm{T}}( bold_italic_u , bold_italic_v ) start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT is an eigenvector of the Bogoliubov matrix written in Eq. 4 for the eigenvalue \mathcal{E}caligraphic_E. The set of all possible Re()Re\operatorname{Re}(\mathcal{E})roman_Re ( caligraphic_E ) values gives rise to the Bogoliubov spectrum of excitations. It is convenient to write the Bogoliubov eigenmodes in reciprocal space versus wavevector k𝑘kitalic_k, using the basis of Bloch states (𝒖~(k),𝒗~(k))Tsuperscriptbold-~𝒖𝑘bold-~𝒗𝑘T(\boldsymbol{\tilde{u}}(k),\boldsymbol{\tilde{v}}(k))^{\mathrm{T}}( overbold_~ start_ARG bold_italic_u end_ARG ( italic_k ) , overbold_~ start_ARG bold_italic_v end_ARG ( italic_k ) ) start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT. For each site i𝑖iitalic_i of the unit cell, 𝒖~i(k)subscriptbold-~𝒖𝑖𝑘\boldsymbol{\tilde{u}}_{i}(k)overbold_~ start_ARG bold_italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_k ) and 𝒗~i(k)subscriptbold-~𝒗𝑖𝑘\boldsymbol{\tilde{v}}_{i}(k)overbold_~ start_ARG bold_italic_v end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_k ) are called the “normal” and “ghost” branch amplitudes.

Figure 3a shows the resulting Bogoliubov spectrum versus k𝑘kitalic_k. Because of the 5a5𝑎5a5 italic_a spatial period imposed by the drive, k𝑘kitalic_k takes values within ]π/5a,π/5a]]-\pi/5a,\pi/5a]] - italic_π / 5 italic_a , italic_π / 5 italic_a ]. We clearly observe the presence of 10 bands separated by energy gaps. By construction, the bands are symmetric with respect to the pump energy Re()=0Re0\operatorname{Re}(\mathcal{E})=0roman_Re ( caligraphic_E ) = 0 (particle-hole symmetry). The five low energy bands correspond to the “normal” modes (|𝒖~(k)|2|𝒗~(k)|2=1superscriptbold-~𝒖𝑘2superscriptbold-~𝒗𝑘21|\boldsymbol{\tilde{u}}(k)|^{2}-|\boldsymbol{\tilde{v}}(k)|^{2}=1| overbold_~ start_ARG bold_italic_u end_ARG ( italic_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | overbold_~ start_ARG bold_italic_v end_ARG ( italic_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1), while the five upper energy bands correspond to the “ghost” modes (|𝒖~(k)|2|𝒗~(k)|2=1superscriptbold-~𝒖𝑘2superscriptbold-~𝒗𝑘21|\boldsymbol{\tilde{u}}(k)|^{2}-|\boldsymbol{\tilde{v}}(k)|^{2}=-1| overbold_~ start_ARG bold_italic_u end_ARG ( italic_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | overbold_~ start_ARG bold_italic_v end_ARG ( italic_k ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = - 1). This spectrum highlights that spatially engineering the driving field enables tailoring the Bogoliubov band structure [26].

Refer to caption
Figure 3: a. Bogoliubov bands on top of the nonlinear steady-state ψ(s)superscript𝜓𝑠\psi^{(s)}italic_ψ start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT. The normal (ghost) branches are shown in solid (dashed) lines. b. 1+1D Bogoliubov bands as a function of k𝑘kitalic_k and φ𝜑\varphiitalic_φ. The computed Chern numbers are given aside each band. γ=44.7μeV,J/γ=2.2,g/γ=103,ΔE=5.14J,P/Pth=0.79formulae-sequence𝛾44.7𝜇eVformulae-sequence𝐽𝛾2.2formulae-sequence𝑔𝛾superscript103formulae-sequenceΔ𝐸5.14𝐽𝑃subscript𝑃th0.79\gamma=44.7~{}{\rm\mu eV},\,J/\gamma=2.2,\,g/\gamma=10^{-3},\,\Delta E=5.14J,% \,P/P_{\rm th}=0.79italic_γ = 44.7 italic_μ roman_eV , italic_J / italic_γ = 2.2 , italic_g / italic_γ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , roman_Δ italic_E = 5.14 italic_J , italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 0.79

.

We now consider a protocol where φ𝜑\varphiitalic_φ is adiabatically varied between 00 and 2π2𝜋2\pi2 italic_π. As we vary φ𝜑\varphiitalic_φ, the diagonal elements of the Bogoliubov matrix are periodically modulated according to ^n,n(φ)=E0+2g|ψn(s)(Fn(φ))|2(ωP+iγ2)subscript^𝑛𝑛𝜑subscript𝐸02𝑔superscriptsuperscriptsubscript𝜓𝑛𝑠subscript𝐹𝑛𝜑2Planck-constant-over-2-pisubscript𝜔𝑃𝑖𝛾2\hat{\mathcal{M}}_{n,n}(\varphi)=E_{0}+2g\left|\psi_{n}^{(s)}(F_{n}(\varphi))% \right|^{2}-\big{(}\hbar\omega_{P}+i\frac{\gamma}{2}\big{)}over^ start_ARG caligraphic_M end_ARG start_POSTSUBSCRIPT italic_n , italic_n end_POSTSUBSCRIPT ( italic_φ ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 2 italic_g | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT ( italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_φ ) ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( roman_ℏ italic_ω start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ). Strikingly, ^^\hat{\mathcal{M}}over^ start_ARG caligraphic_M end_ARG has the form of an Aubry-André-Harper Hamiltonian [34, 35], where the onsite energies are modulated as a function of φ𝜑\varphiitalic_φ, while the coupling coefficients J𝐽Jitalic_J stay constant. As highlighted in Ref. [36, 7], one can see φ𝜑\varphiitalic_φ as a parametric dimension in a 1+1D space. Every realization of ^(φ)^𝜑\hat{\mathcal{M}}(\varphi)over^ start_ARG caligraphic_M end_ARG ( italic_φ ) is a Fourier component of a 2D operator ^2Dsubscript^2D\hat{\mathcal{M}}_{\rm 2D}over^ start_ARG caligraphic_M end_ARG start_POSTSUBSCRIPT 2 roman_D end_POSTSUBSCRIPT acting on the 1+1D system. For a pure sine-modulation of the on-site energies, ^2Dsubscript^2D\hat{\mathcal{M}}_{\rm 2D}over^ start_ARG caligraphic_M end_ARG start_POSTSUBSCRIPT 2 roman_D end_POSTSUBSCRIPT exactly coincides with the Harper-Hofstadter Hamiltonian [37] describing particles on a square lattice with a non-zero flux per plaquette equal to 2πα2𝜋𝛼2\pi\alpha2 italic_π italic_α.

In Fig. 2b, we show the nonlinear potential 2g|ψn(s)(Fn(φ))|22𝑔superscriptsuperscriptsubscript𝜓𝑛𝑠subscript𝐹𝑛𝜑22g\left|\psi_{n}^{(s)}(F_{n}(\varphi))\right|^{2}2 italic_g | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT ( italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_φ ) ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT evolution versus φ𝜑\varphiitalic_φ. We observe that it defines a series of potential wells that get deformed as we vary φ𝜑\varphiitalic_φ, in a way that they exhibit a unidirectional motion, as expected for an adiabatic pump. The nonlinear response of the Kerr resonators leads to anharmonic modulations of the nonlinear onsite energies, so that the mapping to the Harper-Hofstadter Hamiltonian is not exact. This is illustrated in Fig. 2c, where |ψ3(s)|2superscriptsuperscriptsubscript𝜓3𝑠2\left|\psi_{3}^{(s)}\right|^{2}| italic_ψ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_s ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (population in the third site of the unit cell) is plotted for P/Pth=79%𝑃subscript𝑃thpercent79P/P_{\rm th}=79~{}\%italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 79 % (blue solid line) and P/Pth=99.6%𝑃subscript𝑃thpercent99.6P/P_{\rm th}=99.6~{}\%italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 99.6 % (red dashed). The corresponding power spectra are shown on the lower panel, where we evidence the increase of the anharmonicity as P𝑃Pitalic_P approaches Pthsubscript𝑃thP_{\rm th}italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT from below. We show in the following that the band structure topology of the Harper-Hofstadter model persists as long as P/Pth<99.6%𝑃subscript𝑃thpercent99.6P/P_{\rm th}<99.6~{}\%italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT < 99.6 %, illustrating its robustness against anharmonicities in the potential.

We plot, in Fig. 3b, the 1+1D (versus k𝑘kitalic_k and φ𝜑\varphiitalic_φ) Bogoliubov energy bands outlined during the drive protocol. We observe that all gaps described in Fig. 2 remain open for all φ𝜑\varphiitalic_φ values. For the five normal and the five ghost branches, we then compute Chern numbers [38, 39] and find 𝒞=1, 1,4, 1, 1𝒞11411\mathcal{C}=1,\,1,\,-4,\,1,\,1caligraphic_C = 1 , 1 , - 4 , 1 , 1. These values correspond to the Chern numbers expected for the Harper-Hofstadter model with magnetic flux α=1/5𝛼15\alpha=1/5italic_α = 1 / 5. Adjusting the drive protocol, we point out that one can straightforwardly extend this demonstration to any rational flux α=p/q𝛼𝑝𝑞\alpha=p/qitalic_α = italic_p / italic_q, where q𝑞qitalic_q is the number of sites per unit cell. These results establish adiabatic drive modulation of a nonlinear fluid of light as a powerful tool to induce non-trivial topology in an effectively 2D space with one synthetic parametric dimension.

Refer to caption
Figure 4: Intensity distributions |un|2superscriptsubscript𝑢𝑛2\left|u_{n}\right|^{2}| italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of a Wannier state located at a chosen lattice position as a function of φ𝜑\varphiitalic_φ. From bottom to top, we represent Wannier states in each of the five normal bands ordered by increasing energies. The red lines trace the barycenter motion of each Wannier state during one protocol period. The white dashed lines delimit one unit cell. The intensity distributions are normalized to their maximum values. All numerical parameters are identical to the ones used in Fig. 3.

To evidence topological pumping for the Bogoliubov excitations, we now investigate the Wannier centers and their evolution during the drive protocol. For every value of φ𝜑\varphiitalic_φ, we compute the maximally-localized Wannier states associated to the five lowest energy (“normal”) bands. For each normal band, we show in Fig 4, the intensity distribution |un|2superscriptsubscript𝑢𝑛2\left|u_{n}\right|^{2}| italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the Wannier state located at a chosen unit cell, and we monitor its evolution as a function of φ𝜑\varphiitalic_φ. The solid lines follow the barycenter trajectories for these Wannier states, and clearly highlight a net motion of the Wannier centers. Moreover, after one modulation period, each Wannier state has moved by a quantized number of unit cells that exactly matches the Chern number of the band. This fully demonstrates that we have realized a topological pump for Bogoliubov excitations. The crucial novelty of our result lies in the fact that the pumping mechanism relies entirely on the presence of inter-particle interactions which modulate the potential acting on Bogoliubov excitations.

Refer to caption
Figure 5: a. Energy gaps ΔijsubscriptΔ𝑖𝑗\Delta_{ij}roman_Δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT versus P/Pth𝑃subscript𝑃thP/P_{\rm th}italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT. b. Logarithmic plot of Δ34/JsubscriptΔ34𝐽\Delta_{34}/Jroman_Δ start_POSTSUBSCRIPT 34 end_POSTSUBSCRIPT / italic_J versus P/Pth𝑃subscript𝑃thP/P_{\rm th}italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT for ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0 (solid line) and ϵ=0.06italic-ϵ0.06\epsilon=0.06italic_ϵ = 0.06 (dashed line). c. Intensity distributions |un|2superscriptsubscript𝑢𝑛2\left|u_{n}\right|^{2}| italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT versus φ𝜑\varphiitalic_φ for the Wannier states associated to the third (bottom) and fourth (top) Bogoliubov bands obtained for ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0, just below Pthsubscript𝑃thP_{\rm th}italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT (P/Pth=99.99%𝑃subscript𝑃thpercent99.99P/P_{\rm th}=99.99~{}\%italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 99.99 %). The red lines trace the barycenter motion of each Wannier state during one protocol period. The intensity distributions are normalized to their maximum values.

So far, we have focused on a value of P𝑃Pitalic_P lying well below Pthsubscript𝑃thP_{\rm th}italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT. As discussed in Fig. 2c, the nonlinear character of the modulation protocol, leads to power dependent anharmonicities in the modulated onsite energies. In Figure 5a, we plot the evolution of the energy gaps ΔijsubscriptΔ𝑖𝑗\Delta_{ij}roman_Δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT between bands i𝑖iitalic_i and j𝑗jitalic_j, as we increase P𝑃Pitalic_P up to Pthsubscript𝑃thP_{\rm th}italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT. Interestingly, at a critical power Pc/Pth=99.6%subscript𝑃𝑐subscript𝑃thpercent99.6P_{c}/P_{\rm{th}}=99.6\%italic_P start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT = 99.6 %, we observe that Δ34subscriptΔ34\Delta_{34}roman_Δ start_POSTSUBSCRIPT 34 end_POSTSUBSCRIPT vanishes, and then the gap reopens (see semi-logarithmic plot in Fig.5b). For power values within the range (Pc<P<Pthsubscript𝑃𝑐𝑃subscript𝑃thP_{c}<P<P_{\rm{th}}italic_P start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT < italic_P < italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT), we repeat our analysis of the resulting pump. We first compute the Chern numbers for the 1+1D Bogoliubov bands, and obtain 𝒞=1, 1, 1,4, 1𝒞11141\mathcal{C}=1,\,1,\,1,\,-4,\,1caligraphic_C = 1 , 1 , 1 , - 4 , 1. Compared to the situation discussed previously in Fig. 3b, we do not recover the same Chern number ordering, as the Chern numbers of the third and fourth bands are inverted. We then calculate the Wannier centers evolution and find that their barycenter motion follows the new Chern number ordering. Indeed, the Wannier state motions in the third and fourth bands (see Fig. 5c), are inverted as compared to Fig. 4, while the motions stay unchanged in the rest of the bands (not shown). These results evidence the occurrence of a topological transition induced by the nonlinearity and the resulting anharmonic modulation of the onsite energies. We note that the n𝑛nitalic_nth harmonic introduces nonzero n𝑛nitalic_nth-neighbor coupling in the parametric dimension, so that deviations in the mapping to the 2D Harper-Hofstadter model without long range couplings can no longer be neglected, and the system experiences a topological transition.

Increasing P𝑃Pitalic_P to values P>Pth𝑃subscript𝑃thP>P_{\rm th}italic_P > italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT, we observe that the anharmonicity of the onsite energy modulation decreases, and we recover the Harper-Hofstadter Chern ordering. These results illustrate an asset of our pump protocol, whereby the drive amplitude can be used as a knob to control the topology that emerges in the Bogoliubov spectrum, in a way that Bogoliubov excitations experience topological transitions. As a further tuning knob, we now show that one can control Pcsubscript𝑃𝑐P_{c}italic_P start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT by modifying the frequency spectrum of the drive protocol. We illustrate this feature by studying the evolution of Δ34subscriptΔ34\Delta_{34}roman_Δ start_POSTSUBSCRIPT 34 end_POSTSUBSCRIPT versus P/Pth𝑃subscript𝑃thP/P_{\rm{th}}italic_P / italic_P start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT when adding a second order harmonic to the drive:

Fn(φ)=F|cos(φ/2+παn)|+ϵF|cos(φ+2παn)|.subscript𝐹𝑛𝜑𝐹𝜑2𝜋𝛼𝑛italic-ϵ𝐹𝜑2𝜋𝛼𝑛F_{n}(\varphi)=F\left|\cos(\varphi/2+\pi\alpha\,n)\right|+\epsilon F\left|\cos% (\varphi+2\pi\alpha\,n)\right|\,.italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_φ ) = italic_F | roman_cos ( start_ARG italic_φ / 2 + italic_π italic_α italic_n end_ARG ) | + italic_ϵ italic_F | roman_cos ( start_ARG italic_φ + 2 italic_π italic_α italic_n end_ARG ) | . (5)

The result is shown in Fig. 5b (dashed line) for ϵ=0.06italic-ϵ0.06\epsilon=0.06italic_ϵ = 0.06. Under the modified drive protocol, the critical power at which the topological transition occurs gets significantly reduced to the value Pc=0.85Pthsubscriptsuperscript𝑃𝑐0.85subscriptsuperscript𝑃thP^{{}^{\prime}}_{c}=0.85P^{{}^{\prime}}_{\rm{th}}italic_P start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.85 italic_P start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_th end_POSTSUBSCRIPT. This confirms that the observed gap closure and ensuing topological transition are rooted in the evolution of modulation anharmonicities as the nonlinearity develops. We note that tuning the number and amplitudes of harmonics included in the drive protocol provides a vast nonlinear optics framework to trigger topological transitions and control the motion of Wannier states in the Bogoliubov excitation spectrum. A systematic study of these effects is beyond the scope of this letter.

In conclusion, starting with a trivial lattice of driven-dissipative Kerr resonators, we establish Bogoliubov excitations as a playground to organize Thouless pumping through spatial engineering of the drive. Adiabatically modulating the drive pattern introduces a parametric dimension leading to 1+1D topological bands. As a consequence of the presence of non-zero Chern numbers, the Bogoliubov Wannier states associated to the bands experience quantized motion through Thouless pumping. Moreover, the nonlinearity modifies the frequency spectrum of the onsite energy modulation, leading to topological transitions that switch the directionality and amplitude of the quantized motion of Wannier states within different bands. We emphasize that all our simulations are performed using realistic parameters that are compatible with the exciton polariton platform. Benefiting from recent progress in probing and manipulating Bogoliubov excitations in these systems [32, 33, 40], we foresee that the experimental implementation of the proposed topological pump is within reach. As a further development, we envision that drive engineering can be used to induce uniform magnetic flux for Bogoliubov excitations in 2D lattices. By extending the perturbative approach of Ref. [26], one could realize Hofstadter-type models with large synthetic flux Φπsimilar-toΦ𝜋\Phi\sim\piroman_Φ ∼ italic_π per plaquette, hence offering new opportunities to explore nonlinear topology in 2D lattices [41, 42].

Acknowledgements.
We thank A. Amo, I. Carusotto, and O. Zilberberg for fruitful discussions. This work was partly supported by the Paris Ile de France Région in the framework of DIM SIRTEQ, by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (project ARQADIA, grant agreement no. 949730), and under Horizon Europe research and innovation programme (ANAPOLIS, grant agreement no. 101054448). NG is supported by the ERC Grant LATIS and the EOS project CHEQS. NM acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – EXC-2111 – 390814868.

References

  • Thouless [1983] D. J. Thouless, Quantization of particle transport, Phys. Rev. B 27, 6083 (1983).
  • Citro and Aidelsburger [2023] R. Citro and M. Aidelsburger, Thouless pumping and topology, Nature Reviews Physics 5, 87 (2023).
  • Lohse et al. [2016] 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 (2016).
  • Nakajima et al. [2016] S. Nakajima, T. Tomita, S. Taie, T. Ichinose, H. Ozawa, L. Wang, M. Troyer, and Y. Takahashi, Topological Thouless pumping of ultracold fermions, Nature Physics 12, 296 (2016).
  • Lu et al. [2016] H.-I. Lu, M. Schemmer, L. M. Aycock, D. Genkina, S. Sugawa, and I. B. Spielman, Geometrical pumping with a bose-einstein condensate, Phys. Rev. Lett. 116, 200402 (2016).
  • Dreon et al. [2022] D. Dreon, A. Baumgärtner, X. Li, S. Hertlein, T. Esslinger, and T. Donner, Self-oscillating pump in a topological dissipative atom-cavity system, Nature 608, 494 (2022).
  • Kraus et al. [2012] Y. E. Kraus, Y. Lahini, Z. Ringel, M. Verbin, and O. Zilberberg, Topological states and adiabatic pumping in quasicrystals, Phys. Rev. Lett. 109, 106402 (2012).
  • Ke et al. [2016] Y. Ke, X. Qin, F. Mei, H. Zhong, Y. S. Kivshar, and C. Lee, Topological phase transitions and thouless pumping of light in photonic waveguide arrays, Laser & Photonics Reviews 10, 995 (2016).
  • Cerjan et al. [2020] A. Cerjan, M. Wang, S. Huang, K. P. Chen, and M. C. Rechtsman, Thouless pumping in disordered photonic systems, Light: Science & Applications 9, 178 (2020).
  • Fedorova et al. [2020] Z. Fedorova, H. Qiu, S. Linden, and J. Kroha, Observation of topological transport quantization by dissipation in fast Thouless pumps, Nature Communications 11, 3758 (2020).
  • Grinberg et al. [2020] I. H. Grinberg, M. Lin, C. Harris, W. A. Benalcazar, C. W. Peterson, T. L. Hughes, and G. Bahl, Robust temporal pumping in a magneto-mechanical topological insulator, Nature Communications 11, 974 (2020).
  • Xia et al. [2021] Y. Xia, E. Riva, M. I. N. Rosa, G. Cazzulani, A. Erturk, F. Braghin, and M. Ruzzene, Experimental observation of temporal pumping in electromechanical waveguides, Phys. Rev. Lett. 126, 095501 (2021).
  • Niu and Thouless [1984] Q. Niu and D. J. Thouless, Quantised adiabatic charge transport in the presence of substrate disorder and many-body interaction, Journal of Physics A: Mathematical and General 17, 2453 (1984).
  • Kuno and Hatsugai [2020] Y. Kuno and Y. Hatsugai, Interaction-induced topological charge pump, Phys. Rev. Res. 2, 042024(R) (2020).
  • Viebahn et al. [2023] K. Viebahn, A.-S. Walter, E. Bertok, Z. Zhu, M. Gächter, A. A. Aligia, F. Heidrich-Meisner, and T. Esslinger, Interaction-induced charge pumping in a topological many-body system (2023), arXiv:2308.03756 [cond-mat.quant-gas] .
  • Tuloup et al. [2023] T. Tuloup, R. W. Bomantara, and J. Gong, Breakdown of quantization in nonlinear thouless pumping (2023), in press.
  • Walter et al. [2023] A.-S. Walter, Z. Zhu, M. Gächter, J. Minguzzi, S. Roschinski, K. Sandholzer, K. Viebahn, and T. Esslinger, Quantization and its breakdown in a Hubbard-Thouless pump, Nature Physics 10.1038/s41567-023-02145-w (2023).
  • Jürgensen et al. [2021] M. Jürgensen, S. Mukherjee, and M. C. Rechtsman, Quantized nonlinear thouless pumping, Nature 596, 63 (2021).
  • Jürgensen and Rechtsman [2022] M. Jürgensen and M. C. Rechtsman, Chern number governs soliton motion in nonlinear thouless pumps, Physical review letters 128, 113901 (2022).
  • Mostaan et al. [2022] N. Mostaan, F. Grusdt, and N. Goldman, Quantized topological pumping of solitons in nonlinear photonics and ultracold atomic mixtures, Nature Communications 13, 5997 (2022).
  • Fu et al. [2022] Q. Fu, P. Wang, Y. V. Kartashov, V. V. Konotop, and F. Ye, Nonlinear thouless pumping: Solitons and transport breakdown, Phys. Rev. Lett. 128, 154101 (2022).
  • Jürgensen et al. [2023] M. Jürgensen, S. Mukherjee, C. Jörg, and M. C. Rechtsman, Quantized fractional thouless pumping of solitons, Nature Physics 19, 420 (2023).
  • Ke et al. [2017] Y. Ke, X. Qin, Y. S. Kivshar, and C. Lee, Multiparticle wannier states and thouless pumping of interacting bosons, Phys. Rev. A 95, 063630 (2017).
  • Tangpanitanon et al. [2016] J. Tangpanitanon, V. M. Bastidas, S. Al-Assam, P. Roushan, D. Jaksch, and D. G. Angelakis, Topological pumping of photons in nonlinear resonator arrays, Phys. Rev. Lett. 117, 213603 (2016).
  • Bogoliubov [1947] N. N. Bogoliubov, On the theory of superfluidity, J.Phys.(USSR) 11, 23 (1947).
  • Bardyn et al. [2016] C.-E. Bardyn, T. Karzig, G. Refael, and T. C. H. Liew, Chiral bogoliubov excitations in nonlinear bosonic systems, Phys. Rev. B 93, 020502(R) (2016).
  • Di Liberto et al. [2016] M. Di Liberto, A. Hemmerich, and C. Morais Smith, Topological varma superfluid in optical lattices, Phys. Rev. Lett. 117, 163001 (2016).
  • Carusotto and Ciuti [2013] I. Carusotto and C. Ciuti, Quantum fluids of light, Rev. Mod. Phys. 85, 299 (2013).
  • Schneider et al. [2016] C. Schneider, K. Winkler, M. D. Fraser, M. Kamp, Y. Yamamoto, E. A. Ostrovskaya, and S. Höfling, Exciton-polariton trapping and potential landscape engineering, Reports on Progress in Physics 80, 016503 (2016).
  • Kohnle et al. [2011] V. Kohnle, Y. Léger, M. Wouters, M. Richard, M. T. Portella-Oberli, and B. Deveaud-Plédran, From single particle to superfluid excitations in a dissipative polariton gas, Phys. Rev. Lett. 106, 255302 (2011).
  • Zajac and Langbein [2015] J. M. Zajac and W. Langbein, Parametric scattering of microcavity polaritons into ghost branches, Phys. Rev. B 92, 165305 (2015).
  • Stepanov et al. [2019] P. Stepanov, I. Amelio, J.-G. Rousset, J. Bloch, A. Lemaître, A. Amo, A. Minguzzi, I. Carusotto, and M. Richard, Dispersion relation of the collective excitations in a resonantly driven polariton fluid, Nature Communications 10, 3869 (2019).
  • Claude et al. [2022] F. Claude, M. J. Jacquet, R. Usciati, I. Carusotto, E. Giacobino, A. Bramati, and Q. Glorieux, High-resolution coherent probe spectroscopy of a polariton quantum fluid, Phys. Rev. Lett. 129, 103601 (2022).
  • Harper [1955] P. G. Harper, Single band motion of conduction electrons in a uniform magnetic field, Proceedings of the Physical Society. Section A 68, 874 (1955).
  • Aubry and André [1980] S. Aubry and G. André, Analyticity breaking and anderson localization in incommensurate lattices, Ann. Israel Phys. Soc 3, 133 (1980).
  • Kraus and Zilberberg [2012] Y. E. Kraus and O. Zilberberg, Topological equivalence between the fibonacci quasicrystal and the harper model, Phys. Rev. Lett. 109, 116404 (2012).
  • Hofstadter [1976] D. R. Hofstadter, Energy levels and wave functions of bloch electrons in rational and irrational magnetic fields, Phys. Rev. B 14, 2239 (1976).
  • Peano et al. [2016] V. Peano, M. Houde, C. Brendel, F. Marquardt, and A. A. Clerk, Topological phase transitions and chiral inelastic transport induced by the squeezing of light, Nature Communications 7, 10779 (2016).
  • [39] The Chern numbers are obtained by integrating the Berry curvature of Bogoliubov bands, taking into account their intrinsic symplectic structure, see Ref. [38].
  • Frérot et al. [2023] I. Frérot, A. Vashisht, M. Morassi, A. Lemaître, S. Ravets, J. Bloch, A. Minguzzi, and M. Richard, Bogoliubov excitations driven by thermal lattice phonons in a quantum fluid of light, Phys. Rev. X 13, 041058 (2023).
  • Tesfaye and Eckardt [2024] I. Tesfaye and A. Eckardt, Quantum geometry of bosonic bogoliubov quasiparticles (2024), arXiv:2406.12981 [cond-mat.quant-gas] .
  • Villa et��al. [2024] G. Villa, J. del Pino, V. Dumont, G. Rastelli, M. Michałek, A. Eichler, and O. Zilberberg, Topological classification of driven-dissipative nonlinear systems (2024), arXiv:2406.16591 [cond-mat.mes-hall] .