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

Waveguide quantum optomechanics: parity-time phase transitions in ultrastrong coupling regime

Ivan Iorsh Department of Physics and Technology, ITMO University, St. Petersburg, 197101, Russia    Alexander Poshakinskiy Ioffe Institute, 194021, St. Petersburg, Russia    Alexander Poddubny Ioffe Institute, 194021, St. Petersburg, Russia Department of Physics and Technology, ITMO University, St. Petersburg, 197101, Russia
Abstract

We develop a rigorous theoretical framework for interaction-induced phenomena in the waveguide quantum electrodynamics (QED) driven by mechanical oscillations of the qubits. Specifically, we predict that the simplest set-up of two qubits, harmonically trapped over an optical waveguide, enables the ultrastrong coupling regime of the quantum optomechanical interaction. Moreover, the combination of the inherent open nature of the system and the strong optomechanical coupling leads to emerging parity-time (𝒫𝒯\mathcal{PT}) symmetry, quite unexpected for a purely quantum system without artificially engineered gain and loss. The 𝒫𝒯\mathcal{PT} phase transition drives long-living subradiant states, observable in the state-of-the-art waveguide QED setups.

Introduction. Waveguide quantum electrodynamics Roy et al. (2017); Chang et al. (2018), studying the interaction of quantum emitters with propagating photons, experiences now a rapid progress driven by novel quantum technologies. Man-made arrays of cold atoms Corzo et al. (2019) and superconducting qubits van Loo et al. (2013); Mirhosseini et al. (2019) allow one to explore novel topological physics Kim et al. (2020); Perczel et al. (2020), collective super-radiance and sub-radiance  Ke et al. (2019); Kornovan et al. (2019); Zhang et al. (2020) with unprecedented tunability and precision and can be used to build future quantum networks Kimble (2008). However, the photon-qubit interaction in the waveguides is intrinsically weaker than in the cavity setup, since light is delocalized in space. It is not clear whether the fundamentally interesting regime of quantum phase transitions in ultra-strong coupling regime Kockum et al. (2019a) can be demonstrated in waveguide QED.

In this Letter we predict that mechanical motion of the qubits coupled to the waveguide opens up a new route to ultrastrong coupling at the quantum level for experimentally accessible parameters. Despite recent theoretical advances Manzoni et al. (2017); Mazza et al. (2020); Sánchez-Burillo et al. (2020) and experimental observations of optomechanical nonreciprocity with superconducting qubits Peterson et al. (2017), the field of waveguide quantum optomechanics still remains practically unchartered. Here we develop a rigorous theoretical framework to describe light-mediated coupling between mechanical oscillations in the array of qubits, harmonically trapped over an optical waveguide. Our calculations reveal an unexpectedly rich behavior of coupled light, qubits and vibrations. Even in the simplest case of just two qubits

Refer to caption
Figure 1: (Upper panel) Geometry of the problem: two oscillating qubits, trapped in the parabolic potentials, and coupled to the waveguide. (Lower panel) Phase diagram showing by color the radiative decay rates missingImE0±-\mathop{\mathrm{missing}}{Im}\nolimits E_{0}^{\pm} of the (anti)symmetric superposition of the photonic qubit excitations in the lowest vibrational state as function of interatomic distance and optomechanical coupling strength calculated from Hamiltonian (6). Red color indicates small decay rates, with color saturation proportional to the lifetime of most long-lived state 1/min(missingImE0+,missingImE0)-1/\text{min}(\mathop{\mathrm{missing}}{Im}\nolimits E_{0}^{+},\mathop{\mathrm{missing}}{Im}\nolimits E_{0}^{-}). Blue color encodes the difference between the decay rates of the two lowest vibrational states |missingIm(E1+E0+)||\mathop{\mathrm{missing}}{Im}\nolimits(E_{1}^{+}-E_{0}^{+})| with higher saturation corresponding to smaller difference realized in the vicinity of unbroken 𝒫𝒯\mathcal{PT} symmetry (solid line).

the system is described by a celebrated quantum Rabi model Casanova et al. (2018), which, in stark contrast to the usual situation, is inherently non-Hermitian due to the possibility of radiative decay into the waveguide. We examine the phase diagram depending on the strength of light-matter and optomechanical couplings and find the phases very sensitive to the inter-qubit spacings, controlling the radiative decay. Namely, when the spacing is equal to the quarter of light wavelength at the qubit resonance λ/4\lambda/4, the system demonstrates an emergence of parity-time (𝒫𝒯\mathcal{PT}) symmetry Feng et al. (2017). The 𝒫𝒯\mathcal{PT} symmetry breaking drives formation of coupled modes of light and vibrations with suppressed radiative decay, that manifest themselves as narrow resonances of single-photon transmission. A very different scenario is observed in case of Dicke superradiance, when the spacing is an integer multiple of λ/2\lambda/2. When the qubits are still, the spectrum has one subradiant and one subradiant mode, while an increase of optomechanical coupling brightens the subradiant mode and enables high-quality resonances in optical spectra. Both of these scenarios are summarized at the lower panel of Fig. 1. Our results demonstrate that the optomechanical coupling opens a new degree of freedom to engineer radiative lifetimes, spatial distributions and quantum correlations in waveguide quantum electrodynamics.

Optomechanical Quantum Rabi Model. We consider an array of qubits schematically shown in Fig. 1. Each qubit is trapped in individual harmonic optical trap which is placed in the vicinity of a single-mode optical fiber. The qubit can absorb or emit a waveguide photon, and the radiative relaxation to the far field outside the waveguide is suppressed. The Hamiltonian of the system can be written as

H^=kωkckck+jωxσj+σj+jΩajaj+H^int,\displaystyle\hat{H}=\sum_{k}\omega_{k}c_{k}^{\dagger}c_{k}+\sum_{j}\omega_{x}\sigma_{j}^{+}\sigma_{j}+\sum_{j}\Omega a_{j}^{\dagger}a_{j}+\hat{H}_{\rm int}, (1)

where ωk=v|k|\omega_{k}=v|k| is the dispersion of the waveguide modes, which is assumed to be linear, ωx\omega_{x} is the qubit resonance frequency, and Ω\Omega is the optical trap phonon energy. The interaction Hamiltonian is given by

H^int=gk,j[σjckeik[zj+u0(aj+aj)]+H.c.],\displaystyle\hat{H}_{\rm int}=g\sum_{k,j}\left[\sigma_{j}^{\dagger}c_{k}^{\vphantom{\dagger}}{\rm e}^{ik[z_{j}+u_{0}(a_{j}+a_{j}^{\dagger})]}+\mathrm{H.c.}\right], (2)

where gg is the Rabi splitting, zjz_{j} is the position of the jj-th optical trap, and u0=/(2MΩ)u_{0}=\sqrt{\hbar/(2M\Omega)} is the quantum of the mechanical motion, where MM is the mass of the qubit.

We integrate out the waveguide degrees of freedom Shi et al. (2011) to obtain the effective Hamiltonian up to the second order of the qubit-photon coupling gg:

H^eff=jωxσj+σj+jΩajaj\displaystyle\hat{H}_{\rm eff}=\sum_{j}\omega_{x}\sigma_{j}^{+}\sigma_{j}+\sum_{j}\Omega a_{j}^{\dagger}a_{j}
\displaystyle- iΓ0jlσj+σleiq|zjzl|eiqu0sign(zjzl)(aj+ajalal),\displaystyle i\Gamma_{0}\sum_{jl}\sigma_{j}^{+}\sigma_{l}e^{iq|z_{j}-z_{l}|}e^{iqu_{0}\,\text{sign}(z_{j}-z_{l})\,(a_{j}+a_{j}^{\dagger}-a_{l}-a_{l}^{\dagger})}, (3)

where Γ0=g2/v\Gamma_{0}=g^{2}/v is the radiative decay rate of a single qubit and we used the Markov approximation neglecting the frequency dispersion in the phase factor, kqωx/vk\approx q\equiv\omega_{x}/v. Hamiltonian (3) can be also obtained from the conventional waveguide-QED Hamiltonian Zhang and Mølmer (2019); Ke et al. (2019) where the coordinates of static qubits zjz_{j} are replaced with the corresponding dynamical value zj+u0(aj+aj)z_{j}+u_{0}(a_{j}+a_{j}^{\dagger}) and u0|zjzl|u_{0}\lesssim|z_{j}-z_{l}| is supposed.

In this work we restrict ourselves to a specific case of only two qubits and only a single photonic excitation in the array. Then, the qubit subspace of the full Hilbert space is a span of only two states |L|L\rangle and |R|R\rangle corresponding to the left/right qubit being in the excited state. We introduce a new rotated basis comprising a symmetric and antisymmetric superposition of the excitations at the individual qubits: |±=12(|L±|R)|\pm\rangle=\frac{1}{\sqrt{2}}(|L\rangle\pm|R\rangle). Moreover, we perform the unitary transformation of the phononic degrees of freedom: x^s(d)=(x^2±x^1)/2\hat{x}_{s(d)}=(\hat{x}_{2}\pm\hat{x}_{1})/\sqrt{2}, where x^i=ai+ai\hat{x}_{i}=a_{i}+a_{i}^{\dagger}. In the new basis the effective Hamiltonian is written as

H^eff=Ωasas+ΩadadiΓ0(1+σzeiφeiηx^d),\displaystyle\hat{H}_{\rm eff}=\Omega a_{s}^{\dagger}a_{s}^{\vphantom{\dagger}}+\Omega a_{d}^{\dagger}a_{d}^{\vphantom{\dagger}}-i\Gamma_{0}\left(1+\sigma_{z}e^{i{\varphi}}e^{i\eta\hat{x}_{d}}\right), (4)

where φ=q|z1z2|{\varphi}=q|z_{1}-z_{2}| and η=2qu0\eta=2qu_{0}. We note, that the symmetric vibrations are decoupled from the rest part of the Hamiltonian since they do not affect the distance between the qubits. Thus, the centre of mass of the qubits oscillates freely and the number of ss phonons is a good quantum number. Moreover, the |+|+\rangle and ||-\rangle states are also decoupled and described by the separate Hamiltonians:

H^eff±=ΩadadiΓ0(1±eiφeiηx^d).\displaystyle\hat{H}_{\rm eff}^{\pm}=\Omega a_{d}^{\dagger}a_{d}^{\vphantom{{\dagger}}}-i\Gamma_{0}\left(1\pm e^{i{\varphi}}e^{i\eta\hat{x}_{d}}\right). (5)

Namely, when η=0,φ=π\eta=0,{\varphi}=\pi the antisymmetric state corresponds to a dark, subradiant state, and the symmetric one to a superradiant state.

The spectrum of the obtained Hamiltonians can be found most conveniently by using the real-space representation of the mechanical motion. Namely, introducing the coordinate xdx_{d} measured in units of 2u0\sqrt{2}u_{0} we obtain the following representation of the effective Hamiltonian

H^eff±=Ω2(d2dxd2+xd21)iΓ0iΓ0eiφeiηxd.\displaystyle\hat{H}_{\rm eff}^{\pm}=\frac{\Omega}{2}\left(-\frac{{\rm d}^{2}}{{\rm d}x_{d}^{2}}+x_{d}^{2}-1\right)-i\Gamma_{0}\mp i\Gamma_{0}e^{i{\varphi}}e^{i\eta x_{d}}.~{} (6)

The obtained Hamiltonian allows for a simple numerical solution with the finite difference schemes.

Before we proceed further, it is important to discuss the range of the experimentally accessible parameters of the model. The relative strength of the quantum optomechanical coupling is defined by the ratio of the length scale of the mechanical atomic movement and the wavelength of the photon in the wavequide, η=4πu0/λ\eta=4\pi u_{0}/\lambda. The length scale of the atomic movement can be roughly estimated via the de Broglie wavelength u0</pthu_{0}<\hbar/p_{\rm th}, where the thermal momentum pth=3MkBTp_{\rm th}=\sqrt{3Mk_{B}T}. For the lithium atoms and the resonant wavelength approximately 700 nm the value of η=1\eta=1 is achieved at T=640T=640 nK, which is a temperature which has been achieved in recent cold atom experiments (see the review Anglin and Ketterle (2002) and references within). The corresponding phonon energy is then approximately 2.42.4 kHz. The radiative decay rate Γ0\Gamma_{0} can be flexibly tuned in a wide range of frequencies from zero to the GHz. Thus, the range of Γ0/Ω,η1\Gamma_{0}/\Omega,\eta\sim 1 is achievable within the modern experimental techniques.

Importantly, the Hamiltonians (5) and (6) have strong similarity with the quantum Rabi model Casanova et al. (2018). Indeed, the equidistant harmonic oscillator eigenstates with energy separation Ω\Omega play the role of the photon number states in the Rabi model. The third term in Eq. (6) mixes the different eigenstates similar to the atom-light interaction term in the Rabi model, and the decay rate Γ0\Gamma_{0} plays the role of the effective Rabi frequency. Since Ω\Omega is small, the ratio Γ0/Ω\Gamma_{0}/\Omega can be arbitrary large and it is possible to access the so-called ultrastrong coupling regime Kockum et al. (2019b). In what follows, the energy will be normalized to the phonon energy Ω\Omega.

Refer to caption
Figure 2: Optomechanical phase diagram for two λ/4\lambda/4-spaced qubits . The green and blue lines show the 𝒫𝒯\mathcal{PT}-phase boundaries for the first and second pair of the lowest symmetric eigenstates, respectively. The white line separates of the 𝒫𝒯\mathcal{PT}-preserved and 𝒫𝒯\mathcal{PT}-broken phase. The colorbar corresponds to max(missingImE)/Γ0\mathrm{max}(\mathop{\mathrm{missing}}{Im}\nolimits E)/\Gamma_{0}.

In stark contrast to the usual Rabi model Casanova et al. (2018), the Hamiltonians (5) and (6) are inherently non-Hermitian since the photons can escape into the waveguide. The strength of the radiative decay is controlled by the distance between the qubits determining the phase φ\varphi. We now set φ=π/2{\varphi}=\pi/2, which means that the qubits are spaced by the quarter of light wavelength at the resonance frequency. In this case, up to the constant decay Γ0\Gamma_{0}, the Hamiltonians in Eq. (6) become parity-time (𝒫𝒯\mathcal{PT}) symmetric. Indeed, the Hamiltonians Heff±+iΓ0H^{\rm\pm}_{\rm eff}+i\Gamma_{0} are invariant under the simultaneous action of the parity (xxx\rightarrow-x) and time reversal (complex conjugation) operators.

𝒫𝒯\mathcal{PT} phase transition. The parity-time symmetric Hamiltonians can either have real eigenvalues (unbroken-𝒫𝒯\mathcal{PT} phase) or pairs of complex conjugated eigenvalues (broken-𝒫𝒯\mathcal{PT} phase) Feng et al. (2017). The 𝒫𝒯\mathcal{PT} phase transition occurs at a certain value of coupling strength. In Fig. 2 we plot the numerically calculated 𝒫𝒯\mathcal{PT} phase diagram for the Hamiltonians (6) depending on the strength of the light-matter coupling Γ0\Gamma_{0} and the optomechanical coupling strength η\eta. The white solid line separates the 𝒫𝒯\mathcal{PT}-preserved phase (black area) where all the eigenvalues are real and the 𝒫𝒯\mathcal{PT}-broken phase, when at least one of the eigenenergies becomes complex. The colorbar shows the ratio of the largest imaginary part to Γ0\Gamma_{0}. The green and blue lines correspond to the 𝒫𝒯\mathcal{PT}-phase boundary for the first and second pair of the lowest symmetric eigenstates, respectively. The phase diagram demonstrates an existence of 𝒫𝒯\mathcal{PT}-unbroken phase for arbitrary large η\eta when Γ0<1/2\Gamma_{0}<1/2. As Γ0\Gamma_{0} increases, the threshold η\eta for the 𝒫𝒯\mathcal{PT}-phase transition decreases slowly.

Refer to caption
Figure 3: (a) Evolution of the real parts of the eigenergies of Heff+H_{\rm eff}^{+} (solid green lines) and HeffH_{\rm eff}^{-} (dashed blue lines). The color map shows the spectrum of the transmission coefficient with absorption of a single dd phonon |S0,1(ω)|2|S^{\rightarrow}_{0,1}(\omega)|^{2} expressed via Eq. (7). (b,c) Evolution of the imaginary part of eigenergy (b) and matrix element xd\langle x_{d}\rangle (c) for the first four eigenstates of Heff+H_{\rm eff}^{+} vs optomechanical coupling η\eta. Calculation has been performed for Γ0/Ω=2.\Gamma_{0}/\Omega=2.

We will now analyze the 𝒫𝒯\mathcal{PT} phase transition in more detail. To this end, we fix the light-matter coupling strength to Γ0=2Ω\Gamma_{0}=2\Omega and plot in Fig. 3(a) the real parts of the eigenergies E±E^{\pm} with green and blue lines, respectively. The spectrum of the imaginary parts of the four lowest eigenenergies of Heff+H_{\rm eff}^{+} is shown in Fig. 3(b). Naturally, at η=0\eta=0 the spectrum is real and it is essentially a pair of the harmonic oscillator ladders with the energies ±Γ0+nΩ\pm\Gamma_{0}+n\Omega, with n=0,1,2n=0,1,2\ldots. As η\eta increases, the energies of the first two pairs of eigenstates of Heff+H_{\rm eff}^{+} (solid green lines) approach each other, and collapse at the corresponding threshold values of η\eta close to 0.50.5. This is the point of 𝒫𝒯\mathcal{PT} phase transition, after which the energies acquire imaginary parts and form complex conjugate pairs, see Fig. 3(b). At even higher values η\eta, the second phase transition occurs, where the imaginary parts of the eigenenergies collapse. Consequently, there exists a certain critical value of η\eta, where the maximum deviation of the eigenenergy imaginary part from Γ0-\Gamma_{0} is achieved. At this point, one of the states, is characterized by the suppressed decay rate. We thus observe the optomechanically induced darkening of the qubit state.

The breaking of the 𝒫𝒯\mathcal{PT} symmetry can be directly observed in the dependence of the average inter-qubit displacement xd=𝑑xdψn,+xdψn,+\langle x_{d}\rangle=\int dx_{d}\psi_{n,+}^{*}x_{d}\psi_{n,+} on η\eta shown in Fig. 3(c). For small η\eta, when the eigenstates are 𝒫𝒯\mathcal{PT}-symmetric, the average value of the 𝒫𝒯\mathcal{PT}-odd quantity xd\langle x_{d}\rangle must be zero, and thus the qubits are placed at the equilibrium positions. At the 𝒫𝒯\mathcal{PT} symmetry breaking point, xd\langle{x_{d}}\rangle changes discontinuously, and a pair of eigenstates connected by 𝒫𝒯\mathcal{PT} symmetry operation acquire opposite nonzero values of xd\langle{x_{d}}\rangle. The qubit pair becomes ’stretched’ for the darker and ’squeezed’ for brighter state. The effect of squeezing or stretching has been considered in the qubit chains in the different regime when both coordinates and qubit polarizations were treated classically Chang et al. (2013)

We will now demonstrate that the subradiant modes, resulting from quantum optomechanical darkening, are manifested as sharp resonances in the spectra of single photon scattering. To calculate the latter, we exploit the Green’s function, described in the Supplementary Material in detail. The coefficients of single photon forward (backward) scattering accompanied by emission of NsN_{s} ss-phonons and NdN_{d} dd-phonons read

SNs,Nd()(ω)=iΓ0Ns=0Nd=0p=±p112𝑑xs𝑑xs𝑑xd𝑑xd\displaystyle S^{\rightarrow(\hookleftarrow)}_{N_{s},N_{d}}(\omega)=-i\Gamma_{0}\sum_{N_{s}^{\prime}=0}^{\infty}\sum_{N_{d}^{\prime}=0}^{\infty}\sum_{p=\pm}p^{\frac{1\mp 1}{2}}\int dx_{s}dx^{\prime}_{s}dx_{d}dx^{\prime}_{d}
×eiη2(xsxs)[cos(η2(xdxd))+pcos(φ+η2(xd+xd))]ωωxENd±NsΩ\displaystyle\times\frac{e^{i\frac{\eta}{2}(x_{s}\mp x_{s}^{\prime})}\left[\cos(\frac{\eta}{2}(x_{d}^{\prime}-x_{d}))+p\cos({\varphi}+\frac{\eta}{2}(x_{d}+x_{d}^{\prime}))\right]}{\omega-\omega_{x}-E_{N_{d}^{\prime}}^{\pm}-N_{s}^{\prime}\Omega}
×φNs(xs)φNs(xs)φNs(xs)φ0(xs)\displaystyle\times{\varphi}_{N_{s}}(x_{s}){\varphi}_{N^{\prime}_{s}}(x_{s}){\varphi}_{N^{\prime}_{s}}({x^{\prime}_{s}}){\varphi}_{0}(x^{\prime}_{s})
×φNd(xd)ψNd±(xd)ψNd±(xd)φ0(xd),\displaystyle\times{\varphi}_{N_{d}}(x_{d})\psi^{\pm}_{N^{\prime}_{d}}(x_{d})\psi^{\pm}_{N^{\prime}_{d}}(x^{\prime}_{d}){\varphi}_{0}(x^{\prime}_{d}), (7)

where φN(x){\varphi}_{N}(x) is the eigenfunction of the free harmonic oscillator corresponding to NN phonons and summation is taken over all possible intermediate states, characterised by the photonic mode parity pp and the number of mechanical excitation, NsN_{s}^{\prime} and NdN_{d}^{\prime}. The energy of the intermediate states is ENd±+NsΩE_{N_{d}^{\prime}}^{\pm}+N_{s}^{\prime}\Omega, where ENd±E_{N_{d}^{\prime}}^{\pm} is the eigenenergy of Heff±H_{\rm eff}^{\pm} and ψNd±(x)\psi^{\pm}_{N_{d}^{\prime}}(x) is the corresponding wave function normalized according to [ψNd±(x)]2𝑑x=1\int[\psi^{\pm}_{N_{d}^{\prime}}(x)]^{2}\,dx=1. Note that that while the phonon reflection probability is simply |SNs,Nd(ω)|2|S^{\hookleftarrow}_{N_{s},N_{d}}(\omega)|^{2}, that of photon transmission reads |δNs,0δNd,0+SNs,Nd(ω)|2|\delta_{N_{s},0}\delta_{N_{d},0}+S^{\rightarrow}_{N_{s},N_{d}}(\omega)|^{2}.

Refer to caption
Figure 4: Evolution of the reflection spectrum |S0,0(ω)|2|S^{\hookleftarrow}_{0,0}(\omega)|^{2} vs optomechanical coupling calculated for φ=π{\varphi}=\pi and Γ0/Ω=2\Gamma_{0}/\Omega=2. Dashed blue lines label the energies of the two lowest eigenstates of HeffH_{\rm eff}^{-}.

The darkening of the dd phonon eigenstates in the 𝒫𝒯\mathcal{PT}-broken regime discussed above should lead to the resonant enhancement of the reflection and transmission as the imaginary part of the denominator in Eq. (7) goes to zero for the corresponding intermediate eigenstates. In order to demonstrate this effect, we show in Fig. 3(b) by color the absolute squared transmission coefficient with emission of 1 dd phonon and zero ss phonons as a function of detuning δ=ωωx\delta=\omega-\omega_{x} and η\eta. The calculation demonstrates enhancement of the transmission in the vicinity of the 𝒫𝒯\mathcal{PT} symmetry breaking transition for the first two eigenstates of Heff+H_{\rm eff}^{+} (red lines), where one of the states becomes almost dark.

In the case of φ=π{\varphi}=\pi (two λ/2\lambda/2-spaced qubits), the effect of optomechanical coupling is quite the opposite. In the absence of optomechanical coupling the eigenmodes of Heff±H_{\rm eff}^{\pm} have the same real energy, however those of HeffH_{\rm eff}^{-} are dark mode with zero decay rate, while those of Heff+H_{\rm eff}^{+} are characterized by the decay rate 2Γ02\Gamma_{0}. The optomechanical coupling leads to the brightening of the dark modes and to the emergence of the sharp resonances in the reflection/transmission spectra. We illustrate this mechanism in Fig. 4. Namely we plot the evolution of the coherent reflectance spectrum |S0,0(ω)|2|S^{\hookleftarrow}_{0,0}(\omega)|^{2} with η\eta. We observe the emergence of the sharp dips in the reflection coefficient at the frequencies corresponding to the eigenergies (labelled with dashed blue lines) as η\eta increases which corresponds to the brightening of the dark modes due to the optomechanical interaction.

In summary, we have demonstrated that the ultrastrong regime of optomechanical coupling for array of qubits near a waveguide can be used to engineer optomechanical 𝒫𝒯\mathcal{PT} symmetry, achieve strongly subradiant states and control single-photon transmission spectra. We believe that the full potential of this setup is far from being explored and a plethora of new physical phenomena can be expected. For instance, the full strength of collective light-matter coupling will be uncovered only for larger number of qubits N>2N>2, where the collective subradiant states start playing role Albrecht et al. (2019). Another degree of freedom will open up when more than one photon excitation is taken into account Zhang and Mølmer (2019); Poshakinskiy et al. (2020). The resulting combination of long coherence of mechanical vibrations and strong quantum nonlinearities can become beneficial for emerging quantum technologies.

Acknowledgements.
I.I. acknowledges the support from the Russian Science Foundation (project 20-12-00224). A.N.P. was supported by the Russian President Grant MD-243.2020.2. A.V.P was supported by the Russian President Grants MD-243.2020.2 and MK-599.2019.2 and the Foundation “BASIS”.

References

*

Appendix A Supplemental Material: calculation of the Raman scattering amplitude

Refer to caption
Figure 5: Diagrammatic representation of the photon reflection or transmission from a qubit array: (a) coherent process, (b) Raman process with one phonon emission, (c) Raman process with emission of two phonons.

The amplitude of the single photon transition or reflection from the system is most conveniently calculated using the Green’s function technique. The diagrams in Fig. 5 show the diagrams corresponding the coherent transmission/reflection [panel (a)] and the Raman scattering processes when one or two vibrational quanta are excited [panels (b) and (c)]. The dot is the photon-phonon-qubit interaction vertex representing Hamiltonian H^int\hat{H}_{\rm int}, Eq. (2) and solid line is the Green’s function Gij(ω){G}_{ij}(\omega) that describes light-mediated propagation of a single exciton from jj-th to ii-th qubit dressed by interaction with qubit vibrations. Such Green’s function is matrix N×NN\times N of operators acting on vibrational states of all NN qubits of the array, that is defined as 𝑮(ω)=[ω𝑯eff]1{\bm{G}}(\omega)=[\omega-{\bm{H}}_{\rm eff}]^{-1}, where 𝑯eff{\bm{H}}_{\rm eff} is the matrix of operator Eq. (3) on the subspace of states with a single qubit excitation,

Heff,ij=ωxδij+kΩakakiΓ0eiq|zizj|eiqu0sign(zizj)[(ai+ai)(aj+aj)].\displaystyle H_{{\rm eff},ij}=\omega_{x}\delta_{ij}+\sum_{k}\Omega a_{k}^{\dagger}a_{k}-{\rm i}\Gamma_{0}{\rm e}^{{\rm i}q|z_{i}-z_{j}|}\,{\rm e}^{{\rm i}qu_{0}\,\text{sign}(z_{i}-z_{j})\,[(a_{i}+a_{i}^{\dagger})-(a_{j}+a_{j}^{\dagger})]}\,. (8)

The amplitude of photon transmission and reflection then reads

SΦ(ω)=δΦ,0iΓ0ijΦ|eik[zi+u0(ai+ai)]Gijeik[zj+u0(aj+aj)]|0,\displaystyle S^{\rightarrow}_{\Phi}(\omega)=\delta_{\Phi,0}-{\rm i}\Gamma_{0}\sum_{ij}\langle\Phi|{\rm e}^{-{\rm i}k[z_{i}+u_{0}(a_{i}+a_{i}^{\dagger})]}G_{ij}{\rm e}^{{\rm i}k[z_{j}+u_{0}(a_{j}+a_{j}^{\dagger})]}|0\rangle\,, (9)
SΦ(ω)=iΓ0ijΦ|eik[zi+u0(ai+ai)]Gijeik[zj+u0(aj+aj)]|0,\displaystyle S^{\hookleftarrow}_{\Phi}(\omega)=-{\rm i}\Gamma_{0}\sum_{ij}\langle\Phi|{\rm e}^{{\rm i}k[z_{i}+u_{0}(a_{i}+a_{i}^{\dagger})]}G_{ij}{\rm e}^{{\rm i}k[z_{j}+u_{0}(a_{j}+a_{j}^{\dagger})]}|0\rangle\>, (10)

where |0|0\rangle is initial vibrational state, which is supposed to be the ground state, and |Φ|\Phi\rangle is the final vibrational state, where some of the qubit vibrations might be excited.

For numerical calculation a coordinate representation of qubit vibrations is more convenient rather than the Fock-state representation. In such representation, the Green’s function Gij(ω;𝒙,𝒙)G_{ij}(\omega;\bm{x},\bm{x}^{\prime}) satisfies the the equation

[ωωxΩ2k(2u02d2dxk2+xk22u021)]Gij(ω;𝒙,𝒙)\displaystyle\left[\omega-\omega_{x}-\frac{\Omega}{2}\sum_{k}\left(-2u_{0}^{2}\frac{d^{2}}{dx_{k}^{2}}+\frac{x_{k}^{2}}{2u_{0}^{2}}-1\right)\right]G_{ij}(\omega;\bm{x},\bm{x}^{\prime}) (11)
+iΓ0leik|zizl|eiksign(zizl)(xixl)Glj(ω;𝒙,𝒙)=δijδ(𝒙𝒙).\displaystyle+{\rm i}\Gamma_{0}\sum_{l}{\rm e}^{{\rm i}k|z_{i}-z_{l}|}{\rm e}^{{\rm i}k\,\text{sign}(z_{i}-z_{l})\,(x_{i}-x_{l})}G_{lj}(\omega;\bm{x},\bm{x}^{\prime})=\delta_{ij}\delta(\bm{x}-\bm{x}^{\prime})\,.

where 𝒙=(x1,x2,xN)\bm{x}=(x_{1},x_{2},\ldots x_{N}) is the vector of coordinates of all qubits. The Green’s function is readily expressed as

Gij(ω;𝒙,𝒙)=νψi(ν)(𝒙)ψj(ν)(𝒙)ωωxϵ(ν)\displaystyle G_{ij}(\omega;\bm{x},\bm{x}^{\prime})=\sum_{\nu}\frac{\psi^{(\nu)}_{i}(\bm{x})\psi^{(\nu)}_{j}(\bm{x}^{\prime})}{\omega-\omega_{x}-\epsilon^{(\nu)}} (12)

where ψi(ν)(x)\psi^{(\nu)}_{i}(x) is the set of eigensolutions of the equation system

Ω2k(2u02d2dxk2+xk22u021)ψi(ν)(𝒙)\displaystyle\frac{\Omega}{2}\sum_{k}\left(-2u_{0}^{2}\frac{d^{2}}{dx_{k}^{2}}+\frac{x_{k}^{2}}{2u_{0}^{2}}-1\right)\psi^{(\nu)}_{i}(\bm{x})
iΓ0keiq|zizk|eiqsign(zizk)(xixk)ψk(ν)(𝒙)=ϵ(ν)ψi(ν)(𝒙)\displaystyle-{\rm i}\Gamma_{0}\sum_{k}{\rm e}^{{\rm i}q|z_{i}-z_{k}|}{\rm e}^{{\rm i}q\,\text{sign}(z_{i}-z_{k})\,(x_{i}-x_{k})}\psi^{(\nu)}_{k}(\bm{x})=\epsilon^{(\nu)}\psi^{(\nu)}_{i}(\bm{x}) (13)

that satisfies the orthogonality condition i𝑑𝒙ψi(ν)(𝒙)ψi(μ)(𝒙)=δνμ\sum_{i}\int d\bm{x}\,\psi^{(\nu)}_{i}(\bm{x})\psi^{(\mu)}_{i}(\bm{x})=\delta_{\nu\mu}. Note that complex conjugation is not involved in this condition, since the effective Hamiltonian is not a Hermitian matrix but a symmetric one. The transmission and reflection coefficients in the coordinate representation read

Sφ(ω)=δφ,φ0iΓ0ijφ(𝒙)eiq(zi+xi)Gij(ω;𝒙,𝒙)eiq(zj+xj)φ0(𝒙)𝑑𝒙𝑑𝒙,\displaystyle S^{\rightarrow}_{{\varphi}}(\omega)=\delta_{{\varphi},{\varphi}_{0}}-{\rm i}\Gamma_{0}\sum_{ij}\int{\varphi}(\bm{x}){\rm e}^{-{\rm i}q(z_{i}+x_{i})}G_{ij}(\omega;\bm{x},\bm{x}^{\prime})\,{\rm e}^{{\rm i}q(z_{j}+x^{\prime}_{j})}{\varphi}_{0}(\bm{x}^{\prime})\,d\bm{x}d\bm{x}^{\prime}\,, (14)
Sφ(ω)=iΓ0ijφ(𝒙)eiq(zi+xi)Gij(ω;𝒙,𝒙)eiq(zj+xj)φ0(𝒙)𝑑𝒙𝑑𝒙,\displaystyle S^{\hookleftarrow}_{{\varphi}}(\omega)=-{\rm i}\Gamma_{0}\sum_{ij}\int{\varphi}(\bm{x}){\rm e}^{{\rm i}q(z_{i}+x_{i})}G_{ij}(\omega;\bm{x},\bm{x}^{\prime})\,{\rm e}^{{\rm i}q(z_{j}+x^{\prime}_{j})}{\varphi}_{0}(\bm{x}^{\prime})\,d\bm{x}d\bm{x}^{\prime}\>, (15)

where φ0(𝒙){\varphi}_{0}(\bm{x}) and φ(𝒙){\varphi}(\bm{x}) are the eigenfunctions of the harmonic oscillator corresponding to initial and final vibrational state.

Applying Eqs. (12), (14), (15) to the case of two-qubit array, as described in the main text, we obtain Eqs. (7)-(8) of the main tex.