Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Simulating topological phases with atom arrays in an optical waveguide

Open Access Open Access

Abstract

In this paper, we employ the atomic arrays in one-dimensional optical waveguides to simulate topological phases, where the waveguide is modeled as a one-dimensional infinitely long coupled cavity array. Under the Markov approximation, the coherent and dissipative coupling between atoms is established by eliminating waveguide modes. When the detuning between atoms and cavity fields lies in the band gap, the dynamics of the system is completely dominated by the coherent interaction. Under this condition, we designed three atomic arrays with different geometries and show that the topologically trivial and non-trivial phases of atomic arrays can be simulated. Furthermore, by introducing periodic atomic driving, the topological phase transition can be induced by adjusting the driving parameters. Finally, we investigate the effect of next-nearest neighbor interactions on topological state transfer and find that the next-nearest neighbor interactions break the degenerated bandgap state and establish a topological state transfer channel.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

With the rapid development of micro- and nano- processing technologies, the waveguide quantum electrodynamics (QED), in which the atoms are coupled to a one-dimensional (1D) propagation field, has become one of the important physical platform for realizing quantum information processing [16] and quantum simulation [711]. The optical modes coupling to the atoms near band edges can be well understood by mapping to the Jaynes-Cummings model [12,13]. This leads to many interesting phenomena and applications such as the generation of long-distance entanglement [14,15], correlated photon scattering [16,17], simulation of topological states [18,19].

Topological insulators, because of many interesting features such as robustness to local decoherence processes and potential applications in quantum information, has attracted a great of interests and attentions in quantum physics [2023]. The Su-Schrieffer-Heeger (SSH) model originally used to describe the transport properties of conducting polyacetylene [24] has also received increasing attention in recent years [4,18,19,2537]. As one of the simplest topological insulator models with its simple structure and abundant physical images, the SSH model has been implemented in several quantum systems including photonic crystals [29,30], spin-phonon crystals [18,19], cavity optomechanical system [3134], circuit QED [3,4,35] , trapped ions [36], plasmonic system [37] et al. Based on these platforms, a lot of interesting phenomena and potential applications have been done for example, topological phase transitions [3840], quantum state transfer [4143], non-Hermitian [44] et al. Recently, M. D’Angelis et al. proposed a faster and higher-fidelity state transfer scheme by considering next-nearest neighbor interactions, which significantly impove of the transfer speed compared to conventional adiabatic processes [45]. Furthermore, Wanjura et al. proposed a method using dynamic matrix topological property to predict and to describe the directional-amplifying transmission in driven-dissipative cavity arrays [46,47], which establishes a connection between directional transmission and topological insulator theory and greatly enhance the links between quantum physical phenomena and topological physics.

In this paper, we propose a scheme to simulate the SSH model by designing the positions of the atoms in the coupled cavity arrays. After eliminated the photon modes, the long-range coherent and correlated dissipative interactions between atoms can be established. When the detuning between atoms and cavity fields lies in the band gap, the dissipative coupling approaches zero, so the dynamics of the system is completely dominated by the coherent interaction. Under this condition, we designed three atomic arrays with different geometries. By considering the positions of atoms in coupled cavity arrays, we can achieve topologically trivial and non-trivial phases of atomic arrays. Moreover, the topological phase transition can be induced by tuning the drive parameters when we introduce periodic atomic drives. Finally, we shows that when photon-induced next-nearest-neighbor interactions were taken into account, the degenerate bandgap states are broken, and a topological state transfer channel is established.

2. Effective coherent and dissipative coupling in atomic array

We consider $2N_a$ two-level atoms coupling to a one-dimensional waveguide as shown in Fig. ( 1). Each atom interacts with the cavity through the local Jaynes-Cummings Hamiltonian with coupling strength $g$. The Hamiltonian can be written as

$$H=H_{w}+H_{s}+H_{I},$$
where
$$H_{w} =\sum\limits_{j={-}N}^{N}\omega _{a}a_{j}^{\dagger }a_{j}+J(a_{j+1}^{\dagger }a_{j}+h.c.),$$
$$H_{s} =\sum\limits_{i=1}^{2N_{a}}\omega _{o}\sigma _{i}^{\dagger }\sigma _{i},$$
$$H_{I} =\sum\limits_{j={-}N}^{N}\sum\limits_{i=1}^{2N_{a}}g(a_{j}^{\dagger }\sigma _{i}+\sigma _{i}^{\dagger }a_{j})\delta _{j,i}.$$
$H_w$ describes an optical waveguide modeled as an array consisting of $2N$ optical cavities, where $a_{j}$ is the annihilation operator of the optical cavity at $x_{j}$ with frequency $\omega _{a}$, $J$ refers to the coupling strength between adjacent cavities. $H_s$ denotes the energy of $2N_{a}$ atoms, where $\sigma _{i}$ is the transition operator of the $i$th atom from the excited state $|e\rangle$ to the ground state $\vert g\rangle$ at $x_{i}$ with the transition frequency $\omega _{o}$. Meanwhile, the number of atoms $2N_a$ is smaller than the cavity field. Although, we need enough atoms to simulate the topological phases in the next section, we consider that the number of atoms is still much smaller than the cavity. $H_{I}$ represents the interaction between the atoms and the cavity fields with coupling strength $g$. Considering $N \to \infty$, we can rewrite the Hamiltonian Eq. (2) in momentum space. By introducing Fourier transformation as $a_{j}=\frac {1}{\sqrt {N}} \sum\limits _{k}F_{jk}a_{k}$, where $F_{jk}=\exp (ikx_{j})$, and $\sum\limits _{j}F_{jk}^{\ast }F_{jl}=N\delta _{lk}$, then, the Hamiltonian $H_{w}$ becomes $H_{k}=\sum\limits _{k}\omega _{k}a_{k}^{\dagger }a_{k}$ with the dispersion relation $\omega _{k}=\omega _{a}+2J\cos (kx_0)$. Therefore, $H_{k}$ can be understood as an optical waveguide with center frequency $\omega _{a}$ and bandwidth $4J$. It is worth noting that the above dispersion relation is a result of strict Fourier transformation and does not include the actual photon band structure and effects such as directional emission [48]. Under the condition $kx_0<<1$, the dispersion relation can be expanded to linear or quadratic dispersion relation [49]. In the momentum space, $H_I$ becomes
$$H_{I}=\sum\limits_{j,k}\frac{g}{\sqrt{N}}(F_{j,k}^{{\ast} }a_{k}^{\dagger }\sigma _{j}+F_{j,k}a_{k}\sigma _{j}^{\dagger }).$$

 figure: Fig. 1.

Fig. 1. Schematic diagram of the $2N_{a}$ two-level atoms coupled into an optical waveguide. The coupling strength and distance between adjacent cavities are $J$ and $x_0$.

Download Full Size | PDF

Next, we consider the weak coupling or broadband limit $g/J\ll 1$. In this case, the photonic waveguide modes simply act as a collective reservoir of the atoms and can be eliminated by using the Born-Markov approximation, which can be guaranteed by the condition that the relaxation time of the environment is much faster than the evolution time scale of the states of the atomic array. The master equation for the reduced density operator of the atoms is as [48]

$$\dot{\rho}=\sum\limits_{j,i=1}^{2N_{a}}A_{ij}(\sigma _{j}\rho \sigma _{i}^{\dagger }-\sigma _{i}^{\dagger }\sigma _{j}\rho )+A_{ij}^{{\ast} }(\sigma _{i}\rho \sigma _{j}^{\dagger }-\rho \sigma _{j}^{\dagger }\sigma _{i}),$$
where
$$A_{ij}=\frac{4g^{2}e^{iK|x_{i}-x_{j}|}}{\sqrt{4J^{2}-(\delta +i\frac{\gamma _{a}}{2})^{2}}},$$
with $K=\pi -\arccos {[\frac {\delta +i\gamma _{a}/2}{2J}]}$, where $\delta =\omega _{o}-\omega _{a}$ is the atom-photon detuning and $\gamma _a$ is the loss of the cavity field. After rewriting Eq. (6) , we obtain
$$\dot{\rho}={-}i[H_{eff},\rho ]+\sum\limits_{j,i=1}^{2N_{a}}\Gamma _{i,j}(2\sigma _{j}\rho \sigma _{i}^{\dagger }- \sigma _{i}^{\dagger }\sigma _{j}\rho-\rho\sigma _{i}^{\dagger }\sigma _{j}),$$
with
$$H_{eff}=\sum\limits_{i,j}^{2N_{a}}U_{i,j}(\sigma _{i}^{\dagger }\sigma _{j}+\sigma _{j}^{\dagger }\sigma _{i}),$$
where $U_{i,j}=Im(A_{i,j})$, $\Gamma _{i,j}=Re(A_{i,j})$. It can be seen that the eliminated photon modes produce long-range coherent interaction $U_{i,j}$ and correlated dissipation $\Gamma _{i,j}$ between atoms. In Eq. (8), we have neglected the decay of the atoms. When the decoherence time of the atom is longer enough than the dynamics evolution time of the system, the atomic decay can be ignored. We will discuss the lifetime of atoms in section 6.

In Fig. (2), we plot the correlated decay rates $\Gamma _{i,j}$ and coherent interactions $U_{i,j}$ versus $\delta$. The distance between atoms $|x_{i}-x_{j}|=0, x_0, 2x_0, 3x_0$ corresponds to the blue ($U_0, \Gamma _0$), red ($U_1, \Gamma _1$), purple ($U_2, \Gamma _2$), and cyan ($U_3, \Gamma _3$) curves, respectively. We take $x_0=1$ in the numerical calculation. The presence of dissipative coupling is harmful for the topological study of the system, so it is necessary to suppress it. As for the atomic detuning range in the $-2J<\delta <2J$, the dissipative couplings cannot be zero simultaneously for different distance between atoms as shown in Fig. 2(b). When the atomic frequency lies in the bandgap region ($\delta >2J$ or $\delta < -2J$), it can be seen that the coherent and dissipative couplings exhibit the same asymptotic behavior and converge to zero with increasing or decreasing $\delta$ for different distance between atoms $|x_{i}-x_{j}|$. This asymptotic behavior offers us the possibility to consider only several-times coupling in terms of $|x_{i}-x_{j}|$, which is obviously different from [38,50] where the strength of the coupling between the atoms caused by the waveguide exhibits periodic oscillations with distance between the atoms. For the fixed detuning ($\delta =2.2J$), the coherent interactions $|U_{i,j}|$ gradually decrease with the atomic distance increases. Under the parameters in Fig. 2(a), for enough value $\delta$ ($\delta >2.8J$), $|U_{i,j}|$ with $|x_{i}-x_{j}|=3x_0$ is smaller than that with $|x_{i}-x_{j}|=0, x_0, 2x_0$ so that it can be ignored. Most importantly, $U_{i,j}$ and $\Gamma _{i,j}$ for $|x_i-x_j|>3x_0$ are all near zero under the group of parameters in Fig. 2, because of the complex $K$ making $A_{ij}$ exponentially decreasing with the increasing of $|x_i-x_j|$. Thus $U_{i,j}$ can be approximated as follows

$$U_{i,j}=U_{0}\delta _{i,j}+U_{1}\delta _{i,j+1}+U_{2}\delta _{i,j+2},$$
where $U_{0}$, $U_{1}$, $U_{2}$ represent the on-site, nearest-neighbor interaction ( the distance between cavities is $x_0$) and next-nearest-neighbor interaction ( the distance between cavities is $2x_0$) are induced by eliminating waveguides, respectively. Simultaneously, we see that the dissipative coupling $\Gamma _{i,j}$ is suppressed when $\delta >2.8J$ as shown in Fig. 2(b). Thus, for appropriate values of detuning, the dissipative coupling can be zero. Plugging Eq. (10) into Eq. (9), the effective Hamiltonian becomes
$$H_{eff}=\sum\limits_{j=1}^{2N_{a}}U_0\sigma_{j}^{\dagger }\sigma _{j}+U_{1}(\sigma _{j}^{\dagger }\sigma _{j+1}+h.c.)+U_{2}(\sigma _{j}^{\dagger }\sigma _{j+2}+h.c.).$$

 figure: Fig. 2.

Fig. 2. Correlated decay rates $\Gamma _{i, j}$ and coherent dipole-dipole interactions $U_{i, j}$ versus $\delta$ for different distance between atoms $|x_{i}-x_{j}|=0, x_0, 2x_0, 3x_0$ corresponds to the blue ($U_0, \Gamma _0$), red ($U_1, \Gamma _1$), purple ($U_2, \Gamma _2$), and cyan ($U_3, \Gamma _3$), respectively. The other parameters are $\gamma _{a}=0.25J$ and $g=0.3J$.

Download Full Size | PDF

Therefore, the dynamics of the system is governed by the Hamiltonian Eq. (11). Because the effective coupling strengths $U_i$ ($i=0,1,2$) depends on the distance between the atoms. In next section, we will design the special position of the atoms in the coupled cavity arrays to simulate the SSH model with different topological phases. Moreover, the on-site potential induced by photons is the same for every atom. This leads to a drift in the energy spectrum and does not affect the quantization of the zak phase as detailed in Appendix A. Therefore, in the following, we neglected this item.

3. Simulated SSH model by designing atomic position

In the previous section, we simulate the SSH model by changing the $2N_{a}$ atoms arrangement inside the waveguide with three different atom arrays. As shown in Fig. (3), we designed atomic arrays with three different geometries. For the A-Type atomic arrangement structure shown in Fig. 3(a), the $2N_a$ atoms are adjacent to each other in the coupled cavity arrays and the intra-cell atomic distance $x_0$ equals to the extra-cell atomic distance. The Hamiltonian of this structure can be written as Eq. (1). For the B-Type atomic arrangement structure in Fig. 3(b), each cell contains two atom-cavity sites, and an empty cavity connects the cells. So the intra-cell atomic distance is $x_0$, and the extra-cell atomic distance is $2x_0$. The interaction Hamiltonian can be written as

$$H_{Ib} =\sum\limits _{j={-}N}^{N }\ ^{\prime}\sum\limits_{i=1}^{2N_{a}}g(a_{j}^{{\dagger}}\sigma _{i}+\sigma _{i}^{\dagger }a_{j})\delta _{j,i},$$
where $\ ^{\prime }$ represents the mod $(j, 3)\ne 0$. As for the atomic array with the structure of Fig. 3(c), each cell contains three cavity fields in which the cavities at each end interacts with atoms. Thus, the intra-cell atomic distance is $2x_0$ and the extra-cell distance is $x_0$. For the C-Type atomic arrangement structure, the interaction Hamiltonian can be expressed as
$$H_{Ic} =\sum\limits_{j={-}N}^{N }\ ^{\prime\prime}\sum\limits_{i=1}^{2N_{a}}g(a_{j}^{\dagger }\sigma _{i}+\sigma _{i}^{\dagger }a_{j})\delta _{j,i},$$
where $\ ^{\prime \prime }$ represents the mod $(j+1,3)\ne 0$. It is worth noting here that we achieve two different geometries by setting the number of cavity fields within each cell. Due to the different number of cavity fields, which leads to different atomic distances within the cell, thus, the two geometries B-Type, C-Type are different. In other words, the different structures are not just resulted from artificial classification but stem from the difference at the two ends.

 figure: Fig. 3.

Fig. 3. (a) (b) (c) Three different arrangements of atomic array in coupled cavity arrays. We label the odd number of atoms as $A_{n}$ and the even number of atoms as $B_{n}$, $n=1,2,\ldots,N_{a}$.

Download Full Size | PDF

No matter what kinds of structures, the Fourier transformation for the Hamiltonian Eq. (3) is exactly applicable, and the method of eliminating the waveguide is the same as that achieving Eq. (11). The only difference is the position of the atoms, which is exhibited by the effective coupling $U_j (j=1,2)$ under $\delta >2.8J$. For convenience, here we label the odd number of atoms as $A_{n}$ (gray ball) and the even number of atoms as $B_{n}$ ( blue ball) , $n=1,2,\ldots,N_{a}$. Thus, the same subscript is in one cell. For A-Type atomic array, since the atoms are equally spaced, according to Eq. (11), there will be next-nearest neighbor interactions between the atoms. In terms of $A_n$ and $B_n$, the Hamiltonian of A-Type atomic array can be rewritten as

$$H_{a}=\sum\limits_{n}^{N_{a}}[U_{1}(A_{n}^{\dagger }B_{n}+A_{n+1}^{\dagger }B_{n})+U_{2}(A_{n+1}^{\dagger }A_{n}+B_{n+1}^{\dagger }B_{n})+h.c.].$$

It describes a one-dimensional chain of ordinary atoms, and $U_{1}$ as well as $U_{2}$ represents the nearest-neighbor and next-nearest-neighbor interactions between atoms. As for the atomic array with the structure of Fig. 3(b), due to the presence of empty cavity, the next-nearest neighbor interactions between atoms $U_3$ can be neglected according to the conclusions in Fig. (2). Meanwhile $U_{1}$ and $U_{2}$ are both nearest-neighbor interactions. Then, the Hamiltonian B-Type atomic array can be written as

$$H_{b}=\sum\limits_{n}^{N_{a}}(U_{1}A_{n}^{\dagger }B_{n}+U_{2}A_{n+1}^{\dagger }B_{n}+h.c.),$$
where $U_{1}$ and $U_{2}$ are the intra-cell coupling and extra-cell coupling strength, respectively. Similarly, for the atomic array with the structure of Fig. 3(c), the Hamiltonian can be described as
$$H_{c}=\sum\limits_{n}^{N_{a}}(U_{2}A_{n}^{\dagger }B_{n}+U_{1}A_{n+1}^{\dagger }B_{n}+h.c.),$$
where $U_{2}$ and $U_{1}$ are the intra-cell coupling and extra-cell coupling strength, respectively. It is obvious that for the atomic array with B- type structure, the coupling intensity of the intra-cell is larger than extra-cell coupling ($U_{1}>U_{2}$). However, for C-Type atomic array, the relation of intensity between inter-cell and extra-cell is reversed. In Fig. (4), we plot the energy spectrum of the three structures and the distribution of the zero-energy states. In Fig. 4(a), we can see that for A-Type atomic array the energy bands are continuous, therefore it is topologically trivial. For B-Type atomic array, as shown in Fig. 4(b), the energy band is open but doesn’t exist the bandgap state, so it is still topologically trivial. However, for the C-Type atomic array, it can be clearly seen that there are two degenerated zero-energy states in the bandgap shown in Fig. 4(c). Comparing Fig. 4(b) with Fig. 4(c), we can see that the two atomic arrays B-Type, C-Type are indeed different. We also plot the probability distribution for one of the bandgap states $\Psi$ (corresponding to the $N_a$th eigenvalue of the C-Type atomic array) in Fig. 4(d). It can be found that this bandgap state is distributed with equal probability at both ends of the atomic arrays. The above special zero-energy states, which are usually called edge states, are topologically protected as well as immune to a certain degree of disorder. The internal and external couplings are related with the spacing distance $x_0$. If $x_0\pm \xi$, as long as the slight disorder intensity $\xi$ does not cause a change in the ratio of internal to external coupling strength from greater than one to less than one, it has no effect on the existence of the bandgap state. The same conclusion is also found in [38,51]. In the current scheme, this bandgap state is not immune to atomic intrinsic loss. We still require the long-life of the atoms, which we discuss the long-life artificial atoms in section 6. In Appendix A, we calculated the winding numbers for the three structures. It turns out that for C-type atomic arrays, we can obtain a non-zero winding number, implying the existence of edge states under open boundary conditions as shown in Fig. 4(d).

 figure: Fig. 4.

Fig. 4. (a)(b)(c) Energy spectra corresponding to A-Type, B-Type and C-Type atomic array, respectively. (d) The probability distribution of the zero energy state at the lattice point corresponds to (c). The parameters are set to $\delta =2.9J$, $U_{1}=-0.017J$, $U_{2}=0.006J$ and $N_{a}=30$.

Download Full Size | PDF

From the above results, it can be found that the topologically trivial phase and non-trivial phase can be achieved by designing the position of the atoms in the coupled cavity array. However, it is noted that the coupling coefficient $U_{i,j}$ decreases with the increasing of atomic distance when $\delta >2.2J$. Therefore, it is impossible to adjust the other parameters of the system to change the intra-cell coupling strength from greater than the extra-cell coupling strength to less than the extra-cell coupling strength. This means that the system does not undergo the topological phase transition by changing in the parameters of the current system.

4. Topological phase transition induced by atomic periodic driving

As mentioned above, the topological phase can be simulated by designing different geometry structures of the atomic arrays, but the topological phase transition can not be achieved due to the fixed intra-cell and extra-cell coupling strength. Fortunately, if we introduce periodic modulation of the atoms, then the topological phase transition can be induced by adjusting the external field parameters.

For atomic periodic driving [36,52], it can be read as

$$H_{d}=\frac{\eta \omega _{d}}{2}\sum\limits_{j=1}^{2N_{a}}\cos {(\Delta kx_{j}+\phi )}\cos {\omega _{d}t}\sigma _{j}^{z},$$
where $\omega _{d}$ is the driving frequency, $\eta$ express the dimensionless driving strength with a pumping frequency scale, $\Delta k$ is the wave vector along the array direction, $x_{j}$ is the position of atoms, and $\phi$ is a global optical shift. We consider the atomic periodic drives are near resonant with the atoms or $\omega _{d}> \omega _{o}$, while the cavity field is far from resonance with the atoms ( $\delta >J>g$). Therefore, the introduction of the pumping field does not affect cavity fields as well as the atom-cavity interactions. Then we can obtain
$$H_{eff}=\sum\limits_{j,i}^{2N_{a}}{U_{j,i}}(\sigma _{j}^{\dagger }\sigma _{i}+\sigma _{i}^{\dagger }\sigma _{j})+\frac{\eta \omega _{d}}{2} \sum\limits_{j=1}^{2N_{a}}\cos {(\Delta kx_{j}+\phi )}\cos {\omega _{d}t} \sigma _{j}^{z}.$$

Now, we perform a rotating transformation $U^{\dagger }HU-iU^{\dagger }\dot {U}$ with $U=\exp \{i\sum\limits _{j=1}^{2N_{a}}\frac {\eta }{2}$ $\sin {\omega _{d}t} \cos {(\Delta kx_{j}+\phi )}\sigma _{j}^{z}\}$. After that, the Hamiltonian reads

$$H_{eff}=\sum\limits_{j,i}^{2N_{a}}{U_{j,i}}(\sigma _{j}^{\dagger }\sigma _{i}e^{iz\sin {\omega _{d}t}}+h.c.),$$
where $z=2\eta \sin {[\frac {\Delta k}{2}(x_j+x_i)+\phi ]}\sin {[\frac {\Delta k}{2}(x_j-x_i)]}$. Next, we exploit the Jacobi-Anger expansions $e^{iz\sin {\omega _d t} }=\sum\limits _{m=-\infty }^{+\infty }\mathcal {J}_{m}(z)e^{im\omega _d t}$ with the $m$th order of the first kind of Bessel function $\mathcal {J}_{m}(z)$. Assuming that $\omega _{d}\gg U_{i,j}$, the only non-fast oscillatory term contribution stems from $m=0$, so under the rotating wave approximation, we can neglect the rapid oscillatory terms, then the Hamiltonian becomes
$$H^{\prime }_{eff}=\sum\limits_{j,i}^{2N_{a}}{U_{i,j}\varXi_{j,i}}(\sigma _{j}^{\dagger }\sigma _{i}+\sigma _{i}^{\dagger }\sigma _{j}),$$
where $\varXi _{j,i}=\mathcal {J}_{0}(z)$. Plugging Eq. (10) into Eq. (20), then we obtain
$$H^{\prime }_{eff}=\sum\limits_{j=1}^{2N_{a}}U_{0}\varXi_{j,j}\sigma _{j}^{\dagger }\sigma _{j}+U_{1}\varXi_{j,j+1}(\sigma _{j}^{\dagger }\sigma _{j+1}+h.c.)+U_{2}\varXi_{j,j+2}(\sigma _{j}^{\dagger }\sigma _{j+2}+h.c.).$$

For A-Type atomic arrays, since the distance between atoms is homogeneous, so we can fix $\Delta k=\frac {\pi }{2x_0}$ to achieve the periodic couplings. According to the expression of Bessel function, we find that $\varXi _{j,i}$ has the following relation

$$\varXi_{j,j}=1, \quad \varXi_{j,j+2}=\left\{ \begin{array}{rl} \mathcal{J}_{0}(2\eta \sin {\phi }), & j=odd \\ \mathcal{J}_{0}(2\eta \cos {\phi }), & j=even \end{array} \right.$$
$$\varXi_{j,j+1}=\varXi_{j+2,j+3}=\left\{ \begin{array}{rl} \mathcal{J}_{0}(\sqrt{2}\eta \sin ({-\frac{\pi }{4}+\phi })), & j=odd \\ \mathcal{J}_{0}(\sqrt{2}\eta \sin ({\frac{\pi }{4}+\phi })), & j=even \end{array} \right.$$

Combining the above relationship and Eq. (14), then the Hamiltonian of the A-Type atomic arrays can be written in the following form

$$H_{a}^{\prime }=\sum\limits_{n}^{N_{a}}(v_{a}A_{n}^{\dagger }B_{n}+w_{a}A_{n+1}^{\dagger }B_{n}+T_{a}A_{n+1}^{\dagger }A_{n}+T_{b}B_{n+1}^{\dagger }B_{n}+h.c.) ,$$
with $v_{a}=\mathcal {J}_{0}(\sqrt {2}\eta \sin ({-\frac {\pi }{4}+\phi } ))U_{1}$, $w_{a}=\mathcal {J}_{0}(\sqrt {2}\eta \sin ({\frac {\pi }{4}+\phi }))U_{1}$, $T_{b}=\mathcal {J}_{0}(2\eta \cos {\phi })U_{2}$, $T_{a} =\mathcal {J}_{0}(2\eta \sin {\phi })U_{2}$. Furthermore, for the inhomogeneous distance between atoms, B-Type and C-Type atomic arrays, we choose $\Delta _k= \frac {\pi } {3x_0}$. After the same method of treatment analogous to the processing of A-Type atomic arrays, we can obtain
$$H_{b}^{\prime } =\sum\limits_{n}^{N_{a}}(v_{b}A_{n}^{\dagger }B_{n}+w_{b}A_{n+1}^{\dagger }B_{n}+h.c.),$$
$$H_{c}^{\prime } =\sum\limits_{n}^{N_{a}}(v_{c}A_{n}^{\dagger }B_{n}+w_{c}A_{n+1}^{\dagger }B_{n}+h.c.),$$
where $v_{b}=\mathcal {J}_{0}(\eta \cos {\phi })U_{1}, w_{b}=\mathcal {J}_{0}(\sqrt {3}\eta \sin {\phi })U_{2}$, $v_{c}=\mathcal {J}_{0}(\sqrt {3}\eta \sin {(\frac {2\pi }{3}+\phi )})U_{2}$, $v_{c}=\mathcal {J}_{0}(\eta \sin {(\frac {\pi }{6}+\phi )}\sin {\frac {\pi }{12}})U_{1}$. Thus, introducing the atomic periodic driving does not cause additional terms, while the coupling coefficients are modulated by controlling the drive parameters.

In Fig. (5), we plot the variation of coupling coefficient $v_{a}$, $w_{a}$, $T_{a}$, $T_{b}$ with the drive parameters $\phi$ and $\eta$. As shown in Fig. 5(a), (b), (c) and (d), we can see that the periods of all coupling coefficients are almost $\pi$, while the yellow area of the coupling coefficient become narrow with the increasing of $\eta$. We also plot the above coupling coefficients varying with the phase $\phi$ when $\eta =2$ in Fig. 5(e) and (f), it is clearly seen that the coupling coefficients $v_{a} (T_b)$ and $w_{a} (T_b)$ both exhibit a sinusoidal function with period $\pi$ but opposite phase. This means that we can periodically control the strength of the intra-cell and extra-cell coupling by adjusting the drive parameters. For B-Type and C-Type atomic arrays, the same conclusion is shown in Fig. 5(g) and (h). In Fig. 6(a), (e) and (h), we plot the energy spectrum as functions of phase $\phi$ for A-Type (a), B-Type (e) and C-Type atomic arrays (h). For A-Type atomic array, we can see that the energy spectrum is no longer symmetrical, due to the presence of next-nearest-neighbor interactions between atoms that break the chiral symmetry. Moreover, the system has a pair of bandgap states are separated from each other in $\phi \in [0.5\pi, \pi ] \cup [1.5\pi, 2\pi ]$, depicted by red lines in the Fig. 6(a). It corresponds to the $N_a$th , ($N_a+1$)th eigenvalue of the A-Type atomic array, respectively. We use $\Psi _L$ and $\Psi _R$ to denote these two bandgap states. We also draw the probability distribution of the bandgap states ($\Psi _L$, $\Psi _R$) and find that different driving parameters $\phi$ affects the existence of edge states. When we select the appropriate drive parameter $\phi =0.8\pi$, these two band gap states are respectively localized at the end of the atomic chain depicted in Fig. 6(b). The presence of edge states implies that the atomic array is in topologically nontrivial phase. Furthermore, when we increase the driving parameter $\phi =1.3\pi$, we see that this state is not localized at the end of the atomic array, which indicates that the system lies in a topologically trivial phase as shown in Fig. 6(c). Eventually when $\phi =1.7\pi$, this state is localized at the right end shown in Fig. 6(d). Thus, we found that by slowly and adiabatically adjusting the driving parameters, the topological phase transition can be induced in the atomic chains.

 figure: Fig. 5.

Fig. 5. The variation of coupling coefficient with the drive parameters $\phi$ and $\eta$. (a) $v_{a}$, (b) $w_{a}$, (c) $T_{a}$, (d) $T_{b}$. (e) (f) (g) (h) $v_{a}, w_{a}, T_{a}, T_{b}$, $v_{b}, w_{b}, v_{c}, w_{c}$ as a function of drive parameters $\phi$. The parameters are $\delta =2.9J$, $U_{1}=-0.017J$ and $U_{2}=0.006J$, $\eta =2$ for (e) (f), $\eta =2.7$ for (g) and $\eta =4.5$ for (h).

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. The energy spectrum for A-Type (a), B-Type (e) and C-Type atomic arrays (h) varying with $\phi$. The probability distribution of the bandgap states ($\Psi _L$, $\Psi _R$) when $\phi =0.8\pi$ (b) , $\phi =1.3\pi$ (c), $\phi =1.7\pi$ (d) and $\eta =2$ for A-Type atomic arrays. The winding number as functions of phase $\phi$ (e) and the probability distribution of the bandgap states $\Psi _L$ (f) when $\phi =\pi$ and $\eta =2.6$ for B-Type atomic array. (i) The winding number varying with $\phi$ and (j) the probability distribution of the bandgap states $\Psi _L$ when $\phi =\pi$ and $\eta =4.5$ for C-Type atomic array. The parameters are $\delta =2.9J$, $U_{1}=-0.017J$, $U_{2}=0.006J$, and $N_{a}=30$.

Download Full Size | PDF

For B-Type and C-Type atomic arrays, we also plot the energy spectrum and the winding number as functions of phase $\phi$, where the definition and the calculation of winding number is given in Appendix A. We can see the symmetry of the energy spectrum, while as $\phi$ changes, zero energy states appear in the bandgap shown in Fig. 6(e) and (h). By calculating the topological invariants of the bulk, such as the winding number $\nu$, we can predict the existence of edge states at the boundary, or vice versa, which is known as the bulk-edge correspondence [20]. In Fig. 6(f) and (i), it can be seen that for both types of atomic arrays, as $\phi$ varies, the winding number in the system changes between 0 and -1, corresponding to the transition from topologically trivial to non-trivial phase. In fact, for SSH chains with an even number of sites, there will be edge states at the end of the long chains in the topologically non-trivial phase. But for finite chains, the two edge modes are hybridized as shown in Fig. 6(g) and (j), where $\Psi _L$ corresponding to the $N_a$th eigenvalue of the B-Type and C-Type atomic array, respectively. Furthermore, for A-Type atomic arrays, $H_{a}^{\prime }$ no longer has a chiral symmetry due to the presence of atomic next-neighbour interactions, in which case the winding number can no longer be used as a topological invariant to describe the topological properties of the system. One can employ the Chen number [20] to indicate the topological properties.

Therefore, by introducing atomic periodic drives, we find that the topological phase transitions can be induced in the three structures of atomic arrays by slowly and adiabatically tuning the drive parameters. Furthermore, we note that the difference between the A-Type, B-Type and C-Type atomic arrays is the presence or absence of next-nearest neighbor interactions between atoms. In conjunction with Fig. 6(b), (g) and (j), it can be seen that the degenerate bandgap state is broken and localized at one end of the chain by considering the next-nearest-neighbor interaction of the atom.

5. Topological state transfer induced by next-nearest-neighbor interactions

In this section, we will show that the introduction of coupling in the next-nearest-neighbor leads to topological channels that allow quantum state transfer. We consider A-Type atomic array with $N_a=15$. For the SSH chain with an odd number of sites, there is only one edge state, which is positioned at the left or right edge of the chain, depending on the ratio of intra-cell to extra-cell. That is, if we can adiabatically change the coupling strength between inside and outside the cell from less than 1 to large than 1, then it is possible to pump the local state from one edge to the other thereby achieving quantum state transfer [5356].

In Fig. 7(a) and (c), we plot the energy spectrum as functions of $\phi$ within the range $[0.75\pi, 1.75\pi ]$ for two defferent detuning $\delta =3.2J$ (a), $2.9J$ (c) corresponding to the absence and presence of next-nearest neighbor interactions, respectively. As shown in Fig. 7(a), the system has a pair of degenerated zero-energy states in the bandgap $\phi \in [0.75\pi, \pi ]\cup [1.5\pi, 1.75\pi ]$. The zero-energy states are distributed with equal probability at two ends of the atomic array depicted in Fig. 7(b). This is because the edge states are not separated but mixed with each other due to the size effect, so it is difficult to achieve the topological quantum state transfer if one wants to use this bandgap state.

 figure: Fig. 7.

Fig. 7. (a) and (c) are the energy spectrum with the variation of drive parameter $\phi$, corresponding to $\delta =3.2J$, $2.9J$, respectively. (b) and (d) are their distribution of the bandgap states with the variation of drive parameter $\phi$. The parameters are set to $\eta =1.5$, and $N_{a}=30$.

Download Full Size | PDF

However, if we appropriately reduce the detuning $\delta$ and consider the next-nearest-neighbor interactions, then the transfer state problem mentioned above will be well resolved. In Fig. 7(c), it can be seen that the previously degenerated bandgap states gradually separate and merge into the bulk state. The probability distribution of bandgap states is given in Fig. 7(b), it can be seen that at $\phi \in [0.75\pi, 1\pi ]$, the bandgap states are mainly distributed on the left side of the atomic array and gradually integrate into the bulk state with the slowly and adiabatically change of $\phi$. When $\phi \in [1.5\pi, 1.75\pi ]$, the bandgap states concentrate in the rightmost. From the above results, it can be seen that topological state transfer channels can be induced by considering the next-nearest-neighbor interactions, and thus state transfer between left-edge states and right-edge states can be achieved by slowly and adiabatically tuning the phase $\phi$. The state transfer channel mentioned above is similar to Rice-Mele pumping [25], where the topological pumping is achieved through periodically modulated on-site potentials.

6. Discussion and conclusion

In this section we discuss the experimental feasibility analysis. In recent years, a large number of systems can be designed with some degree of waveguide QED physics such as solid-state emitters in nanophotonic structures [2,7,57], circuit QED [3,9,58], cold atoms in state-dependent optical lattices [59] et al. As we all know, two counter-propagating lasers can form a standing wave field and the atoms can be trapped in the optical lattice, therefore the equally spaced atomic array can be trapped [60,61]. Meanwhile, each atom is coupled to an additional optical cavity via dipole interactions. Hence, for the A-Type atomic arrays, it will be implemented using the standing-wave mechanism. However, for B-Type and C-Type atomic arrays, controlling inhomogeneous arrays of atoms using optical lattice or optical waveguides will be a challenge. In addition, our proposed scheme maybe implemented in the superconducting circuit, where the superconducting quantum qubit acts as a two-level atom, the cavity and the waveguide can be realized by using coupled superconducting transmission line resonators [4,6265]. Thus, the position of the atoms and cavities is fixed, and the transition frequency of the qubit is modulated by two periodic driving fields [52]. Experimentally, the coupling of adjacent resonator $J$ can be achieved by coupling an auxiliary capacitor, and the coupling strengths of $2\pi \times 31$ MHz can be obtained, which also can be significantly improved by increasing the capacitance [65]. Similarly, the coupling strength $g$ between resonator and qubit can be increased by boosting the impedance of the resonator [62]. It is therefore possible to adjust the capacitance of the auxiliary capacitor, and the inductance of the oscillator to achieve $g\ll J$. Moreover, the strong coupling between atoms and cavities ($g>\gamma _a, \gamma _s$) has been achieved experimentally in superconducting circuit [64]. Additionally, in Eq. (8), we ignore the decoherence rate ($\gamma _s$) of the atoms, which requires atoms to have long lifetimes. State-of-the-art superconducting qubits have coherence times in excess of 100 $\mu s$ (decoherence rates below 10 kHz). Considering the value of $U_2=0.006 \times 2\pi \times 31$ MHz, the dynamical evolution time is 0.85 $\mu s$ which is much shorter than 100 $\mu s$. Therefore, it should be reasonable to ignore the atomic inherent losses.

In conclusion, we propose a simple and feasible scheme to simulate the SSH model in the waveguide QED system. The photon bound state occurs when the atomic transition frequency happens to be in the photon bandgap, which establishes the coherent and dissipative coupling between atoms. Then, the topologically trivial and non-trivial phases can be achieved by considering atomic arrays with different distance structures. Furthermore, by introducing atomic periodic drives the topological phase transitions can be induced in the system by slowly and adiabatically tuning the drive parameters. Finally, for an even number of chains, we find that the introduction of next-nearest-neighbor interactions breaks the degenerated zero-energy state and establishes the topological state transport channel.

A. Appendix A: The bulk Hamiltonian and the topological invariants

To investigate the topological properties of the atomic arrays such as the winding number, topological phases and the existence of edge states, as in [18,25], we consider $N_a \to \infty$ and write the Hamiltonian Eq. (14) in momentum space. For A-Type atomic array, the bulk Hamiltonian as $H_{a}=\sum\limits _{k}\Psi _{k}^{\dagger }H_{a}(k)\Psi _{k}$, where $\Psi _{k}=(A_{k},B_{k})$ [66]. Introducing the vector of Pauli matrices $\vec{\sigma}$ and the identity matrix $I$, we obtain

$$H_{a}(k)=h_{I}(k)I+\vec{h}_a(k)\cdot \vec{\sigma},$$
with $h_{I}(k)=2U_{2}\cos {(kx_{0})}$, $\vec {h}_{a}(k)=(U_{1}+U_{1}\cos { (kx_{0})},U_{1}\sin {(kx_{0})})$. We can derive the eigenvalues of $H_{a}(k)$ as well as the eigenvectors as
$$E_{a}(k,\pm )=2U_{2}\cos {kx_{0}}\pm \sqrt{2U_{1}^{2}\cos {(kx_{0})}},\quad |a({\pm} k)\rangle =\frac{1}{\sqrt{2}}(e^{{-}i\phi _{a}(k)},\pm 1)^{T},$$
where $\phi _{a}(k)=\arctan {\frac {\sin (kx_{0})}{1+1\cos (kx_{0})}}$ (no relation with $U_{1}$ and $U_{2}$ ). The same method also can be used for both B, C-Types atomic chains. In ($h_x, h_y$) space, the direction of the vector $\vec {h}(k)$ represents an eigenstate, and the magnitude of the vector will give its eigenvalue. Due to the periodicity of the Brillouin zone, the vector $\vec {h}(k)$ must form a closed loop. As shown in Fig. ( 8), for A-Type and B-Type atomic arrays, the curve $\vec {h}(k)$ does not enclose the origin, and $\vert \phi (k)\vert <\pi /2$ for all $k$. However, for C-Type atomic array, the vector $\vec {h}(k)$ rotates around the origin and the phase $\phi (k)$ can take any values. The two-dimensional vector $\vec {h}(k)$ defines a trajectory in ($h_x, h_y$) space that is linked to the appearance of topological invariants. To describe the topological properties of the system, we calculated the Zak phase [67], defined as
$$Z=i\oint dk\langle k\vert\frac{d}{dk}\vert k\rangle=\frac{1}{2 }\oint dk\frac{d\phi}{dk},$$
where $\vert k\rangle$ is the eigenvector of the system in momentum space. Additionally, the winding number $\nu$ associated with the Zak phase can be found
$$\nu=\frac{1}{2\pi }\oint dk\frac{d\phi}{dk}=\frac{\Delta\phi}{2\pi},$$
where $\Delta \phi$ is the change in $\phi (k)$ with $k$ varying throughout the Brillouin zone. Therefore the Zak phase is $\pi$ times the winding number of the curve $\vec {h}(k)$ around the origin. So,
$$v=\left\{ \begin{array}{rl} 0, &{\rm A,\ B\ Type} \\ -1. & {\rm C\ Type} \end{array} \right.$$
where the minus sign refers to the clockwise rotation of the vector $\vec {h}(k)$ around the origin. Therefore, the winding number $\nu$ can be found for two different structures of the atomic arrays A, B-Type ($\nu =0$) and C-Type ($\nu =-1$), corresponding to the topological trivial phase and topological non-trivial phases, respectively. Moreover, we note that the first term of the Eq. (25) not breaks the chiral symmetry so that the energy spectrum is shifted but has no effect on the eigenstate Eq. (26) of the system, so the Zak phase is still quantized to multiples of $\pi$. We therefore ignore the on-site potential induced by photons in Eq. (14), Eq. (15) and Eq. (16).

 figure: Fig. 8.

Fig. 8. Trajectories of $\vec {h}(k)$ over the Brillouin zone in the ($h_x, h_y$) space for different atomic arrays (a) A-Type atomic array. (b) B-Type atomic array. (c) C-Type atomic array. The red arrow represents the vector $\vec {h}(k)$, whose direction represents an eigenstate and whose mode indicates its corresponding eigenvalue. The blue circles represents the trajectory of vector $\vec {h}(k)$ in $\vec {k}$ space. The parameters are set to $U_{1}=-0.017J$ and $U_{2}=0.006J$.

Download Full Size | PDF

Funding

Key Technologies Research and Development Program (2021YFE0193500); National Natural Science Foundation of China (12274053, 11874099).

Disclosures

The authors declare no conflict of interest.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

1. A. Goban, C.-L. Hung, S.-P. Yu, J. Hood, J. Muniz, J. Lee, M. Martin, A. McClung, K. Choi, D. E. Chang, O. Painter, and H. Kimble, “Atom-light interactions in photonic crystals,” Nat. Commun. 5(1), 3808 (2014). [CrossRef]  

2. A. Goban, C.-L. Hung, J. D. Hood, S.-P. Yu, J. A. Muniz, O. Painter, and H. J. Kimble, “Superradiance for atoms trapped along a photonic crystal waveguide,” Phys. Rev. Lett. 115(6), 063601 (2015). [CrossRef]  

3. E. Kim, X. Zhang, V. S. Ferreira, J. Banker, J. K. Iverson, A. Sipahigil, M. Bello, A. González-Tudela, M. Mirhosseini, and O. Painter, “Quantum electrodynamics in a topological waveguide,” Phys. Rev. X 11(1), 011015 (2021). [CrossRef]  

4. W. Nie, Z. H. Peng, F. Nori, and Y.-x. Liu, “Topologically protected quantum coherence in a superatom,” Phys. Rev. Lett. 124(2), 023603 (2020). [CrossRef]  

5. M. Bello, G. Platero, J. I. Cirac, and A. González-Tudela, “Unconventional quantum optics in topological waveguide qed,” Sci. Adv. 5(7), eaaw0297 (2019). [CrossRef]  

6. S.-f. Qi and J. Jing, “Generation of bell and greenberger-horne-zeilinger states from a hybrid qubit-photon-magnon system,” Phys. Rev. A 105(2), 022624 (2022). [CrossRef]  

7. J. S. Douglas, H. Habibian, C.-L. Hung, A. V. Gorshkov, H. J. Kimble, and D. E. Chang, “Quantum many-body models with cold atoms coupled to photonic crystals,” Nat. Photonics 9(5), 326–331 (2015). [CrossRef]  

8. M. J. Hartmann, F. G. Brandao, and M. B. Plenio, “Strongly interacting polaritons in coupled arrays of cavities,” Nat. Phys. 2(12), 849–855 (2006). [CrossRef]  

9. M. Bello, G. Platero, and A. González-Tudela, “Spin many-body phases in standard- and topological-waveguide qed simulators,” PRX Quantum 3(1), 010336 (2022). [CrossRef]  

10. I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simulation,” Rev. Mod. Phys. 86(1), 153–185 (2014). [CrossRef]  

11. M. Almokhtar, M. Fujiwara, H. Takashima, and S. Takeuchi, “Numerical simulations of nanodiamond nitrogen-vacancy centers coupled with tapered optical fibers as hybrid quantum nanophotonic devices,” Opt. Express 22(17), 20045–20059 (2014). [CrossRef]  

12. S. Haroche and J.-M. Raimond, Exploring the quantum: atoms, cavities, and photons (Oxford university press, 2006).

13. D. Plankensteiner, C. Sommer, H. Ritsch, and C. Genes, “Cavity antiresonance spectroscopy of dipole coupled subradiant arrays,” Phys. Rev. Lett. 119(9), 093601 (2017). [CrossRef]  

14. H. Pichler, T. Ramos, A. J. Daley, and P. Zoller, “Quantum optics of chiral spin networks,” Phys. Rev. A 91(4), 042116 (2015). [CrossRef]  

15. A. González-Tudela and D. Porras, “Mesoscopic entanglement induced by spontaneous emission in solid-state quantum optics,” Phys. Rev. Lett. 110(8), 080502 (2013). [CrossRef]  

16. M. Ringel, M. Pletyukhov, and V. Gritsev, “Topologically protected strongly correlated states of photons,” New J. Phys. 16(11), 113030 (2014). [CrossRef]  

17. J.-T. Shen and S. Fan, “Strongly correlated multiparticle transport in one dimension through a quantum impurity,” Phys. Rev. A 76(6), 062709 (2007). [CrossRef]  

18. X.-X. Li, B. Li, and P.-B. Li, “Simulation of topological phases with color center arrays in phononic crystals,” Phys. Rev. Res. 2(1), 013121 (2020). [CrossRef]  

19. X.-X. Li, P.-B. Li, H.-R. Li, H. Gao, and F.-L. Li, “Simulation of topological zak phase in spin-phononic crystal networks,” Phys. Rev. Res. 3(1), 013025 (2021). [CrossRef]  

20. M. Z. Hasan and C. L. Kane, “Colloquium: Topological insulators,” Rev. Mod. Phys. 82(4), 3045–3067 (2010). [CrossRef]  

21. 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(1), 015006 (2019). [CrossRef]  

22. L. A. Wray, S.-Y. Xu, Y. Xia, D. Hsieh, A. V. Fedorov, Y. S. Hor, R. J. Cava, A. Bansil, H. Lin, and M. Z. Hasan, “A topological insulator surface under strong coulomb, magnetic and disorder perturbations,” Nat. Phys. 7(1), 32–37 (2011). [CrossRef]  

23. M. Wang, C. Doyle, B. Bell, M. J. Collins, E. Magi, B. J. Eggleton, M. Segev, and A. Blanco-Redondo, “Topologically protected entangled photonic states,” Nanophotonics 8(8), 1327–1335 (2019). [CrossRef]  

24. W. P. Su, J. R. Schrieffer, and A. J. Heeger, “Solitons in polyacetylene,” Phys. Rev. Lett. 42(25), 1698–1701 (1979). [CrossRef]  

25. J. K. Asbóth, L. Oroszlány, and A. Pályi, A Short Course on Topological Insulators (Springer International Publishing, 2016).

26. X.-F. Zhou, X.-W. Luo, S. Wang, G.-C. Guo, X. Zhou, H. Pu, and Z.-W. Zhou, “Dynamically manipulating topological physics and edge modes in a single degenerate optical cavity,” Phys. Rev. Lett. 118(8), 083603 (2017). [CrossRef]  

27. F. Liu and K. Wakabayashi, “Novel topological phase with a zero berry curvature,” Phys. Rev. Lett. 118(7), 076803 (2017). [CrossRef]  

28. B.-Y. Xie, G.-X. Su, H.-F. Wang, H. Su, X.-P. Shen, P. Zhan, M.-H. Lu, Z.-L. Wang, and Y.-F. Chen, “Visualization of higher-order topological insulating phases in two-dimensional dielectric photonic crystals,” Phys. Rev. Lett. 122(23), 233903 (2019). [CrossRef]  

29. M. I. Shalaev, W. Walasik, A. Tsukernik, Y. Xu, and N. M. Litchinitser, “Robust topologically protected transport in photonic crystals at telecommunication wavelengths,” Nat. Nanotechnol. 14(1), 31–34 (2019). [CrossRef]  

30. V. Peano, C. Brendel, M. Schmidt, and F. Marquardt, “Topological phases of sound and light,” Phys. Rev. X 5(3), 031011 (2015). [CrossRef]  

31. L. Qi, Y. Xing, S. Liu, S. Zhang, and H.-F. Wang, “Topological phase induced by distinguishing parameter regimes in a cavity optomechanical system with multiple mechanical resonators,” Phys. Rev. A 101(5), 052325 (2020). [CrossRef]  

32. Y. Xing, L. Qi, J. Cao, D.-Y. Wang, C.-H. Bai, W.-X. Cui, H.-F. Wang, A.-D. Zhu, and S. Zhang, “Controllable photonic and phononic edge localization via optomechanically induced Kitaev phase,” Opt. Express 26(13), 16250 (2018). [CrossRef]  

33. X. Z. Hao, X. Y. Zhang, Y. H. Zhou, C. M. Dai, S. C. Hou, and X. X. Yi, “Topologically protected optomechanically induced transparency in a one-dimensional optomechanical array,” Phys. Rev. A 105(1), 013505 (2022). [CrossRef]  

34. L.-L. Wan, X.-Y. Lü, J.-H. Gao, and Y. Wu, “Controllable photon and phonon localization in optomechanical lieb lattices,” Opt. Express 25(15), 17364–17374 (2017). [CrossRef]  

35. L. Qi, Y. Xing, G.-L. Wang, S. Liu, S. Zhang, and H.-F. Wang, “Topological and nontopological edge states induced by qubit-assisted coupling potentials,” Ann. Phys. 532(7), 2000067 (2020). [CrossRef]  

36. P. Nevado, S. Fernández-Lorenzo, and D. Porras, “Topological edge states in periodically driven trapped-ion chains,” Phys. Rev. Lett. 119(21), 210401 (2017). [CrossRef]  

37. S. R. Pocock, X. Xiao, P. A. Huidobro, and V. Giannini, “Topological plasmonic chain with retardation and radiative effects,” ACS Photonics 5(6), 2271–2279 (2018). [CrossRef]  

38. W. Nie, M. Antezza, Y.-x. Liu, and F. Nori, “Dissipative topological phase transition with strong system-environment coupling,” Phys. Rev. Lett. 127(25), 250402 (2021). [CrossRef]  

39. 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 Rev. 10(6), 995–1001 (2016). [CrossRef]  

40. Z.-X. Zhang, R. Huang, L. Qi, Y. Xing, Z.-J. Zhang, and H.-F. Wang, “Topological phase transition and eigenstates localization in a generalized non-hermitian su–schrieffer–heeger model,” Ann. Phys. 533(1), 2000272 (2021). [CrossRef]  

41. N. Lang and H. P. Büchler, “Topological networks for quantum communication between distant qubits,” npj Quantum Inf 3(1), 47 (2017). [CrossRef]  

42. L. Qi, G.-L. Wang, S. Liu, S. Zhang, and H.-F. Wang, “Engineering the topological state transfer and topological beam splitter in an even-sized su-schrieffer-heeger chain,” Phys. Rev. A 102(2), 022404 (2020). [CrossRef]  

43. Z. C. Shi, X. L. Zhao, and X. X. Yi, “Robust state transfer with high fidelity in spin-1/2 chains by lyapunov control,” Phys. Rev. A 91(3), 032301 (2015). [CrossRef]  

44. E. J. Bergholtz, J. C. Budich, and F. K. Kunst, “Exceptional topology of non-hermitian systems,” Rev. Mod. Phys. 93(1), 015005 (2021). [CrossRef]  

45. F. M. D’Angelis, F. A. Pinheiro, D. Guéry-Odelin, S. Longhi, and F. m. c. Impens, “Fast and robust quantum state transfer in a topological su-schrieffer-heeger chain with next-to-nearest-neighbor interactions,” Phys. Rev. Res. 2(3), 033475 (2020). [CrossRef]  

46. C. C. Wanjura, M. Brunelli, and A. Nunnenkamp, “Topological framework for directional amplification in driven-dissipative cavity arrays,” Nat. Commun. 11(1), 3149 (2020). [CrossRef]  

47. D. Porras and S. Fernández-Lorenzo, “Topological amplification in photonic lattices,” Phys. Rev. Lett. 122(14), 143901 (2019). [CrossRef]  

48. G. Calajó, F. Ciccarello, D. Chang, and P. Rabl, “Atom-field dressed states in slow-light waveguide qed,” Phys. Rev. A 93(3), 033833 (2016). [CrossRef]  

49. L. Zhou, Z. R. Gong, Y.-x. Liu, C. P. Sun, and F. Nori, “Controllable scattering of a single photon inside a one-dimensional resonator waveguide,” Phys. Rev. Lett. 101(10), 100501 (2008). [CrossRef]  

50. C. McDonnell and B. Olmos, “Subradiant edge states in an atom chain with waveguide-mediated hopping,” arXiv 2205.13853 (2022).

51. S. Longhi, G. L. Giorgi, and R. Zambrini, “Landau–zener topological quantum state transfer,” Adv. Quantum Technol. 2(3-4), 1800090 (2019). [CrossRef]  

52. M. P. Silveri, J. A. Tuorila, E. V. Thuneberg, and G. S. Paraoanu, “Quantum systems under frequency modulation,” Rep. Prog. Phys. 80(5), 056002 (2017). [CrossRef]  

53. F. Mei, G. Chen, L. Tian, S.-L. Zhu, and S. Jia, “Robust quantum state transfer via topological edge states in superconducting qubit chains,” Phys. Rev. A 98(1), 012331 (2018). [CrossRef]  

54. K. Eckert, O. Romero-Isart, and A. Sanpera, “Efficient quantum state transfer in spin chains via adiabatic passage,” New J. Phys. 9(5), 155 (2007). [CrossRef]  

55. N. E. Palaiodimopoulos, I. Brouzos, F. K. Diakonos, and G. Theocharis, “Fast and robust quantum state transfer via a topological chain,” Phys. Rev. A 103(5), 052409 (2021). [CrossRef]  

56. S. Longhi, “Topological pumping of edge states via adiabatic passage,” Phys. Rev. B 99(15), 155150 (2019). [CrossRef]  

57. R. E. Evans, M. K. Bhaskar, D. D. Sukachev, C. T. Nguyen, A. Sipahigil, M. J. Burek, B. Machielse, G. H. Zhang, A. S. Zibrov, E. Bielejec, H. Park, M. Lončar, and M. D. Lukin, “Photon-mediated interactions between quantum emitters in a diamond nanocavity,” Science 362(6415), 662–665 (2018). [CrossRef]  

58. M. Mirhosseini, E. Kim, V. S. Ferreira, M. Kalaee, A. Sipahigil, A. J. Keller, and O. Painter, “Superconducting metamaterials for waveguide quantum electrodynamics,” Nat. Commun. 9(1), 3706 (2018). [CrossRef]  

59. I. de Vega, D. Porras, and J. Ignacio Cirac, “Matter-wave emission in optical lattices: Single particle and collective effects,” Phys. Rev. Lett. 101(26), 260404 (2008). [CrossRef]  

60. D. Jaksch, C. Bruder, J. I. Cirac, C. W. Gardiner, and P. Zoller, “Cold bosonic atoms in optical lattices,” Phys. Rev. Lett. 81(15), 3108–3111 (1998). [CrossRef]  

61. R. Grimm, M. Weidemüller, and Y. B. Ovchinnikov, “Optical dipole traps for neutral atoms,” (Academic Press, 2000), pp. 95–170.

62. A. Blais, A. L. Grimsmo, S. M. Girvin, and A. Wallraff, “Circuit quantum electrodynamics,” Rev. Mod. Phys. 93(2), 025005 (2021). [CrossRef]  

63. Y. Liu and A. A. Houck, “Quantum electrodynamics near a photonic bandgap,” Nat. Phys. 13(1), 48–52 (2017). [CrossRef]  

64. A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, “Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics,” Nature 431(7005), 162–167 (2004). [CrossRef]  

65. A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, “Cavity quantum electrodynamics for superconducting electrical circuits: An architecture for quantum computation,” Phys. Rev. A 69(6), 062320 (2004). [CrossRef]  

66. P. Delplace, D. Ullmo, and G. Montambaux, “Zak phase and the existence of edge states in graphene,” Phys. Rev. B 84(19), 195452 (2011). [CrossRef]  

67. J. Zak, “Berry’s phase for energy bands in solids,” Phys. Rev. Lett. 62(23), 2747–2750 (1989). [CrossRef]  

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1.
Fig. 1. Schematic diagram of the $2N_{a}$ two-level atoms coupled into an optical waveguide. The coupling strength and distance between adjacent cavities are $J$ and $x_0$.
Fig. 2.
Fig. 2. Correlated decay rates $\Gamma _{i, j}$ and coherent dipole-dipole interactions $U_{i, j}$ versus $\delta$ for different distance between atoms $|x_{i}-x_{j}|=0, x_0, 2x_0, 3x_0$ corresponds to the blue ($U_0, \Gamma _0$), red ($U_1, \Gamma _1$), purple ($U_2, \Gamma _2$), and cyan ($U_3, \Gamma _3$), respectively. The other parameters are $\gamma _{a}=0.25J$ and $g=0.3J$.
Fig. 3.
Fig. 3. (a) (b) (c) Three different arrangements of atomic array in coupled cavity arrays. We label the odd number of atoms as $A_{n}$ and the even number of atoms as $B_{n}$, $n=1,2,\ldots,N_{a}$.
Fig. 4.
Fig. 4. (a)(b)(c) Energy spectra corresponding to A-Type, B-Type and C-Type atomic array, respectively. (d) The probability distribution of the zero energy state at the lattice point corresponds to (c). The parameters are set to $\delta =2.9J$, $U_{1}=-0.017J$, $U_{2}=0.006J$ and $N_{a}=30$.
Fig. 5.
Fig. 5. The variation of coupling coefficient with the drive parameters $\phi$ and $\eta$. (a) $v_{a}$, (b) $w_{a}$, (c) $T_{a}$, (d) $T_{b}$. (e) (f) (g) (h) $v_{a}, w_{a}, T_{a}, T_{b}$, $v_{b}, w_{b}, v_{c}, w_{c}$ as a function of drive parameters $\phi$. The parameters are $\delta =2.9J$, $U_{1}=-0.017J$ and $U_{2}=0.006J$, $\eta =2$ for (e) (f), $\eta =2.7$ for (g) and $\eta =4.5$ for (h).
Fig. 6.
Fig. 6. The energy spectrum for A-Type (a), B-Type (e) and C-Type atomic arrays (h) varying with $\phi$. The probability distribution of the bandgap states ($\Psi _L$, $\Psi _R$) when $\phi =0.8\pi$ (b) , $\phi =1.3\pi$ (c), $\phi =1.7\pi$ (d) and $\eta =2$ for A-Type atomic arrays. The winding number as functions of phase $\phi$ (e) and the probability distribution of the bandgap states $\Psi _L$ (f) when $\phi =\pi$ and $\eta =2.6$ for B-Type atomic array. (i) The winding number varying with $\phi$ and (j) the probability distribution of the bandgap states $\Psi _L$ when $\phi =\pi$ and $\eta =4.5$ for C-Type atomic array. The parameters are $\delta =2.9J$, $U_{1}=-0.017J$, $U_{2}=0.006J$, and $N_{a}=30$.
Fig. 7.
Fig. 7. (a) and (c) are the energy spectrum with the variation of drive parameter $\phi$, corresponding to $\delta =3.2J$, $2.9J$, respectively. (b) and (d) are their distribution of the bandgap states with the variation of drive parameter $\phi$. The parameters are set to $\eta =1.5$, and $N_{a}=30$.
Fig. 8.
Fig. 8. Trajectories of $\vec {h}(k)$ over the Brillouin zone in the ($h_x, h_y$) space for different atomic arrays (a) A-Type atomic array. (b) B-Type atomic array. (c) C-Type atomic array. The red arrow represents the vector $\vec {h}(k)$, whose direction represents an eigenstate and whose mode indicates its corresponding eigenvalue. The blue circles represents the trajectory of vector $\vec {h}(k)$ in $\vec {k}$ space. The parameters are set to $U_{1}=-0.017J$ and $U_{2}=0.006J$.

Equations (31)

Equations on this page are rendered with MathJax. Learn more.

H = H w + H s + H I ,
H w = j = N N ω a a j a j + J ( a j + 1 a j + h . c . ) ,
H s = i = 1 2 N a ω o σ i σ i ,
H I = j = N N i = 1 2 N a g ( a j σ i + σ i a j ) δ j , i .
H I = j , k g N ( F j , k a k σ j + F j , k a k σ j ) .
ρ ˙ = j , i = 1 2 N a A i j ( σ j ρ σ i σ i σ j ρ ) + A i j ( σ i ρ σ j ρ σ j σ i ) ,
A i j = 4 g 2 e i K | x i x j | 4 J 2 ( δ + i γ a 2 ) 2 ,
ρ ˙ = i [ H e f f , ρ ] + j , i = 1 2 N a Γ i , j ( 2 σ j ρ σ i σ i σ j ρ ρ σ i σ j ) ,
H e f f = i , j 2 N a U i , j ( σ i σ j + σ j σ i ) ,
U i , j = U 0 δ i , j + U 1 δ i , j + 1 + U 2 δ i , j + 2 ,
H e f f = j = 1 2 N a U 0 σ j σ j + U 1 ( σ j σ j + 1 + h . c . ) + U 2 ( σ j σ j + 2 + h . c . ) .
H I b = j = N N   i = 1 2 N a g ( a j σ i + σ i a j ) δ j , i ,
H I c = j = N N   i = 1 2 N a g ( a j σ i + σ i a j ) δ j , i ,
H a = n N a [ U 1 ( A n B n + A n + 1 B n ) + U 2 ( A n + 1 A n + B n + 1 B n ) + h . c . ] .
H b = n N a ( U 1 A n B n + U 2 A n + 1 B n + h . c . ) ,
H c = n N a ( U 2 A n B n + U 1 A n + 1 B n + h . c . ) ,
H d = η ω d 2 j = 1 2 N a cos ( Δ k x j + ϕ ) cos ω d t σ j z ,
H e f f = j , i 2 N a U j , i ( σ j σ i + σ i σ j ) + η ω d 2 j = 1 2 N a cos ( Δ k x j + ϕ ) cos ω d t σ j z .
H e f f = j , i 2 N a U j , i ( σ j σ i e i z sin ω d t + h . c . ) ,
H e f f = j , i 2 N a U i , j Ξ j , i ( σ j σ i + σ i σ j ) ,
H e f f = j = 1 2 N a U 0 Ξ j , j σ j σ j + U 1 Ξ j , j + 1 ( σ j σ j + 1 + h . c . ) + U 2 Ξ j , j + 2 ( σ j σ j + 2 + h . c . ) .
Ξ j , j = 1 , Ξ j , j + 2 = { J 0 ( 2 η sin ϕ ) , j = o d d J 0 ( 2 η cos ϕ ) , j = e v e n
Ξ j , j + 1 = Ξ j + 2 , j + 3 = { J 0 ( 2 η sin ( π 4 + ϕ ) ) , j = o d d J 0 ( 2 η sin ( π 4 + ϕ ) ) , j = e v e n
H a = n N a ( v a A n B n + w a A n + 1 B n + T a A n + 1 A n + T b B n + 1 B n + h . c . ) ,
H b = n N a ( v b A n B n + w b A n + 1 B n + h . c . ) ,
H c = n N a ( v c A n B n + w c A n + 1 B n + h . c . ) ,
H a ( k ) = h I ( k ) I + h a ( k ) σ ,
E a ( k , ± ) = 2 U 2 cos k x 0 ± 2 U 1 2 cos ( k x 0 ) , | a ( ± k ) = 1 2 ( e i ϕ a ( k ) , ± 1 ) T ,
Z = i d k k | d d k | k = 1 2 d k d ϕ d k ,
ν = 1 2 π d k d ϕ d k = Δ ϕ 2 π ,
v = { 0 , A ,   B   T y p e 1. C   T y p e
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.