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

Controllable single-photon scattering in a Rydberg-atom-waveguide system via van der Waals interactions

Peng-Fei Wang College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Lei Huang College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Yi-Long-Yue Guo College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Jing Wang College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Han-Xiao Zhang College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Hong Yang College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China    Dong Yan [email protected] College of Physics and Electronic Engineering, Hainan Normal University, Haikou 571158, People’s Republic of China
Abstract

We investigate single-photon scattering in a system comprising a waveguide coupled to a pair of Rydberg atoms, illuminated by a coherent field. By adjusting the interatomic distance, we can transition between the Rydberg blockade and Rydberg antiblockade regimes, as the van der Waals interaction strength varies with distance. These distinct regimes, manifesting themselves in single-photon scattering, allow flexible reflection control due to their analogy to those of a small- and giant-atom interactions with the waveguide. We also derive scattering criteria for Rydberg blockade and Rydberg antiblockade, corresponding to specific single-photon reflection spectrum. Based on these criteria, the blockade and antiblockade distances can be estimated.

I Introduction

Recently, there has been significant interest in light-matter interactions within waveguide structures, leading to the emergence of waveguide quantum electrodynamics (waveguide QED) [1, 2, 3, 4]. In quantum networks [5, 6], waveguides are typically regarded as quantum channels for photons, while atoms act as quantum nodes. By adjusting the properties of these quantum nodes, the transmission of photons within the waveguide can be controlled. Typically, atoms are three to four orders of magnitude smaller than the wavelength of photons traveling through the waveguide, allowing them to be effectively modeled as point-like dipoles. However, giant atoms [7, 8] have dimensions comparable to or even larger than the wavelength of photons propagating in the waveguide, rendering the traditional model inapplicable. Compared to traditional atomic systems, giant atoms offer several potential advantages, such as frequency-dependent relaxation rates and Lamb shifts [8], chiral and oscillating bound states [9, 10, 11, 12], anomalous single-photon scattering spectra [13, 14, 15, 16, 17, 18, 19, 20], decoherence-free interactions (DFIs) [21, 22, 23, 24, 25], nonexponential decay [26, 27, 28, 29, 30]. Currently, systems suitable for giant atoms include superconducting quantum circuits [22, 27, 31, 32], coupled waveguide arrays [28], synthetic frequency dimension [12], spin ensembles [33], and matter waves in optical lattices [34].

To explore new features and broaden the range of applications for giant atoms, it is crucial to develop additional methods for constructing them. In the search for approaches to build giant atoms, the goal is to find a controllable building block that exhibits the properties of giant atoms under specific conditions, while demonstrating distinct characteristics under others, and enabling the integration of multiple systems. Fortunately, Rydberg atoms precisely meet these requirements due to their controllability and compatibility, as they possess strong long-range dipole-dipole interactions and are highly sensitive to external fields [35]. Strong van der Waals interactions cause dipole blockade or antiblockade effects depending on the choice of detuning between the laser field and the atomic transition frequency. Both dipole blockade and antiblockade mechanisms have been successfully employed in various applications, including quantum information processing, quantum computation, and quantum simulation [36, 37, 38, 39, 40].

In pioneering work, a phase-dependent decay of the double Rydberg excitation, an archetype of the giant-atom effect, was observed using the Rydberg antiblockade effect [41, 42]. To the best of our knowledge, this represents the first real atoms realization of giant atoms in the optical regime. Rydberg-atom-assisted giant atoms open the door to exploring richer physics and unlocking more potential applications in quantum optics. Inspired by the core ideas in there pioneering works [41, 42], we investigate single-photon scattering in a coupled-resonator waveguide (CRW) system coupled to a pair of Rydberg atoms. Unlike the studies in [41, 42], our approach enables switching between the dipole blockade regime and the antiblockade regime [43, 44, 45] by directly changing the interatomic distance. In contrast, their systems are consistently operated within the antiblockade regime. Additionally, we focus on the single-photon scattering behaviors manifested through reflection spectrum in different regimes. This approach allows us to characterize the blockade and the antiblockade distances without measuring two-photon correlation function.

Refer to caption
Figure 1: (a) Schematic diagram of a coupled-resonator waveguide (CRW) containing a pair of two-level Rydberg atoms interacting via the van der Waals (vdW) potential VdV_{d}. An external coherent field with the Rabi frequency Ωc\Omega_{c}, perpendicular to the CRW, also illuminates the Rydberg atoms. (b) Band structure. The atomic transition frequency ωe\omega_{e} lies within the propagating band of the CRW with the central frequency ω0\omega_{0}, while the external field frequency ωc\omega_{c} falls within the band gap at ωe+Vd\omega_{e}+V_{d} (green shaded). Without loss of generality, we set ωeω0\omega_{e}\equiv\omega_{0}. (c1) A pair of Rydberg atoms modeled in the two-atom basis. The single photon propagating in the CRW with frequency ωk\omega_{k} drives the lower transitions, while the external coherent field drives the upper transitions. The vdW interaction VdV_{d} introduces a detuning on the double Rydberg state |e1e2|e_{1}e_{2}\rangle. (c2) Superatom model. The energy-level diagram of Fig. 1(c1) is replotted in the collective-state basis with ground state |G\mathrm{|}G\rangle, the single-excitation state |E(1)|E^{(1)}\rangle, and the double-excitation state |E(2)|E^{(2)}\rangle. Here, the coupling strengths are set as J0=JN=JJ_{0}=J_{N}=J. (c3) Truncated superatom model in the blockade regime. Only the single photon propagating in the CRW effectively drives the transition from the collective ground state |G\mathrm{|}G\rangle to the single-excitation state |E(1)|E^{(1)}\rangle due to the rigid Rydberg dipole blockade effect. (c4) Equivalent superatom model in the Rydberg antiblockade regime. The superatom model reduces to a two-level system consisting only of the ground state |G|G\rangle and the double-excitation state |E(2)|E^{(2)}\rangle when the system enters the Rydberg antiblockade regime, as there is almost no single Rydberg population in |E(1)|E^{(1)}\rangle. Note that the accumulated phase of π\pi arises from the external coherent field.

In Section II, we present the model and the equations governing the system. In Sections III and IV, we analyze the single-photon scattering phenomena arising from the coupling of a point-like Rydberg superatom and a giant atom with waveguide, respectively. In Section V, we discuss single-photon scattering during the transition between blockade and the antiblockade regimes by varying the interatomic distance. And obtain the effective criteria for Rydberg blockade and Rydberg antiblockade. By adjusting the ratio between the coupling strengths of the two atoms to their respective coupled resonators, we further verify our results. Finally, we conclude in Section VI.

II Model and equations

As illustrated in Fig. 1(a), the system under consideration comprises an infinitely long coupled-resonator waveguide (CRW) with a pair of two-level Rydberg atoms trapped in the 0th and NNth resonators, respectively. The two Rydberg atoms, sharing a resonant frequency ωe\omega_{e}, are driven both by the single photon propagating in the CRW (characterized by coupling frequency ωk\omega_{k} and coupling strengths J0J_{0} and JNJ_{N}) and by an external coherent field with frequency ωc\omega_{c} and Rabi frequency Ωc\Omega_{c}. When both atoms are excited into the Rydberg state, they interact via the van der Waals (vdW) potential Vd=C6/d6V_{d}=C_{6}/d^{6}, where C6C_{6} is the vdW coefficient and dd is the interatomic distance. It should be emphasized that dd and NN are mutually independent here. As shown in Figs. 1(b) and  1(c1), when ωe\omega_{e} lies within CRW band and ωc\omega_{c} resides outside it, the single photon in the CRW drives the lower transitions from |g1g2|g_{1}g_{2}\rangle to |e1g2|e_{1}g_{2}\rangle or |g1e2|g_{1}e_{2}\rangle with detuning Δk=ωkωe\Delta_{k}=\omega_{k}-\omega_{e}, while the external field drives the upper transitions from |e1g2|e_{1}g_{2}\rangle or |g1e2|g_{1}e_{2}\rangle to |e1e2|e_{1}e_{2}\rangle with detuning Δc=Δ+Vd\Delta_{c}=\Delta+V_{d}, where Δ=ωeωc\Delta=\omega_{e}-\omega_{c}. The energy levels are represented as an effective four-level configuration in the two-atom basis (Fig. 1(c1)). By redefining the collective ground state as |G=|g1g2|G\rangle=|g_{1}g_{2}\rangle, the single-excitation state as |E(1)=(|e1g2+|g1e2)/2|E^{(1)}\rangle=(|{e_{1}}{g_{2}}\rangle+|{g_{1}}{e_{2}}\rangle)/{\sqrt{2}}, and the double-excitation state as |E(2)=|e1e2|E^{(2)}\rangle=|e_{1}e_{2}\rangle, we further obtain a superatomic configuration (Fig. 1(c2)). Varying the interatomic distance dd to be less than the blockade distance (such that ΔcΩc\Delta_{c}\gg\Omega_{c}) enforces a strict dipole blockade effect. This closes the double-excitation channels, effectively truncating the configuration in Fig. 1(c2) to a two-level Rydberg superatom (Fig. 1(c3)). In the antiblockade regime, where the double Rydberg excitation occurs, the energy levels reduce approximately to a two-level system comprising |G|G\rangle and |E(2)|E^{(2)}\rangle, as the single-excitation state |E(1)|E^{(1)}\rangle becomes negligibly populated. During the dynamics of the excitation and de-excitation driven by the external coherent field, an accumulated phase of π\pi is obtained (Fig. 1(c4)).

In a rotating frame, the Hamiltonian of the entire system can be written explicitly as three terms (see Appendix A)

H=HC+HJC+HAC.\begin{split}H=&H_{C}+H_{JC}+H_{AC}.\end{split} (1)

The Hamiltonian of the CRW reads (1\hbar\equiv 1 hereafter)

HC=(ω0ωe)jajajξj=(aj+1aj+ajaj+1),H_{C}=\left(\omega_{0}-\omega_{e}\right)\sum_{j}a_{j}^{\dagger}a_{j}-\xi\sum_{j=-\infty}^{\infty}(a_{j+1}^{\dagger}a_{j}+a_{j}^{\dagger}a_{j+1}), (2)

where ω0\omega_{0} and ωe\omega_{e} denote the central frequency of the resonators and the atomic transition frequency, respectively. aja_{j} (aja^{\dagger}_{j}) represents the bosonic annihilation (creation) operator on site jj, and ξ\xi is the hopping strength between the nearest-neighbor resonators.

The interaction between a pair of atoms and their respective resonators is described by a Jaynes-Cummings Hamiltonian

HJC=(J0a0|e1g2g1g2|+JNaN|g1e2g1g2|+H.c.).H_{JC}=(J_{0}{a_{0}}|{e_{1}}{g_{2}}\rangle\langle{g_{1}}{g_{2}}|+J_{N}{a_{N}}|{g_{1}}{e_{2}}\rangle\langle{g_{1}}{g_{2}}|+\mathrm{H}.\mathrm{c}.). (3)

The interaction between a pair of atoms and the external coherent field, as well as the vdW interaction between two Rydberg atoms, is described as follows:

HAC=Δc|e1e2e1e2|+Ωc(|e1e2e1g2|+|e1e2g1e2|+H.c.).H_{AC}=\Delta_{c}|{e_{1}}{e_{2}}\rangle\langle{e_{1}}{e_{2}}|+{\Omega_{c}}(|{e_{1}}{e_{2}}\rangle\langle{e_{1}}{g_{2}}|+|{e_{1}}{e_{2}}\rangle\langle{g_{1}}{e_{2}}|+\mathrm{H}.\mathrm{c}.). (4)

If the frequency ωk\omega_{k} of anincident single photon lies within the CRW band, while the frequency ωc\omega_{c} of the coherent field lies outside it. The two atoms by the incident single photon with wave vector kk and energy ωk=ω02ξcos(k)\omega_{k}=\omega_{0}-2\xi\cos(k), straightforwardly excites can be excited by the collective ground state |G|G\rangle to the single-excitation state |E(1)|E^{(1)}\rangle. Meanwhile, the external coherent field exclusively drives the transition from |E(1)|E^{(1)}\rangle to the double-excitation state |E(2)|E^{(2)}\rangle (see Fig. 1(c2)). Due to the conservation of the number of excitations in the system, the stationary eigenstate of the entire system in the superatom basis can be written as

|E=(jcjaj|g1g2+ue1g2|e1g2+ug1e2|g1e2+ue1e2|e1e2)|0,|E\rangle=(\sum_{j}c_{j}a_{j}^{\dagger}|g_{1}g_{2}\rangle+u_{e_{1}g_{2}}|e_{1}g_{2}\rangle+u_{g_{1}e_{2}}|g_{1}e_{2}\rangle+u_{e_{1}e_{2}}|e_{1}e_{2}\rangle)\otimes|0\rangle, (5)

where cjc_{j} represents the probability amplitude for finding a photonic excitation in resonator jj. ue1g2(ug1e2)u_{e_{1}g_{2}}(u_{g_{1}e_{2}}) denotes the probability amplitude of a single Rydberg atom being in the excited state |e1g2|e_{1}g_{2}\rangle (|g1e2|g_{1}e_{2}\rangle), while ue1e2u_{e_{1}e_{2}} represents the probability amplitude of both atoms being in the double-excitation state |e1e2|e_{1}e_{2}\rangle. In addition, |0|0\rangle indicates that all of the resonators are in the vacuum state. For a 1D scattering problem, the probability amplitude cjc_{j} can be written as:

cj={eikj+reikj,j<0αeikj+βeikj,0<j<Nteikj,j>N,c_{j}=\begin{cases}e^{ikj}+re^{-ikj},&\phantom{<}j<0\\ \alpha e^{ikj}+\beta e^{-ikj},&\phantom{<}0<j<N\\ te^{ikj},&\phantom{<}j>N\\ \end{cases}\\ , (6)

where rr and tt are the reflection and transmission amplitudes, respectively. In the range 0<j<N0<j<N between the two Rydberg atoms, the photon propagates forward with a probability amplitude α\alpha and backward with a probability amplitude β\beta between the resonators. By substituting j=0j=0 and j=Nj=N into Eq. 6, we obtain the continuity boundary conditions 1+r=α+β1+r=\alpha+\beta and αeikN+βeikN=teikN\alpha e^{ikN}+\beta e^{-ikN}=te^{ikN} at the respective sites. Together with the two continuity boundary conditions and solving the Schrödinger equation H|E=(ωkωe)|EH|E\rangle=(\omega_{k}-\omega_{e})|E\rangle. In a rotating frame, the real-space Hamiltonian H1H_{1} transforms to HH (see Appendix A), while simultaneously, the energy of the incident photon changes to ωkωe\omega_{k}-\omega_{e}. The reflection rate R=|r|2R=|r|^{2} can be obtained with (see Appendix B):

r=J2(eikN1)24iξΔksink+2J2(eikN1)+ηJ2(eikN+1)24iξsink(ηΔk2Ωc2)2ηJ2(eikN+1),\begin{split}r=\frac{J^{2}(e^{ikN}-1)^{2}}{4i\xi\Delta_{k}\sin k+2J^{2}(e^{ikN}-1)}+\frac{\eta J^{2}(e^{ikN}+1)^{2}}{4i\xi\sin k(\eta\Delta_{k}-2\Omega_{c}^{2})-2\eta J^{2}(e^{ikN}+1)},\end{split} (7)

where η=ΔkΔc\eta=\Delta_{k}-\Delta_{c}, and J0=JN=JJ_{0}=J_{N}=J.

III Features of single-photon scattering in a system where a CRW is coupled to a small atom

In this paper, we select realistic parameters of cold Rb87{}^{\mathrm{87}}\mathrm{Rb} atoms, such as the transition frequency ωe=2π×1009×1012\omega_{e}=2\pi\times 1009\times 10^{12}Hz and the van der Waals coefficient C6=2π×2.8×1012s1μm6C_{6}=2\pi\times 2.8\times 10^{12}\mathrm{s}^{-1}\mu m^{6} with the ground state |g1,2=|5S1/2,F=2,mF=2|g_{1,2}\rangle=|5S_{1/2},F=2,m_{F}=2\rangle and the Rydberg state |e1,2=75P3/2,mJ=3/2|e_{1,2}\rangle=\mid 75P_{3/2},m_{J}=3/2\rangle. To ensure that the frequency ωc\omega_{c} lies outside the CRW band while the atomic transition frequency ωe\omega_{e} remains within it, we set Δ0\Delta\neq 0 throughout this paper. Other parameters are given in the captions of the figures.

Refer to caption
Figure 2: The single-photon reflection rate RR as a function of detuning Δk/ξ\Delta_{k}/\xi for (a) N=0N=0, (b) N=21N=21, (c) even NN, and (d) odd NN in the blockade regime (d=3.1μmd=3.1\mu\text{m}). The thick line represents the reflection rate RR, calculated using the single-photon reflection formula in Ref. [46] with coupling strength 2J\sqrt{2}J, for a hybrid system of a small atom coupled to a CRW. The parameters are Ωc=5MHz\Omega_{c}=5~{}\text{MHz}, Δ=24MHz\Delta=-24~{}\text{MHz}, ξ=0.4MHz\xi=0.4~{}\text{MHz}, J=0.5ξJ=0.5~{}\xi.

When the distance between a pair of Rydberg atoms is d=3.1μmd=3.1\mu\mathrm{m}, they experience a significant vdW potential of Vd=19.8V_{d}=19.8GHz. In this case, since Vd|Δ|V_{d}\gg|\Delta|, the pair of Rydberg atoms behaves as a truncated superatom, excluding the double-excitation state due to the blockade effect. The blockade effect on single-photon scattering can be demonstrated through the single-photon reflection spectrum. In Fig. 2, we compare two types of single-photon reflection spectrum: one from a system with a small atom coupled to the CRW  [46], and the other from our system consisting of a pair of Rydberg atoms coupled to the CRW. For the case of N=0N=0, we observe in Fig. 2(a) that their behaviors match perfectly, indicating that the effect produced by the pair of Rydberg atoms is equivalent to that of a small atom. The reason is that for N=0N=0, the pair of Rydberg atoms is located in the same resonator, and therefore can be regarded as a truncated superatom without a double-excitation state due to the blockade effect, regardless of the size of the resonator.

We now consider the case of N0N\neq 0 while keeping the distance d constant. When N takes a relatively large value (e.g. N=21N=21), the accumulated phase ϕ=kN\phi=kN becomes significant, leading to a pronounced non-Markovian retardation effect. This phenomenon is manifested in Fig. 2(b) through strong oscillations in the single-photon reflection spectrum [19]. However, the envelope of the oscillating lines remains identical to that shown in Fig. 2(a). In Figs. 2(c) and  2(d), for smaller NN, it is clearly observed that the width of the reflection window decreases with increasing NN for even NN, while it remains nearly constant regardless of NN for odd NN. Furthermore, the reflection window in the latter case is narrower than that in the former. From the fully consistent envelope of all oscillating lines in Fig. 2, we conclude that the underlying physics is governed by the regime in which the CRW is coupled to the truncated superatom (blockade), exhibiting point-like behavior akin to a small atom, independent of the internal structure of the CRW (N0N\neq 0). In this case, when the incident photon resonates with the atomic transition, i.e., Δk=0\Delta_{k}=0, it is completely reflected with R=1R=1.

IV Features of single-photon scattering in a system where a CRW is coupled to a giant atom

Refer to caption
Figure 3: The single-photon reflection rate RR as a function of detuning Δk/ξ\Delta_{k}/\xi for (a) even NN, and (b) odd NN in the antiblockade regime (d=9.5μmd=9.5\mu\text{m}). Other parameters, such as Ωc\Omega_{c}, Δ\Delta, ξ\xi, and JJ, are the same as in Fig. 2.

When the interatomic distance is d=9.5μd=9.5\mum, the two atoms experience a vdW potential of Vd=24V_{d}=24MHz. In this case, Rydberg antiblockade occurs because the vdW interaction-induced shift is well compensated by the detuning [43, 44, 45], i.e., Δ+Vd=0\Delta+V_{d}=0. In the antiblockade regime, as shown in Fig. 3, we observe completely different behaviors in the single-photon scattering arising from the Rydberg double-excitation (see Fig. 1(c4)). These characteristics align perfectly with the single-photon scattering of a giant atom coupled to the CRW [13], apart from a π\pi-phase accumulation. This phenomenon can be clearly explained by the effective relaxation rate of the giant atom. Under the Markovian limit, the effective relaxation rate of a giant atom can be expressed as  [8, 22, 47, 48]

Γeff=4πJ2[1+cos(ϕ)cos(kaN)]D(ωa),\Gamma_{\mathrm{eff}}=4\pi J^{2}[1+\cos(\phi)\cos(k_{a}N)]D(\omega_{a}), (8)

where kak_{a} and D(ωa)D(\omega_{a}) are, respectively, the wave vector and the density of states of the field at the atomic frequency ωa\omega_{a}. ϕ\phi denotes the accumulated phase arising from the coherent driving, i.e., ϕ=π\phi=\pi in this context. Γeff\Gamma_{\mathrm{eff}} indicates that the single-photon reflection rate RR is generally depends on NN with ka/π=k/π=0.5k_{a}/\pi=k/\pi=-0.5 and ka/π=0.5k_{a}/\pi=0.5, respectively (see Fig. 1(b)). Specifically, as shown in Fig. 3(a), when N=2,6,,4m+2,N=2,6,\dots,4m+2,\dots (m=0,1,m=0,1,\dots), kNkN becomes an odd multiple of π\pi, leading to maximum dissipation for the giant atom. This results in the complete reflection of the single photon with R=1R=1; When N=4,8,,4m,N=4,8,\dots,4m,\dots (m=1,2,m=1,2,\dots), kNkN becomes an even multiple of π\pi, rendering the giant atom dissipationless. In this case, the giant atom and the resonator are effectively decoupled, allowing a single photon to travel through the waveguide without exciting any atoms, leading to R=0R=0. As shown in Fig. 3(b), when N=1,5,,4m+1,N=1,5,\dots,4m+1,\dots or N=3,7,,4m+3,N=3,7,\dots,4m+3,\dots (m=0,1,m=0,1,\dots), kNkN becomes a odd multiple of π/2\pi/2, resulting in normal dissipation for the giant atom, with R=0.5R=0.5. However, the reflection windows for these cases are separated and symmetrically distributed on both sides of Δk=0\Delta_{k}=0. Based on the analysis above, we conclude that a pair of Rydberg atoms in the antiblockade regime, coupled to the CRW, exhibits the giant-atom effect as described in Refs. [41, 42].

Refer to caption
Figure 4: (a) The single-photon reflection rate RR as a function of the interatomic distance dd with Δ=24MHz\Delta=-24\text{MHz}. (b) The blockade distance dbd_{b} and the antiblockade distance dabd_{ab} as a function of detuning Δ\Delta. Here, Δk=0\Delta_{k}=0. Other parameters, such as Ωc\Omega_{c}, ξ\xi, and JJ, are the same as in Fig. 2.

V Transitions in single-photon scattering: From small atoms to giant atoms in CRW

In Fig. 4(a), we plot the single-photon reflection rate RR as a function of the interatomic distance dd with detuning Δk=0\Delta_{k}=0 for different values of NN. Clearly, as the interatomic distance dd varies, three distinct regimes emerge: the blockade regime, the transition regime, and the antiblockade regime. The most characteristic feature of the blockade and antiblockade effects is that the behavior of the single-photon scattering remains almost unchanged as the distance varies. In contrast, in the transition regime, the scattering behavior changes significantly, except for N=4m+2N=4m+2 (m=0,1,m=0,1,\dots). Here, we focus on the blockade and antiblockade regimes. Specifically, in the blockade regime, for all values of NN, the incident photon is completely reflected (R=1R=1) up to the blockade distance db=3.5μd_{b}=3.5\mum due to the small-atom characteristics of the pair of Rydberg atoms, as demonstrated in Fig. 2. In the antiblockade regime, when N=4mN=4m (m=1,2,m=1,2,\dots) and N=4m+2N=4m+2 (m=0,1,m=0,1,\dots), the incident photon remains completely reflected (R=1R=1) and completely transmitted (R=0R=0), respectively. However, when N=4m+1N=4m+1 and N=4m+3N=4m+3, R=0.5R=0.5 emerges at the antiblockade distance dabd_{ab}, and beyond dabd_{ab}, RR is slightly above 0.50.5 and slightly below 0.50.5, respectively, due to the imperfect antiblockade effect (Δ+Vd0\Delta+V_{d}\neq 0). These behaviors are consistent with the single-photon scattering characteristics in the antiblockade regime, as shown in Fig. 3. Based on the results, the effective criteria for Rydberg excitation blockade and antiblockade are as follows: Blockade: R=1R=1 for all NN. Antiblockade: R=0.5R=0.5 for odd NN; R=0R=0 for N=4mN=4m (m=1,2,m=1,2,\dots); and R=1.0R=1.0 for N=4m+2N=4m+2 (m=0,1,m=0,1,\dots). Using these criteria, the maximally entangled state (|GE(1)+|E(1)G)/2(|GE^{(1)}\rangle+|E^{(1)}G\rangle)/\sqrt{2} (blockade regime) and non-entangled state |E(2)|E^{(2)}\rangle (antiblockade regime), can be accurately identified via the single-photon scattering.

In Fig. 4(b), we plot the blockade distance dbd_{b} and the antiblockade distance dabd_{ab} as functions of the detuning Δ\Delta with Δk=0\Delta_{k}=0. As Δ\Delta varies, the blockade distance dbd_{b} remains unchanged. This is due to two factors: first and foremost, the resonant excitation condition Δk=0\Delta_{k}=0 is consistently maintained, and second the single-excitation Rydberg state is unaffected by the external field. In contrast, the antiblockade distance dabd_{ab} decreases as the detuning Δ\Delta increases, following dab=(C6/Δ)1/6d_{ab}=(-C_{6}/\Delta)^{1/6}.

Refer to caption
Figure 5: The single-photon reflection rate RR as a function of detuning Δk\Delta_{k} and the interatomic distance dd for different values of NN. Other parameters, such as Ωc\Omega_{c}, ξ\xi, JJ and Δ\Delta, are the same as in Fig. 2.

In Fig. 5, we plot the single-photon reflection rate RR as a function of detuning Δk\Delta_{k} and interatomic distance dd for different values of NN. In each plot, a branch appears for Δk<0\Delta_{k}<0, governed by the dependence of the blockade distance db|Δk|1/6d_{b}\propto|\Delta_{k}|^{1/6}  [49]. Along this branch, except at discontinuity points (R=0R=0) occurring at Δk=2ξcos[(2n+1)π/N]\Delta_{k}=-2\xi\cos[(2n+1)\pi/N] (where nn is a natural number), the blockade effect dominates despite the presence of non-Markovian effect present at those discontinuities. For instance, at the left end of the branch, R=1R=1 persists uniformly for all NN. Here, the blockade distance dbd_{b} spans from d=3.5μmd=3.5\mu\mathrm{m} at Δk=0\Delta_{k}=0 to d=7.6μmd=7.6\mu\mathrm{m} at Δk=2ξ\Delta_{k}=-2\xi. For Δk>0\Delta_{k}>0, the blockade branch does not appear because the incident single photon with Δk>0\Delta_{k}>0 is decoupled from the resonator due to the external field’s detuning Δ<0\Delta<0. This behavior is illustrated in Fig. 6, where increasing the detuning Δ\Delta gives rise to another blockade branch that gradually converges toward the initial branch. In the limit of Δ\Delta\rightarrow-\infty (here, Δ=4.5GHz\Delta=-4.5~{}\text{GHz}), corresponding to an effective Rabi frequency Ωc2/Δ0\Omega_{c}^{2}/\Delta\rightarrow 0, the two branches merge and exhibit symmetry with respect to Δk=0\Delta_{k}=0, as the influence of the external field vanishes. Under these conditions, the blockade distance dbd_{b} for Δk0\Delta_{k}\neq 0 diminishes as Δ\Delta increases, following db|Δ|1/6d_{b}\propto|\Delta|^{-1/6}. In addition, Fig. 6(d) reveals a narrow gap corresponding to the antiblockade regime where Δ+Vd=0\Delta+V_{d}=0 (see the white dashed line).

Refer to caption
Figure 6: The single-photon reflection rate RR as a function of detuning Δk\Delta_{k} and the interatomic distance dd for N=1N=1. In (a) Δ=24MHz\Delta=-24~{}\text{MHz}, (b) Δ=100MHz\Delta=~{}-100\text{MHz}, (c) Δ=160MHz\Delta=~{}-160\text{MHz}, and (d) Δ=4.5GHz\Delta=-4.5~{}\text{GHz}. Other parameters, such as Ωc\Omega_{c}, ξ\xi, and JJ, are the same as in Fig. 2.
Refer to caption
Figure 7: The single-photon reflection rate RR as a function of the ratio of coupling strengths JN/J0J_{N}/J_{0}. In (a) d=3.1μmd=~{}3.1\mu\text{m}, and in (b) d=9.5μmd=~{}9.5\mu\text{m}. Here, Δk=0\Delta_{k}=~{}0 and Δ=24MHz\Delta=~{}-24\text{MHz}. Other parameters, such as Ωc\Omega_{c}, ξ\xi, and JJ, are the same as in Fig. 2.

Finally, we plot the single-photon reflection rate RR as a function of the ratio JN/J0J_{N}/J_{0} with detuning Δk=0\Delta_{k}=0 for different values of NN. As shown in Fig. 7(a), in the blockade regime, RR at Δk\Delta_{k} is independent of JN/J0J_{N}/J_{0} and the coupling sites, due to the point-like nature of the system, which holds regardless of NN. In the antiblockade regime, as shown in Fig. 7(b): for N=4m+1N=4m+1 and N=4m+3N=4m+3, the single-photon reflection rate starts at R=1R=1, decreases rapidly to R=0.5R=0.5 at JN/J0=1J_{N}/J_{0}=1, and then gradually increases back to R=1R=1 as JN/J0J_{N}/J_{0}\rightarrow\infty. This indicates that the coupling site of the giant atom plays a critical role because the size of the giant atom becomes significant in the antiblockade regime. For N=4m+2N=4m+2, the single-photon reflection rate RR remains constant at R=1R=1, independent of JN/J0J_{N}/J_{0}. For N=4mN=4m, the single-photon reflection rate RR remains at R=1R=1 except for R=0R=0 at JN/J0=1J_{N}/J_{0}=1. These results align well with the behaviors observed in Figs. 3 and 4.

VI Conclusions

We investigate single-photon scattering in system consisting of a waveguide coupled to a pair of Rydberg atoms, illuminated by an external coherent field. By tuning the van der Waals interactions between the Rydberg atoms through changes in the interatomic distance, a transition between the Rydberg blockade and antiblockade regime can be achieved, resulting in significant changes in single-photon scattering spectrum. In the blockade regime, a pair of Rydberg atoms behaves as a small atom, whereas in the antiblockade regime, it acts as a giant atom. Remarkably, both the blockade and the antiblockade distances can be estimated directly from the single-photon scattering spectrum without directly requiring calculations of two-photon correlation function. Our work enables the realization of controllable quantum networks with simplified architectures, offering new pathways for complex quantum-network integration. Furthermore, it advances precise quantum control and quantum information processing in Rydberg-atom-waveguide quantum electrodynamics systems, while also guiding the design of single-photon quantum devices.

Acknowledgments

We thank Dr. Lei Du, Prof. Jin-Hui Wu, and Prof. Zhi-Hai Wang for their fruitful discussions. This work is supported by the National Natural Science Foundation of China (Grant No.11874004, No.11204019, and No.12204137). It is also supported by the specific research fund of The Innovation Platform for Academicians of Hainan Province (Grant No.YSPTZX202207) and the Hainan Provincial Natural Science Foundation of China (Grant No.122QN302).

APPENDIX A THE HAMILTONIAN IN THE ROTATING WAVE FRAME

The original Hamiltonian of the system is given by:

H1=ω0jajajξj=(aj+1aj+ajaj+1)+J(a0|e1g2g1g2|+aN|g1e2g1g2|+H.c.)+ωe(|e1g2e1g2|+|g1e2g1e2|)+(2ωe+V6)|e1e2e1e2|+Ωceiωct(|e1e2e1g2|+|e1e2g1e2|+H.c.).\begin{split}H_{1}=&{\omega_{0}}\sum\limits_{j}{a_{j}^{\dagger}{a_{j}}}-\xi\sum\limits_{j=-\infty}^{\infty}{(a_{j+1}^{\dagger}{a_{j}}+a_{j}^{\dagger}{a_{j+1}})}+J{\rm{(}}{a_{0}}|{e_{1}}{g_{2}}\rangle\langle{g_{1}}{g_{2}}|+{a_{N}}|{g_{1}}{e_{2}}\rangle\langle{g_{1}}{g_{2}}|+\mathrm{H}.\mathrm{c}.)\\ &+{\omega_{e}}(|{e_{1}}{g_{2}}\rangle\langle{e_{1}}{g_{2}}|+|{g_{1}}{e_{2}}\rangle\langle{g_{1}}{e_{2}}|)+(2{\omega_{e}}+{V_{6}})|{e_{1}}{e_{2}}\rangle\langle{e_{1}}{e_{2}}|+{\Omega_{c}}{e^{-i{\omega_{c}}t}}(|{e_{1}}{e_{2}}\rangle\langle{e_{1}}{g_{2}}|+|{e_{1}}{e_{2}}\rangle\langle{g_{1}}{e_{2}}|+\mathrm{H}.\mathrm{c}.).\end{split} (9)

We now apply the Fourier transformation ak=jeikjaj/Nca_{k}^{\dagger}=\sum_{j}e^{-ikj}a_{j}^{\dagger}/\sqrt{N_{c}}, where NcN_{c}\to\infty represents the length of the CRW. In the momentum space, the Hamiltonian becomes:

Hp=kωkakak+[kJNcak(|e1g2g1g2|+|g1e2g1g2|eikN)+H.c.]+ωe(|e1g2e1g2|+|g1e2g1e2|)+(2ωe+V6)|e1e2e1e2|+Ωceiωct(|e1e2e1g2|+|e1e2g1e2|+H.c.),\begin{split}H_{p}=&\sum_{k}\omega_{k}a_{k}^{\dagger}a_{k}+[\sum_{k}\frac{J}{\sqrt{N_{c}}}a_{k}\left(|e_{1}g_{2}\rangle\langle g_{1}g_{2}|+|g_{1}e_{2}\rangle\langle g_{1}g_{2}|e^{ikN}\right)+\mathrm{H.c.}]\\ &+\omega_{e}\left(|e_{1}g_{2}\rangle\langle e_{1}g_{2}|+|g_{1}e_{2}\rangle\langle g_{1}e_{2}|\right)+\left(2\omega_{e}+V_{6}\right)|e_{1}e_{2}\rangle\langle e_{1}e_{2}|+\Omega_{c}e^{-i\omega_{c}t}(|e_{1}e_{2}\rangle\langle e_{1}g_{2}|+|e_{1}e_{2}\rangle\langle g_{1}e_{2}|+\mathrm{H.c.}),\end{split} (10)

where ωk=ω02ξcos(k)\omega_{k}=\omega_{0}-2\xi\cos(k). Using the unitary operator U=eiHtU=e^{-i{H^{{}^{\prime}}}t}, where H=kω1akak+ω2(|e1g2e1g2|+|g1e2g1e2|)+ω3|e1e2e1e2|{H^{{}^{\prime}}}=\sum_{k}{\omega_{1}a_{k}^{\dagger}a_{k}}+\omega_{2}\left(|e_{1}g_{2}\rangle\langle e_{1}g_{2}|+|g_{1}e_{2}\rangle\langle g_{1}e_{2}|\right)+\omega_{3}|e_{1}e_{2}\rangle\langle e_{1}e_{2}|, we perform a rotating frame transformation on Eq. (10) to obtain the effective Hamiltonian:

Hs=UHpU+idUdtU=UHpUH=k(ωkω1)akak+[kJNcei(ω1ω2)tak(|e1g2g1g2|+|g1e2g1g2|eikN)+H.c.]+(ωeω2)(|e1g2e1g2|+|g1e2g1e2|)+(2ωe+V6ω3)|e1e2e1e2|+Ωcei(ωc+ω2ω3)t(|e1e2e1g2|+|e1e2g1e2|+H.c.).\begin{split}{H_{s}}&={U^{\dagger}}H_{p}U+i\frac{{d{U^{\dagger}}}}{{dt}}U\\ &={U^{\dagger}}H_{p}U-{H^{{}^{\prime}}}\\ &=\sum_{k}{\left(\omega_{k}-\omega_{1}\right)a_{k}^{\dagger}a_{k}}+[\sum_{k}{\frac{J}{\sqrt{N_{c}}}e^{-i(\omega_{1}-\omega_{2})t}a_{k}\left(|e_{1}g_{2}\rangle\langle g_{1}g_{2}|+|g_{1}e_{2}\rangle\langle g_{1}g_{2}|e^{ikN}\right)+\mathrm{H}.\mathrm{c}.]}\\ &\phantom{=}\,\,+\left(\omega_{e}-\omega_{2}\right)(|e_{1}g_{2}\rangle\langle e_{1}g_{2}|+|g_{1}e_{2}\rangle\langle g_{1}e_{2}|)+(2\omega_{e}+V_{6}-\omega_{3})|e_{1}e_{2}\rangle\langle e_{1}e_{2}|\\ &\phantom{=}\,\,+\Omega_{c}e^{-i(\omega_{c}+\omega_{2}-\omega_{3})t}(|e_{1}e_{2}\rangle\langle e_{1}g_{2}|+|e_{1}e_{2}\rangle\langle g_{1}e_{2}|+\mathrm{H.c.}).\end{split} (11)

In Eq. (11), to eliminate the time-dependent terms, we set ω1=ω2=ωe\omega_{1}=\omega_{2}=\omega_{e} and ω3=ωc+ωe\omega_{3}=\omega_{c}+\omega_{e}. By applying an inverse Fourier transform to HsH_{s}, we obtain the real-space Hamiltonian in Eq. (1).

APPENDIX B SINGLE-PHOTON REFLECTION AMPLITUDE

This H|E=(ωkωe)|EH|E\rangle=(\omega_{k}-\omega_{e})|E\rangle results in the discrete scattering equation

{(ω0ωk)cj+Jδ0,jue1g2+JδN,jug1e2=ξ(cj+1+cj1)(ωkωe)ue1g2=Jc0+Ωcue1e2(ωkωe)ug1e2=JcN+Ωcue1e2(ωkωe)ue1e2=Δcue1e2+Ωcue1g2+Ωcug1e2.\begin{cases}(\omega_{0}-\omega_{k}){c_{j}}+J{\delta_{0,j}}{u_{e_{1}g_{2}}}+J{\delta_{N,j}}{u_{g_{1}e_{2}}}=\xi({c_{j+1}}+{c_{j-1}})\\ (\omega_{k}-{\omega_{e}}){u_{e_{1}g_{2}}}=J{c_{0}}+{\Omega_{c}}{u_{e_{1}e_{2}}}\\ (\omega_{k}-{\omega_{e}}){u_{g_{1}e_{2}}}=J{c_{N}}+{\Omega_{c}}{u_{e_{1}e_{2}}}\\ (\omega_{k}-{\omega_{e}}){u_{e_{1}e_{2}}}={\Delta_{c}}{u_{e_{1}e_{2}}}+{\Omega_{c}}{u_{e_{1}g_{2}}}+{\Omega_{c}}{u_{g_{1}e_{2}}}&\end{cases}. (12)

Substituting j=0j=0 and j=Nj=N into Eq. (12) results in two separate equations from the first equation due to the presence of the δ0,j\delta_{0,j} and δN,j\delta_{N,j} functions. At the same time, substituting Eq. (6) into Eq. (12) and applying the continuity boundary condition 1+r=α+β1+r=\alpha+\beta and αeikN+βeikN=teikN\alpha e^{ikN}+\beta e^{-ikN}=te^{ikN} leads to the following equations:

{(ω0ωk)(1+r)+Jue1g2=ξ[eik+reik+Aeik+Beik](ω0ωk)teikN+Jug1e2=ξ[Aeik(N1)+Beik(N1)+teik(N+1)]1+r=A+BAeikN+BeikN=teikN(ωkωe)ue1g2=J(1+r)+Ωcue1e2(ωkωe)ug1e2=JteikN+Ωcue1e2(ωkωe)ue1e2=Δcue1e2+Ωcue1g2+Ωcug1e2.\begin{cases}(\omega_{0}-\omega_{k})(1+r)+J{u_{e_{1}g_{2}}}=\xi[{e^{-ik}}+r{e^{ik}}+A{e^{ik}}+B{e^{-ik}}]\\ (\omega_{0}-\omega_{k})t{e^{ikN}}+J{u_{g_{1}e_{2}}}=\xi[A{e^{ik(N-1)}}+B{e^{-ik(N-1)}}+t{e^{ik(N+1)}}]\\ 1+r=A+B\\ A{e^{ikN}}+B{e^{-ikN}}=t{e^{ikN}}\\ (\omega_{k}-{\omega_{e}}){u_{e_{1}g_{2}}}=J(1+r)+{\Omega_{c}}{u_{e_{1}e_{2}}}\\ (\omega_{k}-{\omega_{e}}){u_{g_{1}e_{2}}}=Jt{e^{ikN}}+{\Omega_{c}}{u_{e_{1}e_{2}}}\\ (\omega_{k}-{\omega_{e}}){u_{e_{1}e_{2}}}={\Delta_{c}}{u_{e_{1}e_{2}}}+{\Omega_{c}}{u_{e_{1}g_{2}}}+{\Omega_{c}}{u_{g_{1}e_{2}}}&\end{cases}. (13)

By solving the equations related to rr, tt, AA, BB, ue1g2u_{e_{1}g_{2}}, ug1e2u_{g_{1}e_{2}}, and ue1e2u_{e_{1}e_{2}}, the expression for the reflection amplitude rr is derived in Eq. (7).

References

  • Haroche and Kleppner [1989] S. Haroche and D. Kleppner, Cavity quantum electrodynamics, Phys. Today 42, 24 (1989).
  • Mabuchi and Doherty [2002] H. Mabuchi and A. Doherty, Cavity quantum electrodynamics: coherence in context, Science 298, 1372 (2002).
  • Walther et al. [2006] H. Walther, B. T. Varcoe, B.-G. Englert, and T. Becker, Cavity quantum electrodynamics, Rep. Prog. Phys. 69, 1325 (2006).
  • Holzinger et al. [2022] R. Holzinger, R. Gutierrez-Jauregui, T. Hönigl-Decrinis, G. Kirchmair, A. Asenjo-Garcia, and H. Ritsch, Control of localized single-and many-body dark states in waveguide qed, Phys. Rev. Lett. 129, 253601 (2022).
  • Kuhn et al. [2002] A. Kuhn, M. Hennrich, and G. Rempe, Deterministic single-photon source for distributed quantum networking, Phys. Rev. Lett. 89, 067901 (2002).
  • Kimble [2008] H. J. Kimble, The quantum internet, Nature 453, 1023 (2008).
  • Kockum [2021] A. F. Kockum, Quantum Optics with Giant Atoms—the First Five Years, in International Symposium on Mathematics, Quantum Theory, and Cryptography, edited by T. Takagi, M. Wakayama, K. Tanaka, N. Kunihiro, K. Kimoto, and Y. Ikematsu (Springer Singapore, Singapore, 2021) pp. 125–146.
  • Kockum et al. [2014] A. F. Kockum, P. Delsing, and G. Johansson, Designing frequency-dependent relaxation rates and Lamb shifts for a giant artificial atom, Phys. Rev. A 90, 013837 (2014).
  • Wang et al. [2020] X. Wang, T. Liu, A. F. Kockum, H.-R. Li, and F. Nori, Tunable Chiral Bound States with Giant Atoms, Phys. Rev. Lett. 126, 043602 (2020).
  • Guo et al. [2020a] L. Guo, A. F. Kockum, F. Marquardt, and G. Johansson, Oscillating bound states for a giant atom, Phys. Rev. Res. 2, 043014 (2020a).
  • Lim et al. [2023] K. H. Lim, W.-K. Mok, and L.-C. Kwek, Oscillating bound states in non-markovian photonic lattices, Phys. Rev. A 107, 023716 (2023).
  • Du et al. [2022a] L. Du, Y. Zhang, J.-H. Wu, A. F. Kockum, and Y. Li, Giant atoms in a synthetic frequency dimension, Phys. Rev. Lett. 128, 223602 (2022a).
  • Zhao and Wang [2020] W. Zhao and Z.-H. Wang, Single-photon scattering and bound states in an atom-waveguide system with two or multiple coupling points, Phys. Rev. A 101, 053855 (2020).
  • Du and Li [2021] L. Du and Y. Li, Single-photon frequency conversion via a giant Λ\mathrm{\Lambda}-type atom, Phys. Rev. A 104, 023712 (2021).
  • Du et al. [2021] L. Du, Y.-T. Chen, and Y. Li, Nonreciprocal frequency conversion with chiral Λ\mathrm{\Lambda}-type atoms, Phys. Rev. Res. 3, 043226 (2021).
  • Cai and Jia [2021] Q.-Y. Cai and W.-Z. Jia, Coherent single-photon scattering spectra for a giant-atom waveguide-qed system beyond the dipole approximation, Phys. Rev. A 104, 033710 (2021).
  • Yin et al. [2022] X.-L. Yin, Y.-H. Liu, J.-F. Huang, and J.-Q. Liao, Single-photon scattering in a giant-molecule waveguide-qed system, Phys. Rev. A 106, 013715 (2022).
  • Zhu et al. [2022] Y.-T. Zhu, S. Xue, R.-B. Wu, W.-L. Li, Z.-H. Peng, and M. Jiang, Spatial-nonlocality-induced non-markovian electromagnetically induced transparency in a single giant atom, Phys. Rev. A 106, 043710 (2022).
  • Chen et al. [2022] Y.-T. Chen, L. Du, L. Guo, Z. Wang, Y. Zhang, Y. Li, and J.-H. Wu, Nonreciprocal and chiral single-photon scattering for giant atoms, Commun. Physc. 5, 215 (2022).
  • Zhou et al. [2023] J. Zhou, X.-L. Yin, and J.-Q. Liao, Chiral and nonreciprocal single-photon scattering in a chiral-giant-molecule waveguide-qed system, Phys. Rev. A 107, 063703 (2023).
  • Kockum et al. [2018] A. F. Kockum, G. Johansson, and F. Nori, Decoherence-Free Interaction between Giant Atoms in Waveguide Quantum Electrodynamics, Phys. Rev. Lett. 120, 140404 (2018).
  • Kannan et al. [2020] B. Kannan, M. Ruckriegel, D. Campbell, A. F. Kockum, J. Braumüller, D. Kim, M. Kjaergaard, P. Krantz, A. Melville, B. M. Niedzielski, A. Vepsäläinen, R. Winik, J. Yoder, F. Nori, T. P. Orlando, S. Gustavsson, and W. D. Oliver, Waveguide quantum electrodynamics with superconducting artificial giant atoms, Nature (London) 583, 775 (2020).
  • Carollo et al. [2020] A. Carollo, D. Cilluffo, and F. Ciccarello, Mechanism of decoherence-free coupling between giant atoms, Phys. Rev. Res. 2, 043184 (2020).
  • Du et al. [2023a] L. Du, L. Guo, and Y. Li, Complex decoherence-free interactions between giant atoms, Phys. Rev. A 107, 023705 (2023a).
  • Du et al. [2023b] L. Du, L. Guo, Y. Zhang, and A. F. Kockum, Giant emitters in a structured bath with non-hermitian skin effect, Phys. Rev. Res. 5, L042040 (2023b).
  • Guo et al. [2017] L. Guo, A. Grimsmo, A. F. Kockum, M. Pletyukhov, and G. Johansson, Giant acoustic atom: A single quantum system with a deterministic time delay, Phys. Rev. A 95, 053821 (2017).
  • Andersson et al. [2019] G. Andersson, B. Suri, L. Guo, T. Aref, and P. Delsing, Non-exponential decay of a giant artificial atom, Nature Physics 15, 1123 (2019).
  • Longhi [2020] S. Longhi, Photonic simulation of giant atom decay, Opt. Lett. 45, 3017 (2020).
  • Guo et al. [2020b] S. Guo, Y. Wang, T. Purdy, and J. Taylor, Beyond spontaneous emission: Giant atom bounded in the continuum, Phys. Rev. A 102, 033706 (2020b).
  • Du et al. [2022b] L. Du, Y.-T. Chen, Y. Zhang, and Y. Li, Giant atoms with time-dependent couplings, Phys. Rev. Res. 4, 023198 (2022b).
  • Gustafsson et al. [2014] M. V. Gustafsson, T. Aref, A. F. Kockum, M. K. Ekström, G. Johansson, and P. Delsing, Propagating phonons coupled to an artificial atom, Science 346, 207 (2014).
  • Vadiraj et al. [2021] A. Vadiraj, A. Ask, T. McConkey, I. Nsanzineza, C. S. Chang, A. F. Kockum, and C. Wilson, Engineering the level structure of a giant artificial atom in waveguide quantum electrodynamics, Phys. Rev. A 103, 023710 (2021).
  • Wang et al. [2022] Z.-Q. Wang, Y.-P. Wang, J. Yao, R.-C. Shen, W.-J. Wu, J. Qian, J. Li, S.-Y. Zhu, and J.-Q. You, Giant spin ensembles in waveguide magnonics, Nat. Commun. 13, 7580 (2022).
  • González-Tudela et al. [2019] A. González-Tudela, C. S. Muñoz, and J. I. Cirac, Engineering and harnessing giant atoms in high-dimensional baths: a proposal for implementation with cold atoms, Phys. Rev. Lett. 122, 203603 (2019).
  • Saffman et al. [2010] M. Saffman, T. G. Walker, and K. Mølmer, Quantum information with rydberg atoms, Rev. Mod. phys. 82, 2313 (2010).
  • Lukin et al. [2001] M. D. Lukin, M. Fleischhauer, R. Cote, L. Duan, D. Jaksch, J. I. Cirac, and P. Zoller, Dipole blockade and quantum information processing in mesoscopic atomic ensembles, Phys. Rev. Lett. 87, 037901 (2001).
  • Nguyen et al. [2018] T. L. Nguyen, J.-M. Raimond, C. Sayrin, R. Cortinas, T. Cantat-Moltrecht, F. Assemat, I. Dotsenko, S. Gleyzes, S. Haroche, G. Roux, et al., Towards quantum simulation with circular rydberg atoms, Phys. Rev. X 8, 011032 (2018).
  • Zhao et al. [2017] Y.-J. Zhao, B. Liu, Y.-Q. Ji, S.-Q. Tang, and X.-Q. Shao, Robust generation of entangled state via ground-state antiblockade of rydberg atoms, Scientific Reports 7, 16489 (2017).
  • Bienias et al. [2020] P. Bienias, J. Douglas, A. Paris-Mandoki, P. Titum, I. Mirgorodskiy, C. Tresp, E. Zeuthen, M. J. Gullans, M. Manzoni, S. Hofferberth, et al., Photon propagation through dissipative rydberg media at large input rates, Phys. Rev. Res. 2, 033049 (2020).
  • Petrosyan et al. [2017] D. Petrosyan, F. Motzoi, M. Saffman, and K. Mølmer, High-fidelity rydberg quantum gate via a two-atom dark state, Phys. Rev. A 96, 042306 (2017).
  • Chen et al. [2023] Y.-T. Chen, L. Du, Y. Zhang, L. Guo, J.-H. Wu, M. Artoni, and G. C. La Rocca, Giant-atom effects on population and entanglement dynamics of rydberg atoms in the optical regime, Phys. Rev. Res. 5, 043135 (2023).
  • Chen et al. [2024] Y.-T. Chen, L. Du, Z. Wang, M. Artoni, G. La Rocca, and J.-H. Wu, Single-photon manipulations based on optically controlled chiral couplings in waveguide structures of rydberg giant atoms, Phys. Rev. A 109, 063710 (2024).
  • Wu et al. [2022] J.-L. Wu, Y. Wang, J.-X. Han, S.-L. Su, Y. Xia, Y. Jiang, and J. Song, Unselective ground-state blockade of rydberg atoms for implementing quantum gates, Front. Phys. 17, 22501 (2022).
  • Liu et al. [2022] F. Liu, Z.-C. Yang, P. Bienias, T. Iadecola, and A. V. Gorshkov, Localization and criticality in antiblockaded two-dimensional rydberg atom arrays, Phys. Rev. Lett. 128, 013603 (2022).
  • Valencia-Tortora et al. [2024] R. J. Valencia-Tortora, N. Pancotti, M. Fleischhauer, H. Bernien, and J. Marino, Rydberg platform for nonergodic chiral quantum dynamics, Phys. Rev. Lett. 132, 223201 (2024).
  • Zhou et al. [2008] L. Zhou, Z. 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, 100501 (2008).
  • Roccati and Cilluffo [2024] F. Roccati and D. Cilluffo, Controlling markovianity with chiral giant atoms, Phys. Rev. Lett. 133, 063603 (2024).
  • Du et al. [2023c] L. Du, Y.-T. Chen, Y. Zhang, Y. Li, and J.-H. Wu, Decay dynamics of a giant atom in a structured bath with broken time-reversal symmetry, Quantum Sci. Technol. 8, 045010 (2023c).
  • Ma et al. [2019] D. Ma, D. Yu, X.-D. Zhao, and J. Qian, Unidirectional and controllable higher-order diffraction by a rydberg electromagnetically induced grating, Phys. Rev. A 99, 033826 (2019).