Skip to main content
Nature Communications logoLink to Nature Communications
. 2024 Nov 14;15:9848. doi: 10.1038/s41467-024-54199-w

Experimental realization of on-chip few-photon control around exceptional points

Pengtao Song 1,2,#, Xinhui Ruan 1,2,3,#, Haijin Ding 3, Shengyong Li 3, Ming Chen 1, Ran Huang 1, Le-Man Kuang 1, Qianchuan Zhao 3, Jaw-Shen Tsai 4,5, Hui Jing 1, Lan Yang 6, Franco Nori 4,7,8, Dongning Zheng 2,9,10,, Yu-xi Liu 11, Jing Zhang 12,13,, Zhihui Peng 1,10,
PMCID: PMC11561106  PMID: 39537631

Abstract

Non-Hermitian physical systems have attracted considerable attention in recent years for their unique properties around exceptional points (EPs), where the eigenvalues and eigenstates of the system coalesce. Phase transitions near exceptional points can lead to various interesting phenomena, such as unidirectional wave transmission. However, most of those studies are in the classical regime and whether these properties can be maintained in the quantum regime is still a subject of ongoing studies. Using a non-Hermitian on-chip superconducting quantum circuit, here we observe a phase transition and the corresponding exceptional point between the two phases. Furthermore, we demonstrate that unidirectional microwave transmission can be achieved even in the few-photon regime within the broken symmetry phase. This result holds some potential applications, such as on-chip few-photon microwave isolators. Our study reveals the possibility of exploring the fundamental physics and practical quantum devices with non-Hermitian systems based on superconducting quantum circuits.

Subject terms: Superconducting devices, Quantum optics, Quantum information


The authors observe an exceptional point and the corresponding phase transition in a superconducting non-Hermitian circuit. They find non-reciprocal microwave transmission can be achieved within the broken symmetry phase in the few-photon regime.

Introduction

Non-Hermitian physical systems with exceptional points15 (EPs) have attracted considerable attention in recent years. The coalescence of eigenvalues and eigenstates of the non-Hermitian Hamiltonian around EPs leads to various interesting properties and numerous applications, such as unidirectional transmission and invisibility612, chiral behavior13, control of lasers1417 or electromagnetic waves18, high-precision sensors1922, and topological energy transfer3,23. While non-Hermitian physics has been studied extensively in classical systems, the quantum behavior of non-Hermitian systems is not fully understood and there is still much to be explored in the quantum realm. Although there has been interesting work to mimic gain-loss-balance mechanisms in two-level systems (e.g., superconducting quantum circuits24,25, nitrogen-vacancy centers in diamonds26 and trapped ions27), various interesting problems related to quantum non-Hermitian systems in the few-photon regime are still left open28,29. One interesting topic is unidirectional control of light in the few-photon regime for non-Hermitian systems. Such kinds of quantum nonreciprocal effects are very helpful for constructing quantum devices30, such as quantum diodes31 and quantum circulators32, which are key components for quantum information processing. Although it is well known that non-reciprocal wave propagations can be achieved by breaking the symmetries of non-Hermitian physical systems in the vicinity of the exception point, it is challenging to show such kinds of symmetry breaking in the few-photon regime, because the enhanced quantum fluctuations at the exceptional point will disrupt this process.

Demonstrating nonreciprocity in the few-photon regime requires strong nonlinearity, which is quite difficult to achieve in optical systems. This issue could be solved by using superconducting systems, in which strong nonlinearity has been successfully demonstrated at the single-photon level in the microwave frequency regime33. Superconducting quantum information processing has made great progresses in recent years3437. Highly-efficient widely-tunable on-demand single-photon sources and highly-efficient single-photon detectors in the microwave-frequency regime have been realized in superconducting quantum systems3840. With the quantum state transfer by a single microwave photon between two superconducting qubits housed in two refrigerators separated apart by 5 meters, it is possible to realize a microwave quantum network and distributed superconducting quantum computing41. There has been some research on non-Hermitian phenomena in superconducting circuits4244. However, the control of the transmission of a single microwave photon on a chip is still ongoing research, and is one of the crucial open problems for quantum networks.

In this work, we construct a non-Hermitian system using a superconducting circuit consisting of two cavities and three tunable-gap flux qubits. The two cavities have different decay rates and nearly identical resonance frequencies, with tunability within a range of a few tens of MHz. The effective coupling strength between the cavities can be tuned by a quantum coupler, implemented here as a tunable-gap flux qubit. Using this non-Hermitian system, we observe the phase transitions and the corresponding exceptional points. The qubits introduce additional nonlinearity to the two cavities, enabling the generation of unidirectional microwave transmission. By tuning the effective coupling strength between the cavities and the input photon numbers, we investigate the switchable unidirectional transmission of the system at the few-photon level.

Results

Experimental setup

Our experiment is carried out in a dilution refrigerator with a base temperature of about 28 mK. As shown in Fig. 1a–c, the device consists of two superconducting coplanar waveguide (CPW) cavities (marked as Cavity a and Cavity b, respectively) that are capacitively coupled to a tunable-gap flux qubit (working as a quantum coupler and marked as qubit QC) with two nominal identical capacitors (see Fig. 1b, c). Tunable coupling between two CPW cavity modes or two CPW cavities based on a flux qubit have been demonstrated45,46. The Hamiltonian of a qubit coupled to a cavity in the dispersive regime is H(ωc+χ~σz)aa+(ωq+χ~)σz/2, where ωc is the cavity frequency, ωq is the qubit transition frequency, χ~=g02/δ is the qubit-state-dependent dispersive cavity shift33, δ=ωqωcg0, g0 is the coupling strength between the qubit and the cavity, σz=0011 with the ground 0 and excited 1 states of the qubit, and a and a are the creation and annihilation operators of the cavity, respectively. The eigenfrequencies of the two CPW cavities (ωa for cavity a and ωb for cavity b, respectively) can be shifted for a few tens of MHz in the dispersive coupling regime, because the cavities are capacitively coupled to other two auxiliary tunable-gap flux qubits (working as cavity frequency shifters and marked as qubit QL and qubit QR, respectively). Therefore, it is easy to control the detuning frequency between the frequency of the qubit QL and the frequency of the cavity a, and that of the qubit QR and the cavity b, respectively. In this way, ωa and ωb are tuned to be equal to each other, although the bare eigenfrequencies of the cavities are deviated for few MHz due to the limitation of fabrication. As shown in Fig. 1d, we can apply probing signals through port 1 (port 3) of the left (right) circulator and detect the output signals from port 2 or port 4. It yields four complex elements Sji of the scattering matrix S, where i{1,3} and j{2,4}. For example, the normalized S-matrix S21 (S43) is the reflection coefficient of the left (right) cavity, and S41 (S23) is the forward (backward) transmission coefficient. The left cavity (cavity a) supports a high-Q cavity mode a with eigenfrequency ωa/2π6.474 GHz and decay rate κa/2π7.7 MHz, and the right one (cavity b) supports a low-Q cavity mode b with the same eigenfrequency ωb/2π6.474 GHz but different decay rate κb/2π20.9 MHz (see Fig. 1e, f).

Fig. 1. Schematic diagram of the superconducting quantum chip.

Fig. 1

a Optical micrograph of the coupling qubit and the two indirectly-coupled coplanar waveguide (CPW) cavities. Two CPW cavities have the same resonant frequency of ωa/2πωb/2π6.474 GHz but different decay rates of κa/2π=7.7MHz and κb/2π=20.9 MHz. There are three tunable-gap flux qubits. The left (right) qubit QL (QR) only couples to the left (right) CPW cavity and works as a cavity frequency shifter to tune the resonant frequency of the left (right) cavity. The central qubit QC couples to the two cavities as a quantum coupler to tune the coupling strength between the two cavities. Optical micrograph (b) and scanning electron micrograph (c) of the coupling flux qubit. d Equivalent schematic diagram in which two cavities couple with each other by a quantum coupler. Each cavity is connected to a circulator for extending ports. Reflection spectra of the left high-Q cavity (e) and the right low-Q cavity (f) that display Lorentzian lineshapes detected at low probing power to keep the average photon numbers inside the cavities less than one.

The difference between the decay rates of the two CPW cavities are large enough, such that we can tune the effective coupling strength between the two cavities by changing the transition frequency of the quantum coupler to operate the system in three regimes, i.e., before, after, and in the vicinity of the exceptional point. The coupling flux qubit can be considered as a two-level artificial atom (see Fig. 1d) with transition frequency modulated by the injected external magnetic flux Φm in the main loop or Φα in the α loop45. It works as a quantum coupler to adjust the effective coupling strength between the two cavities (see section 2 of the Supplementary materials). The effective eigenfrequency of the flux qubit is Δeff=(2IpδΦ)2+[Δ(Φα)]2, where δΦ=ΦmΦ0/2 and Φ0 is the flux quantum, Ip is the persistent current in the qubit loop, Δ(Φα)/h2.0 GHz is the energy gap of the flux qubit which can be turned up by Φα. By adiabatically eliminating the degree of freedom of the coupling flux qubit, the effective coupling strength between the two cavities is g=(2Δqgagb)/(γ2+Δq2), where γ is the relaxation rate of the qubit, Δq=Δeffωa and ga=gb=71 MHz are the coupling strength between the cavity a (b) and the coupling qubit, respectively.

Observation of the exceptional point

As shown in Fig. 2, the two cavity modes a and b generate two supermodes a± with complex eigenfrequencies ω±ω0giχ±β, where ωa=ωb=ω0, χ=(κa+κb)/4, β=g2κ2, and κ=(κbκa)/4. The point g=κ corresponds to an EP at which both the eigenvalues and the eigenvectors coalesce. The applied probing power P is sufficiently low to keep the average photon numbers ni=P/ωiκi (i=a,b) inside the cavity less than one. It can be calibrated by the ac-Stark shift, which is the transition frequency shift of the qubit depending on na, i.e., ΔωL=2naga2/δ (see section 1 of the Supplementary materials). After fitting the normalized reflection spectra (S212 or S432) with the Lorentzian curves, we obtain the resonance frequencies (the real parts of) and the decay rates (the imaginary parts of) of the supermodes. As shown in Fig. 2a, b, in the strong-coupling regime g>κ, the two supermodes a± have the same damping rate 2χ but different resonance frequencies ω0g±β. The additional frequency shift from ω0 to ω0g for supermodes is induced by the interaction between the quantum coupler (the qubit QC) and the cavity modes in the strong dispersive limit, and thus causes the asymmetric frequency evolution in Fig. 2a. In the weak-coupling regime g<κ, the resonance frequencies of the two supermodes will be degenerate as ω0, while the damping rates of the two supermodes 2χiβ will be different. In our experiments, the damping rate κ is kept fixed and the effective coupling strength between the two cavity modes can be tuned by varying the bias flux δΦ in the main loop of the qubit QC, by which we can operate the system in the above three different regimes. Fig. 2c–e show the output spectra of the supermodes in three different regimes, in which the cavity modes a and b have fixed decay rates but different effective coupling strengths g/κ= 0.29, 1.04, and 3.11, respectively, where κ=2π×3.3MHz. It is shown in Fig. 2d that the supermodes are degenerate in the vicinity of the exceptional point such that gκ.

Fig. 2. Bifurcation diagram of the supermodes in the vicinity of the exceptional points.

Fig. 2

a The real part of the supermodes (frequency). b the imaginary part of the supermodes (decay rate). a, b The blue circles and crosses represent the experimental data and the dashed lines represent the theoretical curves. Normalized reflection spectra from the right low-Q cavity with different coupling strength g. (c) Before the exceptional point (g/κ=0.29) in which two supermodes overlap with each other but with different decay rates. (d) In the vicinity of the exceptional point (g/κ=1.04) in which two supermodes are degenerate. (e) After the exceptional point (g/κ=3.11) in which two supermodes are splitting. The applied probing power is low enough to keep na0.06 with a precision better than ±3 dB inside the high-Q cavity.

Nonreciprocal transmission at the few-photon level

Strong nonlinearity can result in nonreciprocal transmission of photons. In our system, the auxiliary and coupling qubits will induce third-order nonlinear terms to the cavity modes. However, since the frequency detunings between the qubits and cavity modes are large in our experiments, the coefficients of the third-order nonlinear terms are of the order of several kHz when the system is away from EP, and thus is too weak to induce nonreciprocal transmission of photons (see section 3 of the Supplementary materials). Fortunately, in the weak-coupling regime and in the vicinity of EP, the supermodes are localized in the high-Q cavity and thus the nonlinearity of the system is efficiently enhanced. This enhanced nonlinearity enables nonreciprocal transmission of photons in the weak-coupling regime (see Fig. 3c). In fact, in the experiments, we apply probing fields with average photon numbers na3 inside the high-Q cavity from port 1 (forward with a transmittance TF) and port 3 (backward with a transmittance TB) respectively. The backward output field is stronger than the forward one when the effective coupling strength is weak. However, this nonreciprocal effect is suppressed when the effective coupling strength between the two cavities is strong enough to enter the strong-coupling regime (Fig. 3d). We demonstrate the nonreciprocal transmission of photons under different coupling strengths, and the maximum isolation ratio defined as I=10log(TB/TF) can be up to 10.1 dB with na3 (see Fig. 3a). It is shown in Fig. 3e that the forward output power will increase nonlinearly with the increase of average photon numbers na in the weak-coupling regime, while it will increase linearly in the strong coupling regime.

Fig. 3. Few-photon nonreciprocity.

Fig. 3

a Isolation ratio versus the effective coupling strength g and the detuning frequency between the frequency ω of the probe field and the frequency ωo of the cavity mode with na3. b Isolation ratio versus effective coupling strength g/κ and the average photon numbers na inside the cavity with (ωω0)/2π2.7MHz. c Comparison of forward (blue, 14 in Fig. 1a) and backward (red, 32 in Fig. 1a) transmission rates in the weak-coupling regime. d Same as c in the strong-coupling regime. The transmission rates in c and d are defined as 10log(Vout2/Vin2) and have been normalized, where Vin and Vout are the input and output complex amplitudes of the probing fields. e The forward output power versus the average photon numbers na with (ωω0)/2π2.7 MHz. The red circle data are measured with effective coupling strength g/κ2.70. The blue triangle data are measured with coupling strength g/κ0.09. The sweep direction for ae is from low frequency to high frequency, from forward to backward, and from low power to high power.

We also study the isolation ratio depending on the average photon number na inside the high-Q cavity (see Fig. 3b). As we mentioned earlier, the maximum isolation ratio of around 10.1 dB appears in the weak-coupling regime for na3. In the following analysis, we focus on the variation of this maximum value as a function of na. In Fig. 3b, the isolation ratio decreases when reducing na and the nonreciprocal transmission phenomenon becomes nearly invisible when we keep na<1.5. The reason is that the nonlinearity of the cavity modes and the amplification of the nonlinearity induced by the EP phenomenon decrease simultaneously with decreasing na. Due to the strong nonlinearity of the superconducting quantum circuits at few-photon regime, there is still obvious nonreciprocal transmission effect when we keep na2 (corresponding to the input power less than −122 dBm on chip), which are much smaller than that shown in the classical PT-symmetric system11.

Discussion

We have observed the exceptional points and demonstrated nonreciprocal behavior, i.e., unidirectional transmission of microwave at a few-photon level, in a non-Hermitian superconducting coupled-cavities system with a tunable quantum coupler. When the system is in the weak-coupling regime, the isolation ratio of unidirectional transmission can achieve 10.1 dB, while almost zero isolation ratio is observed in the strong-coupling regime. Our device can be a switchable few-photon isolator by tuning the effective coupling strength between the two coupled-cavities using an external magnetic field. This switchable few-photon isolator has potential applications for the transmission of quantum information.

Methods

The System Hamiltonian

Using Jaynes-Cummings-type description, the total Hamiltonian of the system shown in Fig. 1d is

Hc=ωaaa+ωbbb+ωqσ+σ+ga(aσ+aσ+)+gb(bσ+bσ+) 1

where ga(gb) is the coupling strength between the cavity mode a (b) and the central flux qubit. In our experiment, gagb. are the resonance frequencies of the two cavities and the central qubit. By applying adiabatic elimination to the central qubit, we obtain the system’s effective Hamiltonian as

Heff=ωagiκa2aa+ωbgiκb2bbg(ab+ab)+μkerr(aaaa+bbbb), 2

where κa(κb) is the decay rate of cavity mode a (b). The effective coupling strength between the two cavity modes g and the effective self-Kerr coefficient μkerr can be written as

g=ΔqgagbΔq2+γ2, 3
μkerr=4ga(b)4Δqγγ//(Δq2+γ2)2. 4

Here, Δq=ωqωa and γ=γ///2+γφ is the central qubit’s total relaxation rate formed by its depolarizing rate γ and its dephasing rate γφ.

At zero detuning (ωa=ωb=ω0), The eigenvalues of linear terms in Eq. (2) are

ω±=ω0gi4(κa+κb)±g2(κaκb)216. 5

It can be shown that there is an exceptional point (EP) at g=κκaκb/4. The above calculation details can be found in Section 2 of the supplementary materials.

Device fabrication

The superconducting quantum chip is fabricated using a well-established multi-layer thin-film process, featuring a four-layer structure. Electron-beam lithography is employed for the Josephson junction layer, while optical lithography is used for the other layers. The fabrication process begins by depositing a 100 nm aluminum layer on a sapphire substrate, followed by etching the aluminum to form the coplanar waveguide (CPW) cavities and control lines using an aluminum etchant. Next, a 60 nm gold layer is deposited to serve as markers for e-beam alignment. For the Josephson junctions (flux qubit), we use a standard double-angle evaporation process47, depositing 40 nm of aluminum at 20° and 60 nm at −20°. Finally, air-bridges are fabricated above the transmission lines and the control lines to minimize crosstalk between them48. The primary challenge in this fabrication process lies in the numerous steps involved, where even a single error at any stage can result in the failure of the entire fabrication.

Measurement

We use a four-port vector network analyzer (VNA) to simultaneously measure the reflection spectra and transmission spectra of the system. We can obtain resonant frequencies and decay rates of the two bare cavities through the reflection spectra when the transition frequencies of the qubits are far away from the cavities. We can apply external magnetic field in α loop to make the transition frequency between the ground state and the excited state of the qubit QL (QR) resonant with the cavity a (cavity b), respectively. As shown in Fig. S2 in the supplementary materials, we scan the normalized amplitudes of the reflection coefficients (|S21| and |S43|) around the resonant frequency of the two cavities as functions of the probing frequency and flux bias δΦ in the main loop of QL and QR. The phenomena of vacuum Rabi splitting are observed. By fitting the curves of vacuum Rabi splitting, we can obtain the coupling strengths between cavities and the two qubits.

Supplementary information

Source data

Source data (454.9KB, xlsx)

Acknowledgements

We thank Long-cheng Gui for help discussions. Devices were made at the Nanofabrication Facilities at the Institute of Physics in Beijing. This work was supported by the National Natural Science Foundation of China under Grant nos. 12074117, 61833010, 62433015, the Innovation Program for Quantum Science and Technology (Grant no. 2021ZD0301800). J.Z. was supported by Laoshan Laboratory (LSKJ202200900), the Leading Scholar of Xi’an Jiaotong University, and the Innovative leading talent project of “Shuangqian plan” in Jiangxi Province. D.Z. was supported by the National Natural Science Foundation of China (Grant nos. 12204528, 92265207, T2121001), and the Micro/nano Fabrication Laboratory of Synergetic Extreme Condition User Facility (SECUF). J.S.T. was supported by the New Energy and Industrial Technology Development Organization (NEDO) under Grant no. JPNP16007. J.H. and L.M.K. were supported by the NSFC under Grant no. 12421005, the XJ-Lab key project, and the STI Program of Hunan Province under Grant no. 2023ZJ1010. F.N. was supported in part by NTT Research, JST (via the Q-LEAP program, and the Moonshot R&D Grant Number JPMJMS2061), AOARD (Grant no. FA2386-20-1-4069), and ONR.

Author contributions

Y.X.L., J.Z., and Z.P. conceived the idea. P.S., D.Z. and Z.P. designed the experiments. P.S. fabricated the device. P.S. and X.R. performed the experiments and processed the data. X.R. provided a theoretical analysis. Y.X.L., J.Z., H.J. and F.N. guided the theory. J.S.T. and L.Y. guided the experiments. J.Z. and Z.P. wrote the main text with contributions from all authors. P.S. and X.R. wrote the supplementary materials with contributions from all authors. H.D., S.L., M.C., R.H., L.M.K., and Q.Z. provided some technical support. Z.P. supervised the project.

Peer review

Peer review information

Nature Communications thanks the anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.

Data availability

The authors declare that the data supporting the findings of this study are available within the paper and its Supplementary Information files. Source data are provided. Source data are provided with this paper.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

These authors contributed equally: Pengtao Song, Xinhui Ruan.

Contributor Information

Dongning Zheng, Email: dzheng@iphy.ac.cn.

Jing Zhang, Email: zhangjing2022@xjtu.edu.cn.

Zhihui Peng, Email: zhihui.peng@hunnu.edu.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-024-54199-w.

References

  • 1.Bender, C. M. Making sense of non-Hermitian Hamiltonians. Rep. Prog. Phys.70, 947–1018 (2007). [Google Scholar]
  • 2.Christodoulides, D. N. & Yang, J. Parity-time Symmetry and Its Applications. 280 (Springer, 2018).
  • 3.Doppler, J. et al. Dynamically encircling an exceptional point for asymmetric mode switching. Nature537, 76–79 (2016). [DOI] [PubMed] [Google Scholar]
  • 4.Dembowski, C. et al. Experimental observation of the topological structure of exceptional points. Phys. Rev. Lett.86, 787–790 (2001). [DOI] [PubMed] [Google Scholar]
  • 5.Gao, T. et al. Observation of non-Hermitian degeneracies in a chaotic exciton-polariton billiard. Nature526, 554–558 (2015). [DOI] [PubMed] [Google Scholar]
  • 6.Guo, A. et al. Observation of PT-symmetry breaking in complex optical potentials. Phys. Rev. Lett.103, 093902 (2009). [DOI] [PubMed] [Google Scholar]
  • 7.Rüter, C. E. et al. Observation of parity–time symmetry in optics. Nat. Phys.6, 192–195 (2010). [Google Scholar]
  • 8.Klaiman, S., Günther, U. & Moiseyev, N. Visualization of branch points in PT-symmetric waveguides. Phys. Rev. Lett.101, 080402 (2008). [DOI] [PubMed] [Google Scholar]
  • 9.Lin, Z. et al. Unidirectional invisibility induced by PT-symmetric periodic structures. Phys. Rev. Lett.106, 213901 (2011). [DOI] [PubMed] [Google Scholar]
  • 10.Regensburger, A. et al. Parity-time synthetic photonic lattices. Nature488, 167–171 (2012). [DOI] [PubMed] [Google Scholar]
  • 11.Peng, B. et al. Parity-time-symmetric whispering-gallery microcavities. Nat. Phys.10, 394–398 (2014). [Google Scholar]
  • 12.Chang, L. et al. Parity-time symmetry and variable optical isolation in active-passive-coupled microresonators. Nat. Photon.8, 524–529 (2014). [Google Scholar]
  • 13.Peng, B. et al. Chiral modes and directional lasing at exceptional points. Proc. Natl Acad. Sci.113, 6845–6850 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Peng, B. et al. Loss-induced suppression and revival of lasing. Science346, 328–332 (2014). [DOI] [PubMed] [Google Scholar]
  • 15.Feng, L., Wong, Z. J., Ma, R.-M., Wang, Y. & Zhang, X. Single-mode laser by parity-time symmetry breaking. Science346, 972–975 (2014). [DOI] [PubMed] [Google Scholar]
  • 16.Hodaei, H., Miri, M.-A., Heinrich, M., Christodoulides, D. N. & Khajavikhan, M. Parity-time–symmetric microring lasers. Science346, 975–978 (2014). [DOI] [PubMed] [Google Scholar]
  • 17.Jing, H. et al. PT-symmetry phonon laser. Phys. Rev. Lett.113, 053604 (2014). [DOI] [PubMed] [Google Scholar]
  • 18.Zhang, G.-Q. et al. Exceptional point and cross-relaxation effect in a hybrid quantum system. PRX Quantum2, 020307 (2021). [Google Scholar]
  • 19.Wiersig, J. Enhancing the sensitivity of frequency and energy splitting detection by using exceptional points: application to microcavity sensors for single-particle detection. Phys. Rev. Lett.112, 203901 (2014). [Google Scholar]
  • 20.Liu, Z.-P. et al. Metrology with PT-symmetric cavities: Enhanced sensitivity near the PT-phase transition. Phys. Rev. Lett.117, 110802 (2016). [DOI] [PubMed] [Google Scholar]
  • 21.Chen, W., Özdemir, S. K., Zhao, G., Wiersig, J. & Yang, L. Exceptional points enhance sensing in an optical microcavity. Nature548, 192–196 (2017). [DOI] [PubMed] [Google Scholar]
  • 22.Hodaei, H. et al. Enhanced sensitivity at higher-order exceptional points. Nature548, 187–191 (2017). [DOI] [PubMed] [Google Scholar]
  • 23.Xu, H., Mason, D., Jiang, L. & Harris, G. E. Topological energy transfer in an optomechanical system with exceptional points. Nature537, 80 (2016). [DOI] [PubMed] [Google Scholar]
  • 24.Naghiloo, M., Abbasi, M., Joglekar, Y. N. & Murch, K. W. Quantum state tomography across the exceptional point in a single dissipative qubit. Nat. Phys.15, 1232–1236 (2019). [Google Scholar]
  • 25.Quijandría, F., Naether, U., Ozdemir, S. K., Nori, F. & Zueco, D. PT-symmetric circuit QED. Phys. Rev. A97, 053846 (2018). [Google Scholar]
  • 26.Wu, Y. et al. Observation of parity-time symmetry breaking in a single-spin system. Science364, 878–880 (2019). [DOI] [PubMed] [Google Scholar]
  • 27.Ding, L. et al. Experimental determination of PT-symmetric exceptional points in a single trapped ion. Phys. Rev. Lett.126, 083604 (2021). [DOI] [PubMed] [Google Scholar]
  • 28.Klauck, F. et al. Observation of PT-symmetric quantum interference. Nat. Photon.13, 883–887 (2019). [Google Scholar]
  • 29.Maraviglia, N. et al. Photonic quantum simulations of coupled PT-symmetric Hamiltonians. Phys. Rev. Res.4, 013051 (2022). [Google Scholar]
  • 30.Huang, R., Miranowicz, A., Liao, Jie-Qiao, Nori, F. & Jing, H. Nonreciprocal photon blockade. Phys. Rev. Lett.121, 153601 (2018). [DOI] [PubMed] [Google Scholar]
  • 31.Abdo, B. et al. Active protection of a superconducting qubit with an interferometric Josephson isolator. Nat. Commun.10, 3154 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Scheucher, M., Hilico, A., Will, E., Volz, J. & Rauschenbeutel, A. Quantum optical circulator controlled by a single chirally coupled atom. Science354, 2118 (2016). [DOI] [PubMed] [Google Scholar]
  • 33.Gu, X., Kockum, A. F., Miranowicz, A., Liu, Y. & Nori, F. Microwave photonics with superconducting quantum circuits. Phys. Rep.718, 1–102 (2017). 719. [Google Scholar]
  • 34.Blais, A., Girvin, S. M. & Oliver, W. D. Quantum information processing and quantum optics with circuit quantum electrodynamics. Nat. Phys.16, 247–256 (2020). [Google Scholar]
  • 35.Naaman, O., Abutaleb, M. O., Kirby, C. & Rennie, M. On-chip Josephson junction microwave switch. Appl. Phys. Lett.108, 112601 (2016). [Google Scholar]
  • 36.Pechal, M. et al. Superconducting switch for fast on-chip routing of quantum microwave fields. Phys. Rev. Appl.6, 024009 (2016). [Google Scholar]
  • 37.Wagner, A., Ranzani, L., Ribeill, G. & Ohki, T. A. Demonstration of a superconducting nanowire microwave switch. Appl. Phys. Lett.115, 172602 (2019). [Google Scholar]
  • 38.Peng, Z. H., de Graaf, S. E., Tsai, J. S. & Astafiev, O. V. Tuneable on-demand single-photon source in the microwave range. Nat. Commun.7, 12588 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Kono, S., Koshino, K., Tabuchi, Y., Noguchi, A. & Nakamura, Y. Quantum non-demolition detection of an itinerant microwave photon. Nat. Phys.14, 546 (2018). [Google Scholar]
  • 40.Hoi, I. C. et al. Generation of nonclassical microwave states using an artificial atom in 1D open space. Phys. Rev. Lett. 108, 263601 (2012). [DOI] [PubMed]
  • 41.Magnard, P. et al. Microwave quantum link between superconducting circuits housed in spatially separated cryogenic systems. Phys. Rev. Lett.125, 260502 (2020). [DOI] [PubMed] [Google Scholar]
  • 42.Chen, W., Abbasi, M., Joglekar, Y. N. & Murch, K. W. Quantum jumps in the non-Hermitian dynamics of a superconducting Qubit. Phys. Rev. Lett.127, 140504 (2021). [DOI] [PubMed] [Google Scholar]
  • 43.Gaikwad, C., Kowsari, D., Chen, W. & Murch, K. W. Observing parity-time symmetry breaking in a Josephson parametric amplifier. Phys. Rev. Res.5, L042024 (2023). [Google Scholar]
  • 44.Teixeira, W. S., Vadimov, V., Mörstedt, T., Kundu, S. & Möttönen, M. Exceptional-point-assisted entanglement, squeezing, and reset in a chain of three superconducting resonators. Phys. Rev. Res.5, 033119 (2023). [Google Scholar]
  • 45.Peng, Z. H. et al. Correlated emission lasing in harmonic oscillators coupled via a single three-level artificial atom. Phys. Rev. Lett.115, 223603 (2015). [DOI] [PubMed] [Google Scholar]
  • 46.Baust, A. et al. Tunable and switchable coupling between two superconducting resonators. Phys. Rev. B91, 014515 (2015). [Google Scholar]
  • 47.Dolan, G. J. Offset masks for lift‐off photoprocessing. Appl. Phys. Lett.31, 337–339 (1977). [Google Scholar]
  • 48.Chen, Z. et al. Fabrication and characterization of aluminum airbridges for superconducting microwave circuits. Appl. Phys. Lett.104, 052602 (2014). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Source data (454.9KB, xlsx)

Data Availability Statement

The authors declare that the data supporting the findings of this study are available within the paper and its Supplementary Information files. Source data are provided. Source data are provided with this paper.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES