Microwave spectroscopy of Majorana vortex modes
Abstract
The observation of zero-bias conductance peaks in vortex cores of certain Fe-based superconductors has sparked renewed interest in vortex-bound Majorana states. These materials are believed to be intrinsically topological in their bulk phase, thus avoiding potentially problematic interface physics encountered in superconductor-semiconductor heterostructures. However, progress toward a vortex-based topological qubit is hindered by our inability to measure the topological quantum state of a non-local vortex Majorana state, i.e., the charge of a vortex pair. In this paper, we theoretically propose a microwave-based charge parity readout of the Majorana vortex pair charge. A microwave resonator above the vortices can couple to the charge allowing for a dispersive readout of the Majorana parity. Our technique may also be used in vortices in conventional superconductors and allows one to probe the lifetime of vortex-bound quasiparticles, which is currently beyond existing scanning tunneling microscopy capabilities.
Introduction. Majorana zero modes were originally proposed within the context of vortices in a topological superconductor (SC) Volovik (1999); Read and Green (2000); Ivanov (2001); Fu and Kane (2008); Sau et al. (2010), and have since emerged as a captivating subject of study in the field of superconductivity. The recent discovery of zero-bias conductance peaks Chen et al. (2018); Wang et al. (2018); Liu et al. (2018); Chen et al. (2019); Machida et al. (2019); Kong et al. (2019); Zhu et al. (2020); Liu et al. (2020); Kong et al. (2021) in the vortex cores of certain Fe-based superconductors Shi et al. (2017); Zhang et al. (2018, 2019a); Peng et al. (2019); Li et al. (2022) has sparked renewed interest in vortex Majorana zero modes (MZMs), which are predicted to be bound in these vortices Sau et al. (2010); Qin et al. (2019); Machida and Hanaguri (2023). The inherent topological nature of vortices as excitations within the superconducting condensate gives hope that the bound states hosted by them would be less susceptible to the disorder, unlike Majorana approaches that require engineered interfaces Banerjee et al. (2023); Aghaee et al. (2023). The key motivation behind studying MZMs is their predicted non-Abelian braiding statistics and possible use in a topologically protected quantum computer Ivanov (2001); Kitaev (2003); Nayak et al. (2008); Grosfeld et al. (2011).
However, the measurement of the topological quantum state of a non-local vortex MZM remains a challenge, hindering progress toward a vortex-based topological qubit. While it is in principle possible to move the vortices and associated MZMs Tewari et al. (2007); Ma et al. (2020, 2021); Hua et al. (2021), it will be challenging to do this adiabatically for a large vortex, but at the same time fast enough to avoid quasiparticle poisoning, the timescale of which in vortices is currently unknown. Alternatively, measurement-based braiding techniques could potentially circumvent the need for moving the MZMs Bonderson et al. (2008). Non-local conductance Liu et al. (2019); Sbierski et al. (2022) and interferometric Fu and Kane (2009); Akhmerov et al. (2009) measurements have been suggested as a means to identify and control Majorana vortex modes. Nevertheless, it is important to note that a microwave-based technique would be optimal for achieving fast readout Smith et al. (2020); Razmadze et al. (2019).
In this paper, we propose a solution to the measurement problem using microwave (MW) techniques, which have been established and demonstrated as an extremely versatile tool to address electronic systems in various experiments Bretheau et al. (2013); van Woerkom et al. (2017); Tosi et al. (2019); Hays et al. (2020, 2021); Fatemi et al. (2022); Matute-Cañadas et al. (2022); Wesdorp et al. (2022). Specifically, we present a microwave-based method for MZM charge parity readout analogous to what has been proposed in different platforms Yavilberg et al. (2015); Ginossar and Grosfeld (2014); Dmytruk et al. (2015); Väyrynen et al. (2015); Yavilberg et al. (2015); Smith et al. (2020).
Our approach focuses on studying the coupling between electrons in the Fe-based superconductor and the microwave photons from a resonator positioned above it. By analyzing the frequency-dependent transmission of the resonator, we can achieve a dispersive readout of the non-local vortex Majorana state. We provide the necessary requirements for the resonator quality factor to enable the parity readout. Importantly, our technique can also be applied to vortices in conventional superconductors, offering insights into the lifetime and coherent manipulation of vortex-bound quasiparticles, surpassing the capabilities of existing scanning tunneling microscopy.
General theory of MW coupling to vortex state. The interaction between the external electromagnetic field with the charge density of the superconductor results in a MW coupling Hamiltonian
(1) |
where is the charge density operator and is the scalar potential of the external electromagnetic field. This electromagnetic field is created by a resonator, which is within a wavelength of the SC surface. Thus the field can be treated in the quasistatic approximation. The sketch of the measurement setup is shown in Fig. 1(a).
In the static field approximation, screening in the superconductor results in a decay of the field, characterized by a screening length . The scalar potential of the external electromagnetic field can be written as , where is the distance from the top surface of the superconductor and is the amplitude of the external field.
In Eq. (1), the charge density can be expressed as , where is the Nambu field operator and . In order to expand the field operator in the exact eigenbasis of the unperturbed Hamiltonian , which is given by Eq. (10) below, we define as the spinor wave function of the eigenstate with energy , and as the second quantized annihilation operators of these quasiparticles.
The eigenstates of the system exhibit a particle-hole symmetry (PHS) that is represented by an antiunitary operator . For each eigenstate with energy , there exists another eigenstate with energy . The corresponding annihilation operator satisfies . We consider energies below the SC gap and include the excited vortex-bound states (Caroli-de Gennes-Matricon states). The lowest energy state in the system is the Majorana state with energy , and its corresponding operator is given by , where and are two Majorana operators as shown in Fig. 1. We aim to read out the occupation number [or its parity, ] of this Majorana zero mode.
Expanding the Nambu spinor in terms of ,
(2) |
the MW coupling (1) can be written as
(3) | ||||
Here we introduced the matrix elements of the (surface) charge operator , e.g.,
(4) |
Because the charge operator preserves PHS, the matrix elements obey the same symmetry, encoded by the relations and .

Microwave readout of Majorana parity. In circuit quantum electrodynamics Blais et al. (2021), the MW coupling between a resonator and the superconductor allows us to read out the Majorana parity Dmytruk et al. (2015). The electromagnetic fields induced by the resonator interact with the superconductor in the manner described by the Hamiltonian , Eq. (3). This interaction influences the complex transmission coefficient that relates the output and input photonic fields of the resonator. Under the limit (see Fig. 1a) and frequency close to the cavity resonance , we find
(5) |
where is the escape rate of the cavity, and denotes the Majorana parity. We denote the parity-dependent charge-charge correlation function. In the time domain, it is given by , where is the Heaviside step function. As shown in Fig. 1a, , where and are the capacitance of the resonator and the superconductor, respectively. The resonator is coupled with capacitance to the input-output transmission line with resistance , and the effective resistance incorporates the coupling strength Göppl et al. (2008).
The interaction between the resonator and the superconductor induces transitions between the Majorana state and the vortex-bound states localized near the top surface. The correlation function contains information of these transitions and can be written as a sum (from hereon we set )
(6) | ||||
here is the transition frequency and , , and are the energies and occupation numbers of the Majorana state and the bound state . The infinitesimal level width accounts for causality and are the charge matrix elements between bound state and occupied/unoccupied () Majorana state. At low temperatures, in the absence of occupied bound states (), we obtain that for and for . The unequal charge matrix elements and transition frequencies result in different , which suggests that the MW coupling can be used to probe the Majorana occupation number .
The critical cavity Q-factor. The parity-dependent correlation function allows for the microwave readout of MZM parity based on the transmission [Eq. (5)]. For simplicity, let us only consider the first vortex-bound state on the surface. Our primary interest lies in the strong coupling regime where the coupling strength greatly exceeds the level width . In this regime, the transmission versus displays two parity-dependent peaks at . Parity readout is contingent upon the ability to distinguish peaks between different parity, which sets limitations for the cavity Q-factor . Here, we define a minimum critical cavity Q-factor required for parity distinction,
(7) |
where is the peak seperation of two parities in Fig. 1(b), is the average of peak positions, and are the shifted resonator frequencies of parity Sup , i.e., .
The Eq. (7) determines the approximate requirement for distinguishing the different parity resonances. There are two variables that affect the critical cavity Q-factor : the parity-dependent charge matrix elements and the parity-dependent transition energies associated with the Majorana energy splitting . We define the dimensionless variable and the capacitive energy .
In the resonant regime, when the resonator frequency is close to the energy of the first bound state, characterized by , where , the transmission curve of each parity exhibits two peaks of width , separated by . The parity difference causes a shift in the position of the peaks by , where is the dimensionless transition matrix element difference, and is the change in resonance frequency given by . It is important to note that these two contributions have opposite effects, which can affect the behavior of the transmission curve in this regime. By setting the shift in peak position equal to the escape rate , we obtain
(8) |
It is worth mentioning that diverges at since the peak position does not shift, and thus parity detection becomes difficult.
In the detuning regime, where the resonator frequency is significantly detuned from the first bound state’s energy (), the full expression for is more complex compared to the resonant regime (for detailed derivation, see Ref. Sup ). Nevertheless, can be approximated as:
, | (9a) | ||||
, | (9b) |
where . The two different forms highlight the parity readout based on the parity-dependence of the charge matrix element () or the transition energy (). The first form [Eq. (9a)] depends only on the change in the dimensionless charge matrix element difference as the parity-dependent factor, while the second form [Eq. (9b)] depends only on the change in transition energy as the parity-dependent factor.
Model for Fe-based superconductor. In order to estimate the feasibility of the parity readout discussed above, we will use a microscopic Hamiltonian to evaluate the transition matrix element between the Majorana state and the vortex-bound states.
We will analyze a two-band effective BdG model for an Fe-based superconductor Qin et al. (2019); Zhang et al. (2019b); Ghazaryan et al. (2020); Chiu et al. (2020); Hou and Klinovaja (2021); Barik and Sau (2022). The Hamiltonian in the Nambu basis can be represented as , where the BdG Hamiltonian is given by
(10) |
here meV represents the chemical potential, and meV is the bulk pairing gap. In our lattice model, with , where and represent the Pauli matrices that account for the orbital and spin degrees of freedom, respectively Zhang et al. (2019b). In this basis, , where represents the Pauli matrix that accounts for the particle-hole degrees of freedom and denotes complex conjugation. In our numerical simulation, we set meV and nm (the lattice constant), , and so that the system is in the topological phase Zhang et al. (2019b); Hou and Klinovaja (2021) which can have vortex Majorana zero modes.
In the context of our model, we consider the s-wave superconducting pairing potential in the presence of vortices that extend along the z-axis. For a single vortex centered at the origin, the pairing term can be expressed as Blatter et al. (1996)
(11) |
where nm represents the characteristic radius of the vortex and .
In our specific model (shown in Fig. 1a), we consider the presence of two vortices, each hosting a pair of MBSs within the Fe-based superconductor. Assuming the vortices are far apart, we can approximate the pairing term as follows,
(12) |
where and correspond to the respective locations of the two vortices.
The two-vortex pairing term and the BdG Hamiltonian exhibit a symmetry represented by . This operator is characterized by a rotation around the z-axis, with respect to the midpoint of the two vortices, taken here as the origin. Its action on a function is given by . The symmetry operator has eigenvalues . The manifestation of this symmetry results in the observation of double degeneracy in the system’s spectrum, as evident in the inset of Fig. 2. The operator commutes with the Hamiltonian (1), establishing a selection rule that governs the allowed MW transitions within the system. According to the selection rule, transitions within the system can only occur between states that share the same eigenvalues of . Since the PHS operator changes the eigenvalue of , at least one of the transition matrix elements vanishes.
However, the presence of random disorder in realistic conditions disrupts the symmetry, resulting in the elimination of the double degeneracy in the spectrum (Fig. 2). Consequently, this compromises the strict adherence to the selection rule. None of the transition matrix elements are generally zero (Fig. 3). Thus, in realistic experimental settings, the selection rule is not rigorously maintained.

Numerical studies of two-vortex systems. We employ a numerical approach to investigate a two-vortex system. To perform the numerical analysis, we discretize the Hamiltonian given by Eq. (10) and utilize the Kwant package Groth et al. (2014) in Python to implement and solve the corresponding tight-binding model. The system under consideration is a cuboid with dimensions 500 nm 250 nm 25 nm (refer to Fig. 1a for an illustration), discretized with a lattice constant nm.
We utilize the results obtained in Ref. Sup to calculate the screened electric potential of two vortices Blatter et al. (1996), which is then included in the real space version of the Hamiltonian in Eq. (10), similar to the way the chemical potential is incorporated. We take the screening length as one lattice constant. Our investigation encompasses both clean and disordered systems. To model the disorder, we introduce a position-dependent random potential into the Hamiltonian. The disorder potential follows a normal distribution, with the standard deviation of this distribution matching the gap . The spectrums and charge matrix elements acquired through numerical computations are depicted in Fig. 2 and Fig. 3.
Discussion. We showed that a microwave coupling enables the parity readout of a non-local Majorana zero mode hosted in a vortex pair. We quantified the sensitivity of the readout by defining a critical cavity Q-factor , Eqs. (8)-(9b), required of the resonant cavity coupled to the vortices. To estimate a typical value of , let us consider a resonant frequency 5 GHz (much below a typical superconducting gap ) and effective capacitance F of a typical coplanar waveguide resonator Göppl et al. (2008). In our simulation, we find that the MZM energy for a system with a large vortex separation can be neglected while the first excited state is approximately at (see Fig. 2), implying the system is in the detuning regime. The relevant charge matrix elements are numerically estimated to be and , the ratio of which is shown in Fig. 3. In this case, Eq. (9a) gives the required critical cavity Q-factor , which is close to state-of-the-art experimental conditions Noguchi et al. (2019). Below distance , the system is still in the detuning regime of Eq. (9a). There, , well within reach of the experiments.
Our method offers a compelling approach to measuring the non-Abelian nature of Majorana zero modes. By employing two resonators to measure quantities and , we can effectively measure two non-commuting parities of MZMs. Through monitoring these observables Vool et al. (2014); Hays et al. (2018), we can estimate quasi-particle poisoning time and MZM hybridization . Additionally, incorporating a third resonator to measure and an ancillary pair of MZMs, would enable measurement-based braiding Bonderson et al. (2008); Liu et al. (2019); Karzig et al. (2017) within timescales shorter than the quasi-particle poisoning time, when the total parity is conserved. Alternatively, braiding can be achieved through time-dependent control of MZM hybridization in a non-topologically protected manner Trif and Simon (2019). Thus, the resonator-based approach not only allows one to measure the essential quasi-particle poisoning time but also enables one to demonstrate the non-Abelian characteristics of vortex-based MZMs, thus holding significant promise for advancing topological quantum computing and related technologies.

Acknowledgements.
Acknowledgments. We thank Yong Chen, Valla Fatemi, Leonid Glazman, Mingi Kim, and Lingyuan Kong for valuable discussions. This work was initiated at Aspen Center for Physics, which is supported by National Science Foundation grant PHY-1607611. This material is based upon work supported by the Office of the Under Secretary of Defense for Research and Engineering under award number FA9550-22-1-0354.References
- Volovik (1999) G. E. Volovik, “Fermion zero modes on vortices in chiral superconductors,” Soviet Journal of Experimental and Theoretical Physics Letters 70, 609–614 (1999).
- Read and Green (2000) N. Read and Dmitry Green, “Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum Hall effect,” Physical Review B 61, 10267–10297 (2000), arXiv:cond-mat/9906453 [cond-mat.mes-hall] .
- Ivanov (2001) D. A. Ivanov, “Non-Abelian Statistics of Half-Quantum Vortices in p-Wave Superconductors,” Phys. Rev. Lett. 86, 268–271 (2001).
- Fu and Kane (2008) L. Fu and C. L. Kane, “Superconducting Proximity Effect and Majorana Fermions at the Surface of a Topological Insulator,” Phys. Rev. Lett. 100, 096407 (2008).
- Sau et al. (2010) Jay D. Sau, Roman M. Lutchyn, Sumanta Tewari, and S. Das Sarma, “Generic New Platform for Topological Quantum Computation Using Semiconductor Heterostructures,” Phys. Rev. Lett. 104, 040502 (2010).
- Chen et al. (2018) Mingyang Chen, Xiaoyu Chen, Huan Yang, Zengyi Du, Xiyu Zhu, Enyu Wang, and Hai-Hu Wen, “Discrete energy levels of Caroli-de Gennes-Matricon states in quantum limit in FeTe0.55Se0.45,” Nature communications 9, 1–7 (2018).
- Wang et al. (2018) Dongfei Wang, Lingyuan Kong, Peng Fan, Hui Chen, Shiyu Zhu, Wenyao Liu, Lu Cao, Yujie Sun, Shixuan Du, John Schneeloch, Ruidan Zhong, Genda Gu, Liang Fu, Hong Ding, and Hong-Jun Gao, “Evidence for Majorana bound states in an iron-based superconductor,” Science 362, 333–335 (2018).
- Liu et al. (2018) Qin Liu, Chen Chen, Tong Zhang, Rui Peng, Ya-Jun Yan, Chen-Hao-Ping Wen, Xia Lou, Yu-Long Huang, Jin-Peng Tian, Xiao-Li Dong, Guang-Wei Wang, Wei-Cheng Bao, Qiang-Hua Wang, Zhi-Ping Yin, Zhong-Xian Zhao, and Dong-Lai Feng, “Robust and Clean Majorana Zero Mode in the Vortex Core of High-Temperature Superconductor ,” Phys. Rev. X 8, 041056 (2018).
- Chen et al. (2019) C. Chen, Q. Liu, T. Z. Zhang, D. Li, P. P. Shen, X. L. Dong, Z.-X. Zhao, T. Zhang, and D. L. Feng, “Quantized Conductance of Majorana Zero Mode in the Vortex of the Topological Superconductor ,” Chinese Physics Letters 36, 057403 (2019).
- Machida et al. (2019) T. Machida, Y. Sun, S. Pyon, S. Takeda, Y. Kohsaka, T. Hanaguri, T. Sasagawa, and T. Tamegai, “Zero-energy vortex bound state in the superconducting topological surface state of Fe(Se,Te),” Nature Materials 18, 811–815 (2019).
- Kong et al. (2019) Lingyuan Kong, Shiyu Zhu, Michał Papaj, Hui Chen, Lu Cao, Hiroki Isobe, Yuqing Xing, Wenyao Liu, Dongfei Wang, Peng Fan, Yujie Sun, Shixuan Du, John Schneeloch, Ruidan Zhong, Genda Gu, Liang Fu, Hong-Jun Gao, and Hong Ding, “Half-integer level shift of vortex bound states in an iron-based superconductor,” Nature Physics 15, 1181–1187 (2019).
- Zhu et al. (2020) Shiyu Zhu, Lingyuan Kong, Lu Cao, Hui Chen, Michał Papaj, Shixuan Du, Yuqing Xing, Wenyao Liu, Dongfei Wang, Chengmin Shen, Fazhi Yang, John Schneeloch, Ruidan Zhong, Genda Gu, Liang Fu, Yu-Yang Zhang, Hong Ding, and Hong-Jun Gao, “Nearly quantized conductance plateau of vortex zero mode in an iron-based superconductor,” Science 367, 189–192 (2020).
- Liu et al. (2020) Wenyao Liu, Lu Cao, Shiyu Zhu, Lingyuan Kong, Guangwei Wang, Michał Papaj, Peng Zhang, Ya-Bin Liu, Hui Chen, Geng Li, Fazhi Yang, Takeshi Kondo, Shixuan Du, Guang-Han Cao, Shik Shin, Liang Fu, Zhiping Yin, Hong-Jun Gao, and Hong Ding, “A new Majorana platform in an Fe-As bilayer superconductor,” Nature Communications 11, 5688 (2020).
- Kong et al. (2021) Lingyuan Kong, Lu Cao, Shiyu Zhu, Michał Papaj, Guangyang Dai, Geng Li, Peng Fan, Wenyao Liu, Fazhi Yang, Xiancheng Wang, Shixuan Du, Changqing Jin, Liang Fu, Hong-Jun Gao, and Hong Ding, “Majorana zero modes in impurity-assisted vortex of LiFeAs superconductor,” Nature Communications 12, 4146 (2021).
- Shi et al. (2017) Xun Shi, Zhi-Qing Han, Pierre Richard, Xian-Xin Wu, Xi-Liang Peng, Tian Qian, Shan-Cai Wang, Jiang-Ping Hu, Yu-Jie Sun, and Hong Ding, “FeTe1-xSex monolayer films: towards the realization of high-temperature connate topological superconductivity,” Science Bulletin 62, 503–507 (2017).
- Zhang et al. (2018) Peng Zhang, Koichiro Yaji, Takahiro Hashimoto, Yuichi Ota, Takeshi Kondo, Kozo Okazaki, Zhijun Wang, Jinsheng Wen, G. D. Gu, Hong Ding, and Shik Shin, “Observation of topological superconductivity on the surface of an iron-based superconductor,” Science 360, 182–186 (2018).
- Zhang et al. (2019a) Peng Zhang et al., “Multiple topological states in iron-based superconductors,” Nature Physics 15, 41–47 (2019a).
- Peng et al. (2019) X.-L. Peng, Y. Li, X.-X. Wu, H.-B. Deng, X. Shi, W.-H. Fan, M. Li, Y.-B. Huang, T. Qian, P. Richard, J.-P. Hu, S.-H. Pan, H.-Q. Mao, Y.-J. Sun, and H. Ding, “Observation of topological transition in high- superconducting monolayer films on ,” Phys. Rev. B 100, 155134 (2019).
- Li et al. (2022) Geng Li, Shiyu Zhu, Peng Fan, Lu Cao, and Hong-Jun Gao, “Exploring Majorana zero modes in iron-based superconductors,” Chinese Physics B 31, 080301 (2022).
- Qin et al. (2019) Shengshan Qin, Lunhui Hu, Xianxin Wu, Xia Dai, Chen Fang, Fu-Chun Zhang, and Jiangping Hu, “Topological vortex phase transitions in iron-based superconductors,” Science Bulletin 64, 1207–1214 (2019).
- Machida and Hanaguri (2023) Tadashi Machida and Tetsuo Hanaguri, “Searching for Majorana quasiparticles at vortex cores in iron-based superconductors,” Progress of Theoretical and Experimental Physics , ptad084 (2023).
- Banerjee et al. (2023) A. Banerjee, O. Lesser, M. A. Rahman, H.-R. Wang, M.-R. Li, A. Kringhøj, A. M. Whiticar, A. C. C. Drachmann, C. Thomas, T. Wang, M. J. Manfra, E. Berg, Y. Oreg, Ady Stern, and C. M. Marcus, “Signatures of a topological phase transition in a planar Josephson junction,” Phys. Rev. B 107, 245304 (2023).
- Aghaee et al. (2023) Morteza Aghaee et al. (Microsoft Quantum), “InAs-Al hybrid devices passing the topological gap protocol,” Phys. Rev. B 107, 245423 (2023).
- Kitaev (2003) A. Y. Kitaev, “Fault-tolerant quantum computation by anyons,” Annals of Physics 303, 2–30 (2003), quant-ph/9707021 .
- Nayak et al. (2008) Chetan Nayak, Steven H. Simon, Ady Stern, Michael Freedman, and Sankar Das Sarma, “Non-abelian anyons and topological quantum computation,” Rev. Mod. Phys. 80, 1083–1159 (2008).
- Grosfeld et al. (2011) Eytan Grosfeld, Babak Seradjeh, and Smitha Vishveshwara, “Proposed Aharonov-Casher interference measurement of non-Abelian vortices in chiral p-wave superconductors,” Physical Review B 83 (2011), 10.1103/physrevb.83.104513.
- Tewari et al. (2007) Sumanta Tewari, S. Das Sarma, Chetan Nayak, Chuanwei Zhang, and P. Zoller, “Quantum Computation using Vortices and Majorana Zero Modes of a Superfluid of Fermionic Cold Atoms,” Phys. Rev. Lett. 98, 010506 (2007).
- Ma et al. (2020) X. Ma, C. J. O. Reichhardt, and C. Reichhardt, “Braiding Majorana fermions and creating quantum logic gates with vortices on a periodic pinning structure,” Phys. Rev. B 101, 024514 (2020).
- Ma et al. (2021) Hai-Yang Ma, Dandan Guan, Shiyong Wang, Yaoyi Li, Canhua Liu, Hao Zheng, and Jin-Feng Jia, “Braiding Majorana zero mode in an electrically controllable way,” Journal of Physics D Applied Physics 54, 424003 (2021).
- Hua et al. (2021) Chengyun Hua, Gábor B. Halász, Eugene Dumitrescu, Matthew Brahlek, and Benjamin Lawrie, “Optical vortex manipulation for topological quantum computation,” Phys. Rev. B 104, 104501 (2021).
- Bonderson et al. (2008) Parsa Bonderson, Michael Freedman, and Chetan Nayak, “Measurement-Only Topological Quantum Computation,” Phys. Rev. Lett. 101, 010501 (2008).
- Liu et al. (2019) Chun-Xiao Liu, Dong E. Liu, Fu-Chun Zhang, and Ching-Kai Chiu, “Protocol for Reading Out Majorana Vortex Qubits and Testing Non-Abelian Statistics,” Phys. Rev. Appl. 12, 054035 (2019).
- Sbierski et al. (2022) Björn Sbierski, Max Geier, An-Ping Li, Matthew Brahlek, Robert G. Moore, and Joel E. Moore, “Identifying Majorana vortex modes via nonlocal transport,” Physical Review B 106 (2022), 10.1103/physrevb.106.035413.
- Fu and Kane (2009) Liang Fu and C. L. Kane, “Probing Neutral Majorana Fermion Edge Modes with Charge Transport,” Phys. Rev. Lett. 102, 216403 (2009).
- Akhmerov et al. (2009) A. R. Akhmerov, Johan Nilsson, and C. W. J. Beenakker, “Electrically Detected Interferometry of Majorana Fermions in a Topological Insulator,” Phys. Rev. Lett. 102, 216404 (2009).
- Smith et al. (2020) Thomas B. Smith, Maja C. Cassidy, David J. Reilly, Stephen D. Bartlett, and Arne L. Grimsmo, “Dispersive Readout of Majorana Qubits,” PRX Quantum 1, 020313 (2020).
- Razmadze et al. (2019) Davydas Razmadze, Deividas Sabonis, Filip K. Malinowski, Gerbold C. Ménard, Sebastian Pauka, Hung Nguyen, David M.T. van Zanten, Eoin C.T. O′Farrell, Judith Suter, Peter Krogstrup, Ferdinand Kuemmeth, and Charles M. Marcus, “Radio-Frequency Methods for Majorana-Based Quantum Devices: Fast Charge Sensing and Phase-Diagram Mapping,” Phys. Rev. Appl. 11, 064011 (2019).
- Bretheau et al. (2013) L. Bretheau, Ç. Ö. Girit, H. Pothier, D. Esteve, and C. Urbina, “Exciting Andreev pairs in a superconducting atomic contact,” Nature 499, 312–315 (2013).
- van Woerkom et al. (2017) David J. van Woerkom, Alex Proutski, Bernard van Heck, Daniël Bouman, Jukka I. Väyrynen, Leonid I. Glazman, Peter Krogstrup, Jesper Nygård, Leo P. Kouwenhoven, and Attila Geresdi, “Microwave spectroscopy of spinful Andreev bound states in ballistic semiconductor Josephson junctions,” Nat. Phys. 13, 876–881 (2017).
- Tosi et al. (2019) L. Tosi, C. Metzger, M. F. Goffman, C. Urbina, H. Pothier, Sunghun Park, A. Levy Yeyati, J. Nygârd, and P. Krogstrup, “Spin-Orbit Splitting of Andreev States Revealed by Microwave Spectroscopy,” Phys. Rev. X 9, 011010 (2019).
- Hays et al. (2020) M Hays, V Fatemi, K Serniak, D Bouman, S Diamond, G de Lange, P Krogstrup, J Nygård, A Geresdi, and MH Devoret, “Continuous monitoring of a trapped superconducting spin,” Nature Physics 16, 1103–1107 (2020).
- Hays et al. (2021) M Hays, V Fatemi, D Bouman, J Cerrillo, S Diamond, K Serniak, T Connolly, P Krogstrup, J Nygård, A Levy Yeyati, et al., “Coherent manipulation of an Andreev spin qubit,” Science 373, 430–433 (2021).
- Fatemi et al. (2022) V Fatemi, PD Kurilovich, M Hays, D Bouman, T Connolly, S Diamond, NE Frattini, VD Kurilovich, P Krogstrup, J Nygard, et al., “Microwave Susceptibility Observation of Interacting Many-Body Andreev States,” Phys. Rev. Lett. 129, 227701 (2022).
- Matute-Cañadas et al. (2022) F. J. Matute-Cañadas, C. Metzger, Sunghun Park, L. Tosi, P. Krogstrup, J. Nygård, M. F. Goffman, C. Urbina, H. Pothier, and A. Levy Yeyati, “Signatures of Interactions in the Andreev Spectrum of Nanowire Josephson Junctions,” Phys. Rev. Lett. 128, 197702 (2022).
- Wesdorp et al. (2022) J. J. Wesdorp, F. J. Matute-Caňadas, A. Vaartjes, L. Grünhaupt, T. Laeven, S. Roelofs, L. J. Splitthoff, M. Pita-Vidal, A. Bargerbos, D. J. van Woerkom, P. Krogstrup, L. P. Kouwenhoven, C. K. Andersen, A. Levy Yeyati, B. van Heck, and G. de Lange, “Microwave spectroscopy of interacting Andreev spins,” (2022), 10.48550/ARXIV.2208.11198.
- Yavilberg et al. (2015) Konstantin Yavilberg, Eran Ginossar, and Eytan Grosfeld, “Fermion parity measurement and control in Majorana circuit quantum electrodynamics,” Phys. Rev. B 92, 075143 (2015).
- Ginossar and Grosfeld (2014) E. Ginossar and E. Grosfeld, “Microwave transitions as a signature of coherent parity mixing effects in the Majorana-transmon qubit,” Nat. Commun. 5, 4772 (2014), arXiv:1307.1159 [cond-mat.mes-hall] .
- Dmytruk et al. (2015) Olesia Dmytruk, Mircea Trif, and Pascal Simon, “Cavity quantum electrodynamics with mesoscopic topological superconductors,” Phys. Rev. B 92, 245432 (2015).
- Väyrynen et al. (2015) Jukka I. Väyrynen, Gianluca Rastelli, Wolfgang Belzig, and Leonid I. Glazman, “Microwave signatures of Majorana states in a topological Josephson junction,” Phys. Rev. B 92, 134508 (2015).
- Blais et al. (2021) Alexandre Blais, Arne L. Grimsmo, S. M. Girvin, and Andreas Wallraff, “Circuit quantum electrodynamics,” Rev. Mod. Phys. 93, 025005 (2021).
- Göppl et al. (2008) M. Göppl, A. Fragner, M. Baur, R. Bianchetti, S. Filipp, J. M. Fink, P. J. Leek, G. Puebla, L. Steffen, and A. Wallraff, “Coplanar waveguide resonators for circuit quantum electrodynamics,” Journal of Applied Physics 104 (2008), 10.1063/1.3010859, 113904.
- (52) See supplementary materials for details.
- Zhang et al. (2019b) Rui-Xing Zhang, William S. Cole, and S. Das Sarma, “Helical Hinge Majorana Modes in Iron-Based Superconductors,” Phys. Rev. Lett. 122, 187001 (2019b).
- Ghazaryan et al. (2020) Areg Ghazaryan, P. L. S. Lopes, Pavan Hosur, Matthew J. Gilbert, and Pouyan Ghaemi, “Effect of Zeeman coupling on the Majorana vortex modes in iron-based topological superconductors,” Phys. Rev. B 101, 020504 (2020).
- Chiu et al. (2020) Ching-Kai Chiu, T. Machida, Yingyi Huang, T. Hanaguri, and Fu-Chun Zhang, “Scalable Majorana vortex modes in iron-based superconductors,” Science Advances 6, eaay0443 (2020).
- Hou and Klinovaja (2021) Zhe Hou and Jelena Klinovaja, “Zero-energy Andreev bound states in iron-based superconductor Fe(Te,Se),” (2021).
- Barik and Sau (2022) Tamoghna Barik and Jay D. Sau, “Signatures of nontopological patches on the surface of topological insulators,” Phys. Rev. B 105, 035128 (2022).
- Blatter et al. (1996) Gianni Blatter, Mikhail Feigel’man, Vadim Geshkenbein, Anatoli Larkin, and Anne van Otterlo, “Electrostatics of Vortices in Type-II Superconductors,” Phys. Rev. Lett. 77, 566–569 (1996).
- Groth et al. (2014) Christoph W Groth, Michael Wimmer, Anton R Akhmerov, and Xavier Waintal, “Kwant: a software package for quantum transport,” New Journal of Physics 16, 063065 (2014).
- Noguchi et al. (2019) Takashi Noguchi, Agnes Dominjon, Matthias Kroug, Satoru Mima, and Chiko Otani, “Characteristics of Very High Q Nb Superconducting Resonators for Microwave Kinetic Inductance Detectors,” IEEE Transactions on Applied Superconductivity 29, 1–5 (2019).
- Vool et al. (2014) U. Vool, I. M. Pop, K. Sliwa, B. Abdo, C. Wang, T. Brecht, Y. Y. Gao, S. Shankar, M. Hatridge, G. Catelani, M. Mirrahimi, L. Frunzio, R. J. Schoelkopf, L. I. Glazman, and M. H. Devoret, “Non-Poissonian Quantum Jumps of a Fluxonium Qubit due to Quasiparticle Excitations,” Phys. Rev. Lett. 113, 247001 (2014).
- Hays et al. (2018) M. Hays, G. de Lange, K. Serniak, D. J. van Woerkom, D. Bouman, P. Krogstrup, J. Nygård, A. Geresdi, and M. H. Devoret, “Direct Microwave Measurement of Andreev-Bound-State Dynamics in a Semiconductor-Nanowire Josephson Junction,” Phys. Rev. Lett. 121, 047001 (2018).
- Karzig et al. (2017) T. Karzig, C. Knapp, R. M. Lutchyn, P. Bonderson, M. B. Hastings, C. Nayak, J. Alicea, K. Flensberg, S. Plugge, Y. Oreg, et al., “Scalable designs for quasiparticle-poisoning-protected topological quantum computation with Majorana zero modes,” Phys. Rev. B 95, 235305 (2017), arXiv:1610.05289 [cond-mat.mes-hall] .
- Trif and Simon (2019) Mircea Trif and Pascal Simon, “Braiding of Majorana Fermions in a Cavity,” Phys. Rev. Lett. 122, 236803 (2019).
- Lubashevsky et al. (2012) Y. Lubashevsky, E. Lahoud, K. Chashka, D. Podolsky, and A. Kanigel, “Shallow pockets and very strong coupling superconductivity in FeSexTe1-x,” Nature Physics 8, 309–312 (2012).
- Rinott et al. (2017) Shahar Rinott, K. B. Chashka, Amit Ribak, Emile D. L. Rienks, Amina Taleb-Ibrahimi, Patrick Le Fevre, François Bertran, Mohit Randeria, and Amit Kanigel, “Tuning across the BCS-BEC crossover in the multiband superconductor Fe1+ySexTe1-x: An angle-resolved photoemission study,” Science Advances 3, e1602372 (2017), https://www.science.org/doi/pdf/10.1126/sciadv.1602372 .
- Xu et al. (2023) Chang Xu, Ka Ho Wong, Eric Mascot, and Dirk K. Morr, “Competing topological superconducting phases in ,” Phys. Rev. B 107, 214514 (2023).
Supplementary materials on
“Microwave spectroscopy of Majorana vortex modes”
Zhibo Ren1, Justin Copenhaver1,2, Leonid Rokhinson1, Jukka I. Väyrynen1,
1 Department of Physics and Astronomy, Purdue University, West Lafayette, Indiana 47907, USA
2 Department of Physics, University of Colorado, Boulder, CO 80309, USA
These supplementary materials contain details about the critical cavity Q-factor and screened potential of a vortex.
S1 Derivation of the critical cavity Q-factor
In the strong coupling limit, where the MW coupling strength greatly exceeds the vortex bound state level width , we can approximate the correlation function in Eq. (6) by neglecting its imaginary part. The approximation leads to an expression for the correlation function given by (take )
(S1) |
where is the Majorana occupation number and its parity. The resonator transmission, Eq. (5) of the main text, then takes the form
(S2) |
We can read off the modified resonance frequencies from the equation,
(S3) |
Next, we solve this equation in both the resonant limit and the detuning limit, and find that there are always 3 solutions, we denote them as , and .
In the resonant region , these solutions are given by
(S4) |
The first solution is negative, so the peak with the lowest positive frequency of the transmission is located at , with a peak width approximately equal to . The shift of the peak positions, denoted as , is given by
(S5) |
Approximate to the zero-order of , the average of peak positions . The critical cavity Q-factor is achieved when the shift in peak position equals the escape rate, . This lead to the expression for ,
(S6) |
giving Eq. (8) of the main text.
In the detuning region , the Eq. (S3) has 3 solutions given by
(S7) |
Similar to the resonant region, the peak with the lowest positive frequency of the transmission is located at , and the peak width is now doubled to . The shift of the peak positions is given by
(S8) |
Approximate to the zero-order of , the average of peak positions The critical cavity Q-factor is achieved when the shift in peak position equals the escape rate, .
(S9) |
S2 Screening of a vortex
Here we model the screened potential of a vortex based on 3D parabolic bulk bands Blatter et al. (1996); Lubashevsky et al. (2012); Rinott et al. (2017); Xu et al. (2023). The opening of a gap in the spectrum results in a displacement of the carrier density by , corresponding to a (non-screened) charge density . For a single vortex of size with order parameter given by Eq. (11),
(S10) |
with the radial distance from the vortex core and Blatter et al. (1996), where is the electron density and is the density of states at chemical potential . With two vortices separated by a distance , the charge density contains contributions from each vortex.
To account for electric screening due to bulk carriers Lubashevsky et al. (2012); Rinott et al. (2017) in the superconductor, we use a Thomas-Fermi approximation and solve the screened Poisson equation for the screened electric potential Blatter et al. (1996):
(S11) |
where is the screening length. In our numerical simulation, we take to be one lattice constant, nm. The Green’s function corresponding to Eq. (S11) is , therefore, on the lattice we have
(S12) |
After performing this calculation with a two-vortex source term, we then use in Hamiltonian given by Eq. (10), by setting .
