Proposal for a nonadiabatic geometric gate with an Andreev spin qubit
Abstract
We study a hybrid structure of a ferromagnetic insulator and a superconductor connected by a weak link, which accommodates Andreev bound states whose spin degeneracy is lifted due to the exchange interaction with the ferromagnet. The resultant spin-resolved energy levels realize a two-state quantum system, provided that a single electron is trapped in the bound state, i.e., an Andreev spin qubit. The qubit state can be manipulated by controlling the magnetization dynamics of the ferromagnet, which mediates the coupling between external fields and the qubit. In particular, our hybrid structure provides a simple platform to manipulate and control the spin qubit using spintronic techniques. By employing a modified Hahn spin echo protocol for the magnetization dynamics, we show that our Andreev spin qubit can realize a nonadiabatic geometric gate.
I Introduction
Motivated by the prospect of scalable device fabrication and circuit design, the search for two-state quantum systems or qubits in solid-state systems has been a major experimental and theoretical endeavor Burkard et al. (2023); Vion et al. (2002); *task2; *task3; *task4; *task5. Among all, Andreev physics-based qubits present a particularly promising route. In the superconducting state at low temperatures, the Fermi-level degrees of freedom are frozen out. Therefore, most of the dissipative mechanisms are eliminated so that the qubit may exhibit long coherence times Bocko et al. (1997); *Ketterson2; Hays et al. (2021); Janvier et al. (2015). In this paper, we study a hybrid structure of a ferromagnetic insulator (FI) connected by a weak link (a normal region N) to an -wave superconductor (S) that accommodates spin-resolved Andreev bound states (ABS) in the weak link [see Fig. 1(a)]. We show that a pair of spin-resolved ABS, when occupied by an electron, creates an Andreev spin qubit (ASQ). On account of a large level spacing in our ASQ, which is due to the exchange field of the FI, the external fields controlling the FI magnetization are strongly coupled to the qubit spin. Due to the strong coupling, we expect that spin-flipping errors originating from weak random fields in the environment, such as in the case of hyperfine coupling to nuclei Khaetskii et al. (2002); *Loss22; *Loss23, will be suppressed.
A hallmark feature of the ABS formed in a Josephson junction is that the occupied levels modify a supercurrent across the junction Chtchelkatchev and Nazarov (2003); Nichele et al. (2020) so that transitions between Andreev levels can be detected by means of transport measurement. This is the physical basis of Andreev level qubits Janvier et al. (2015); Lantz et al. (2002); *Zaz2; *Zaz3. It was also observed that in spin active weak links, e.g., with spin-orbit coupling Chtchelkatchev and Nazarov (2003); Hays et al. (2021), or magnetic impurity Bratus’ et al. (1997); *ASQ2, the spin degeneracy of the levels can be lifted, so that the supercurrent flow would depend on the spin state of the electron trapped in the bound state. In our ASQ, the exchange field of the FI breaks the spin degeneracy of the levels, so that quantum information can be stored in the spin state of the electron Wendin and Shumeiko (2021). Similarly, to detect or read out the qubit state, we study the superconducting current affected by the occupied spin-resolved levels.

Generically, the unwanted qubit coupling to the environment and inaccurate external control can reduce the qubit state coherence time, which hinders the experimental implementation of scalable qubit designs Burkard et al. (2023); Vion et al. (2002); *task2; *task3; *task4. The qubits with gates based on the geometric phases may have the inherently fault-tolerant advantage due to the fact that the geometric gates depend only on some global geometric features of the evolution but independent of evolution details, so that they can be robust against control errors Xiang-Bin and Keiji (2001a); *PRLM2; De Chiara and Palma (2003); *GG2; *GG3; *GG4; Zhao et al. (2017). Moreover, utilizing nonadiabatic gates with high operation speed may reduce exposure time to the environment, rendering a high-fidelity gate Zhang et al. (2020). Here, we will show that our hybrid structure is a natural platform to realize a nonadiabatic geometric gate.
II The hybrid structure with spin-resolved ABS
Consider the FI-N-S hybrid structure with the following Bogoliubov–de Gennes Hamiltonian
(1) |
where , is the chemical potential, is the magnetic exchange field, and is the superconducting pair potential, both written in terms of the Heaviside step function. The scalar potential models the FI electron gap where , and parametrizes an interface barrier potential where is the Fermi velocity. The unit vector gives the time-dependent direction of the magnetization, and are Pauli matrices operating in spin space.
Our goal is to find the subgap () energy spectrum of bound states, which are confined in the normal region. Imposing a condition for constructive quantum interference in a McMillan-Rowell process for the electrons—four times crossing N with two Andreev conversions as well as two reflections from FI, once as electron and once as hole Eschrig (2018)—results in a transcendental equation
(2) |
whose solutions are the discrete bound states energy levels (measured from ). Here, is specified with respect to the FI magnetization direction, , with being the spin-dependent phase acquired by the electron upon reflection from the FIN interface, , where Eschrig (2018). where is the normal region thickness. Since for subgap energies the S supports only Cooper pair tunneling into the S, the incident electrons from the N region on the clean N-S interface reflect back as a hole rat (experiencing an Andreev reflection), where the phase change associated with the Andreev reflection is given by .
When is smaller than the superconducting coherence length and , Eq. (2) can admit a single spin-degenerate positive-energy solution whose degeneracy is removed by increasing [see Fig. 1(b)]. We point out that an oxide layer formed at the FIN interface or a tunable mismatch between the electronic properties across the interface create an effective ultrathin insulating layer that can greatly affect the probability of electron and hole evanescent penetration into the FI region. To capture the essential effect of this insulating layer, we use an interface barrier potential with strength , which can be utilized to tune the effective exchange interaction experienced by electrons and holes [see Fig. 1(c)]. Moreover, the number of ABS found in a clean N region depends on its thickness, that is, by increasing the ABS levels are pushed towards the middle of the gap at , so that more energy levels begin to appear [see Fig. 1(d)].
III The Andreev spin qubit
Let us consider a hybrid structure with only two positive-energy spin-resolved ABS , and ignore the continuum of positive energy levels above the S gap. We define the ground state of the system when these ABS are unoccupied and measure energies with respect to . The energies then are for the ground, and for two spin-, and for excited state. The ground and excited states have even (electron number) parity. The odd parity states, on the other hand, correspond to a single electron in the system realizing an ASQ. ABS wave functions with energies may have different spatial components. However, due to the tunneling treatment of the magnetic insulator, we assume that the ABS wave functions differ mainly in their spin character, with the orbital components being essentially identical. The effective ASQ Hamiltonian written in terms of the instantaneous quantization axis reads
(3) |
where . It is clear that the timescale for the coherent manipulation of the qubit, , is set by the lifetime of the odd-parity sector, . In the temperature regime considered in this paper, , the odd-parity sector can be prepared, for example, by microwave irradiation Lundin et al. (1999); Chtchelkatchev and Nazarov (2003), electron injection by means of voltage gates Wendin et al. (1999), or spontaneously when an electron is stochastically trapped in the bound state. The latter mechanism, employed in the recent experiment Hays et al. (2021), is due to ubiquitous nonequilibrium quasiparticles and likely originated from a background stray photons with energy exceeding the threshold for breaking a Cooper pair Houzet et al. (2019). There are more deterministic approaches to preparing the ASQ in odd parity state Kurilovich et al. (2023); Wesdorp et al. (2023); *Wes2. For example, if the junction is irradiated with the microwave of frequency , the microwave drive can break a Cooper pair placing electrons on the Andreev levels . Moreover, imposing the condition , would guarantee that the microwave drive can only excite the electron from Andreev level into the continuum, which results in the odd parity state Kurilovich et al. (2023).
Once an electron is trapped, the probability of thermally activated parity-switching processes contains an exponentially small factor for tunneling leading to the long lifetime of the trapped electron in the ABS, which is observed to exceed s Zgirski et al. (2011). In general, when there are available subgap states caused by, e.g., spatial variations of the order parameter or impurities Shiba (1968), will be finite provided the trapped electron can tunnel through the gap. We note that the microscopic details of the hybrid structure can greatly modify the residence time for the trapped electron, which can be determined from the “lifetime matrix” lft . As a result, the timescale for the coherent manipulation of the qubit reads
(4) |
where the lower bound is imposed to avoid dynamical mixing with the continuum states when the qubit state is evolving.
IV The qubit manipulation.
Isolated single spins can be coherently manipulated using both electrical and optical techniques Burkard et al. (2023); Kato et al. (2004); *spinc2. In our hybrid structure, however, we accomplish the qubit manipulation by the coherent control of the magnetization dynamics. We study the qubit dynamics when the magnetization precession is both resonant and nonresonant. When resonant, our hybrid structure enhances the Rabi oscillation frequency of the qubit dynamics. When nonresonant, as we discuss, our hybrid structure has the benefit of implementing nonadiabatic single-qubit gates via spintronic techniques. Because of the experimentally achievable high-speed control of the magnetization dynamics Stern et al. (2008); *high-speed2, the nonadiabatic spintronic manipulation of the qubit state renders a shorter qubit evolution time, which is an important advantage in realizing high-fidelity quantum gates Li et al. (2021).
Consider first an FI dynamics at resonance, where the magnetization parametrized as exhibits precessional motion around the axis with ferromagnetic resonance (FMR) frequency , where is the gyromagnetic ratio, and is an effective the magnetic field in the direction. Note that may contain a static external field, demagnetization field, and other crystalline anisotropy fields Suhl (1955); *cone-dyn2, where the effect of the external magnetic field on the qubit, if nonzero, can easily be incorporated into Eq. (2) as a spin-dependent phase shift in the normal part. The cone angle for the magnetization is determined by a transverse rf (microwave) field, , as , where is the dimensionless Gilbert damping constant that parametrizes the inherent spin angular momentum losses of the magnetization dynamics.
The qubit spin, on the other hand, exhibits Rabi oscillations with frequency when , which is the electron-spin-resonance (ESR) frequency. Now, matching the ESR and FMR frequencies and assuming , we get
(5) |
For small damping , . This shows that, for a given microwave stimulus , the magnetization acts as a mediator with the bonus of spatially focusing and intensifying the external field that enhances the qubit Rabi oscillation frequency.
In contrast, when the FI dynamics are nonresonant, , one can control phase to implement arbitrary single-qubit operations. In this regime, a natural spintronic way to maintain a magnetization precession at in the “conical” state is simply by maintaining an out-of-the-plane spin accumulation in an attached normal metal [see Fig. 2(a)]. The normal metal with spin accumulation can exert torque Tserkovnyak and Ochoa (2017) on the magnet at the interface. Note that when the resultant torque is the ordinary Gilbert damping endowed by the normal metal Tserkovnyak et al. (2002). The presence of the spin accumulation can balance the damping torque leading to coherent precession of magnetization at with . In this case, the magnitude of spin accumulation controls the frequency , like in a Josephson relation. As a result, phase changes could be accomplished simply by short pulses of large spin accumulations, which can be within the reach of current spintronic experimental techniques Stern et al. (2008); *high-speed2. To realize the conical state, one could take simply an easy-plane magnet. In a uniaxial crystal, the anisotropy energy contributes to the magnetic free energy with a term , which is minimized when the spontaneous magnetization direction lies in the plane Lifshitz (1995 - 1980). As a result, the out-of-the-plane angle can be controlled by a normal magnetic field.

V Nonadiabatic geometric gate
As an illustrative application, one can adopt a protocol through which the qubit state accumulates only a geometric phase, i.e., the dynamic phase is zero during the whole evolution. To that end, we follow the protocol presented in Ref. Zhao et al. (2017) and show that our ASQ serves as a natural platform to implement a nonadiabatic geometric gate. In the rotating frame, when , the ASQ Hamiltonian can be written as , where . Consider now to undergo a cycle described by the following three fixed orientations of connected by fast quenches:
(6) |
where . As a result, the full qubit evolution operator (or the quantum gate) reads
(7) |
where with . Now, if we impose , , and , the two orthogonal states ,
(8) | ||||
undergo a particular cyclic evolution where the initial and final states are related by a purely geometric phase factor. To see this, one can check that for and , where . This implies that the dynamic phase at each stage of the evolution is zero. As a result, the geometric gate is obtained as
(9) |
In order to elucidate the geometric nature of the angle , we note that the expectation values of spin at times , and are , and , respectively, undergoing a cyclic evolution on the Bloch sphere where the subtended solid angle associated with the enclosed path is given by [see Fig. 2(b)]. Notice that the case of and is equivalent to Hahn spin echo Hahn (1950); *Hahn2 with a -pulse (in the second step). Thus, the protocol given in Eq. (V) can be considered as a generalized spin echo with a -pulse. Since the geometric phase depends on the evolution paths, quantum gates based on the geometric phases are resilient against errors in the evolution details, i.e., control errors Zhao et al. (2017); Jones et al. (2000); *Jones2; Zhang et al. (2015).
VI The qubit readout
To detect or read out the qubit state one can probe a small bias supercurrent, which requires the usage of another weakly coupled superconducting lead to the FI [see the inset in Fig. 2(c)]. Josephson junctions consisting of an FI Senapati et al. (2011) are believed to have interesting properties, such as Josephson state Kawabata et al. (2010); *pi2. Here, to study the ABS in an S-N-FI-S Josephson junction, we model the FI layer as a thin insulating barrier with spin-dependent parameter Tanaka and Kashiwaya (1997). As a result, right- and left-moving electrons (and holes) get a spin-dependent coupling leading to spin-polarized ABS . The discrete spectrum of these bound states can be related to the scattering matrix of the normal region Beenakker (1992) as
(10) |
where is a two by two identity matrix, , the scattering matrix for electrons (holes) with spin , , is given by
(11) |
with , , , and . One can check that these Andreev levels satisfy
(12) |
where with and are spin-dependent reflection and transmission probabilities due to the FI layer, with being the phase shift of the reflected electron from the N-FI interface, and is the phase difference between the superconducting leads. Note that for Eq. (10) produces a similar constraint to that given in Eq. (2). Figures 2(c)-2(e) show the spectrum of ABS for different normal layer thicknesses. As a consequence of the broken spin degeneracy, the supercurrent can depend on the spin state of the occupied Andreev levels Chtchelkatchev and Nazarov (2003). For simplicity, we consider the limit of , where spin-dependent supercurrent can be written as
(13) |
Here, and we introduce the sign convention for spin . Measurement of the spin-dependent supercurrent, e.g., by coupling the supercurrent to a superconducting microwave resonator Janvier et al. (2015); Wallraff et al. (2004); Hays et al. (2020), can provide a qubit state readout.
VII Discussion and conclusion
By fabricating an array of Josephson junctions, our hybrid structure may be generalized to a multi-ASQ system, where the FI magnetizations provide a single-qubit control knob. In a layout with two qubits whose spatial separation is comparable to , the two ABS wave functions with close energies may hybridize and induce an effective exchange interaction between qubits. One way to control this interaction could be to consider a single-channel semiconducting nanowire as the weak link so that a gate voltage might be used to tune the ABS wave functions (via affecting the transmission of the wires) and therefore manipulate the ABS hybridization Kornich et al. (2019); Padurariu and Nazarov (2012). Moreover, in SQUID loops containing these qubits, the interaction between two or multiple qubits can be realized by means of tunable inductive coupling between these SQUID loops Kafri et al. (2017); *talk2. This tunable interaction allows single- or two-qubit operations by preventing unwanted qubit crosstalk. As an immediate application of a tunable exchange interaction, the single-qubit protocol discussed in Eq. (V) can be generalized to a two-qubit geometric gate by periodic manipulation of the exchange interaction Zhang et al. (2020). In a multi-ASQ system, it would be an interesting future study to explore other spin-qubit encodings Burkard et al. (2023); Mishra et al. (2021); *Simon2; *Simon3 that further reduce coupling of the qubit to the environment by storing information in the shared spin space of many qubits.
In this article, we propose a spin qubit in an S-N-FI hybrid structure based on the spin-resolved Andreev-bound states localized in the N region. The qubit state can be coherently controlled by manipulating the FI magnetization dynamics. Our hybrid structure demonstrates the potential for spintronic methods to control the spin qubits. We show that a simple protocol for the magnetization dynamics leads to the realization of a nonadiabatic geometric gate.
This work was supported by the U.S. Department of Energy, Office of Basic Energy Sciences under Award No. DE-SC0012190.
References
- Burkard et al. (2023) Guido Burkard, Thaddeus D. Ladd, Andrew Pan, John M. Nichol, and Jason R. Petta, “Semiconductor spin qubits,” Rev. Mod. Phys. 95, 025003 (2023).
- Vion et al. (2002) D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier, C. Urbina, D. Esteve, and M. H. Devoret, “Manipulating the quantum state of an electrical circuit,” Science 296, 886–889 (2002).
- Devoret et al. (2004) M. H. Devoret, A. Wallraff, and J. M. Martinis, “Superconducting qubits: A short review,” (2004).
- Langer et al. (2005) C. Langer, R. Ozeri, J. D. Jost, J. Chiaverini, B. DeMarco, A. Ben-Kish, R. B. Blakestad, J. Britton, D. B. Hume, W. M. Itano, D. Leibfried, R. Reichle, T. Rosenband, T. Schaetz, P. O. Schmidt, and D. J. Wineland, “Long-lived qubit memory using atomic ions,” Phys. Rev. Lett. 95, 060502 (2005).
- Saffman et al. (2010) M. Saffman, T. G. Walker, and K. Mølmer, “Quantum information with rydberg atoms,” Rev. Mod. Phys. 82, 2313–2363 (2010).
- Zou et al. (2023) Ji Zou, Stefano Bosco, Banabir Pal, Stuart S. P. Parkin, Jelena Klinovaja, and Daniel Loss, “Quantum computing on magnetic racetracks with flying domain wall qubits,” Phys. Rev. Res. 5, 033166 (2023).
- Bocko et al. (1997) M.F. Bocko, A.M. Herr, and M.J. Feldman, “Prospects for quantum coherent computation using superconducting electronics,” IEEE Transactions on Applied Superconductivity 7, 3638–3641 (1997).
- Shafranjuk et al. (2002) S.E Shafranjuk, I.P Nevirkovets, and J.B Ketterson, “A qubit device based on manipulations of andreev bound states in double-barrier josephson junctions,” Solid State Communications 121, 457–460 (2002).
- 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, A. Geresdi, and M. H. Devoret, “Coherent manipulation of an andreev spin qubit,” Science 373, 430–433 (2021).
- Janvier et al. (2015) C. Janvier, L. Tosi, L. Bretheau, Ç. Ö. Girit, M. Stern, P. Bertet, P. Joyez, D. Vion, D. Esteve, M. F. Goffman, H. Pothier, and C. Urbina, “Coherent manipulation of andreev states in superconducting atomic contacts,” Science 349, 1199–1202 (2015).
- Khaetskii et al. (2002) Alexander V. Khaetskii, Daniel Loss, and Leonid Glazman, “Electron spin decoherence in quantum dots due to interaction with nuclei,” Phys. Rev. Lett. 88, 186802 (2002).
- Martins et al. (2016) Frederico Martins, Filip K. Malinowski, Peter D. Nissen, Edwin Barnes, Saeed Fallahi, Geoffrey C. Gardner, Michael J. Manfra, Charles M. Marcus, and Ferdinand Kuemmeth, “Noise suppression using symmetric exchange gates in spin qubits,” Phys. Rev. Lett. 116, 116801 (2016).
- Malinowski et al. (2017) Filip K. Malinowski, Frederico Martins, Peter D. Nissen, Edwin Barnes, Łukasz Cywiński, Mark S. Rudner, Saeed Fallahi, Geoffrey C. Gardner, Michael J. Manfra, Charles M. Marcus, and Ferdinand Kuemmeth, “Notch filtering the nuclear environment of a spin qubit,” Nature Nanotechnology 12, 16 (2017).
- Chtchelkatchev and Nazarov (2003) Nikolai M. Chtchelkatchev and Yu. V. Nazarov, “Andreev quantum dots for spin manipulation,” Phys. Rev. Lett. 90, 226806 (2003).
- Nichele et al. (2020) F. Nichele, E. Portolés, A. Fornieri, A. M. Whiticar, A. C. C. Drachmann, S. Gronin, T. Wang, G. C. Gardner, C. Thomas, A. T. Hatke, M. J. Manfra, and C. M. Marcus, “Relating andreev bound states and supercurrents in hybrid josephson junctions,” Phys. Rev. Lett. 124, 226801 (2020).
- Lantz et al. (2002) J. Lantz, V.S. Shumeiko, E. Bratus, and G. Wendin, “Flux qubit with a quantum point contact,” Physica C: Superconductivity 368, 315–319 (2002).
- Zazunov et al. (2003) A. Zazunov, V. S. Shumeiko, E. N. Bratus’, J. Lantz, and G. Wendin, “Andreev level qubit,” Phys. Rev. Lett. 90, 087003 (2003).
- 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).
- Bratus’ et al. (1997) E. N. Bratus’, V. S. Shumeiko, E. V. Bezuglyi, and G. Wendin, “dc-current transport and ac josephson effect in quantum junctions at low voltage,” Phys. Rev. B 55, 12666–12677 (1997).
- Tosi et al. (2019) L. Tosi, C. Metzger, M. F. Goffman, C. Urbina, H. Pothier, Sunghun Park, A. Levy Yeyati, J. Nygård, and P. Krogstrup, “Spin-orbit splitting of andreev states revealed by microwave spectroscopy,” Phys. Rev. X 9, 011010 (2019).
- Wendin and Shumeiko (2021) Göran Wendin and Vitaly Shumeiko, “Coherent manipulation of a spin qubit,” Science 373, 390–391 (2021).
- Xiang-Bin and Keiji (2001a) Wang Xiang-Bin and Matsumoto Keiji, “Nonadiabatic conditional geometric phase shift with nmr,” Phys. Rev. Lett. 87, 097901 (2001a).
- Zhu and Wang (2002) Shi-Liang Zhu and Z. D. Wang, “Implementation of universal quantum gates based on nonadiabatic geometric phases,” Phys. Rev. Lett. 89, 097902 (2002).
- De Chiara and Palma (2003) Gabriele De Chiara and G. Massimo Palma, “Berry phase for a spin particle in a classical fluctuating field,” Phys. Rev. Lett. 91, 090404 (2003).
- Sjöqvist et al. (2012) Erik Sjöqvist, D M Tong, L Mauritz Andersson, Björn Hessmo, Markus Johansson, and Kuldip Singh, “Non-adiabatic holonomic quantum computation,” New Journal of Physics 14, 103035 (2012).
- Abdumalikov Jr et al. (2013) A. A. Abdumalikov Jr, J. M. Fink, K. Juliusson, M. Pechal, S. Berger, A. Wallraff, and S. Filipp, “Experimental realization of non-abelian non-adiabatic geometric gates,” Nature 496, 482 (2013).
- Zu et al. (2014) C. Zu, W.-B. Wang, L. He, W.-G. Zhang, C.-Y. Dai, F. Wang, and L.-M. Duan, “Experimental realization of universal geometric quantum gates with solid-state spins,” Nature 514, 72 (2014).
- Zhao et al. (2017) P. Z. Zhao, Xiao-Dan Cui, G. F. Xu, Erik Sjöqvist, and D. M. Tong, “Rydberg-atom-based scheme of nonadiabatic geometric quantum computation,” Phys. Rev. A 96, 052316 (2017).
- Zhang et al. (2020) Chengxian Zhang, Tao Chen, Sai Li, Xin Wang, and Zheng-Yuan Xue, “High-fidelity geometric gate for silicon-based spin qubits,” Phys. Rev. A 101, 052302 (2020).
- Eschrig (2018) M. Eschrig, “Theory of andreev bound states in s-f-s junctions and s-f proximity devices,” Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 376, 20150149 (2018).
- (31) When the Fermi velocity between the normal and the superconducting regions is different, the mismatch at the NS interface will result in some normal reflection, even at a clean interface with no barrier present. Although our calculation can easily be generalized to include such a mismatch, here, we consider no mismatch at the interface. For more details, see: G. E. Blonder and M. Tinkham, Phys. Rev. B, 27 112 (1983) .
- Lundin et al. (1999) N.I. Lundin, L.Y. Gorelik, R.I. Shekhter, M. Jonson, and V.S. Shumeiko, “Mesoscopic superconductors under irradiation: microwave spectroscopy of andreev states,” Superlattices and Microstructures 25, 937–947 (1999).
- Wendin et al. (1999) Göran Wendin, Vitaly S. Shumeiko, and Peter Samuelsson, “Controlling josephson transport by manipulation of andreev levels in ballistic mesoscopic junctions,” Superlattices and Microstructures 25, 983–992 (1999).
- Houzet et al. (2019) M. Houzet, K. Serniak, G. Catelani, M. H. Devoret, and L. I. Glazman, “Photon-assisted charge-parity jumps in a superconducting qubit,” Phys. Rev. Lett. 123, 107704 (2019).
- Kurilovich et al. (2023) P. D. Kurilovich, V. D. Kurilovich, A. E. Svetogorov, W. Belzig, M H. Devoret, and L. I. Glazman, “On-demand population of andreev levels by their ionization in the presence of coulomb blockade,” (2023), arXiv:2312.07512.
- Wesdorp et al. (2023) J. J. Wesdorp, L. Grünhaupt, A. Vaartjes, M. Pita-Vidal, A. Bargerbos, L. J. Splitthoff, P. Krogstrup, B. van Heck, and G. de Lange, “Dynamical polarization of the fermion parity in a nanowire josephson junction,” Phys. Rev. Lett. 131, 117001 (2023).
- Ackermann et al. (2023) Nico Ackermann, Alex Zazunov, Sunghun Park, Reinhold Egger, and Alfredo Levy Yeyati, “Dynamical parity selection in superconducting weak links,” Phys. Rev. B 107, 214515 (2023).
- Zgirski et al. (2011) M. Zgirski, L. Bretheau, Q. Le Masne, H. Pothier, D. Esteve, and C. Urbina, “Evidence for long-lived quasiparticles trapped in superconducting point contacts,” Phys. Rev. Lett. 106, 257003 (2011).
- Shiba (1968) Hiroyuki Shiba, “Classical Spins in Superconductors,” Progress of Theoretical Physics 40, 435–451 (1968).
- (40) One way to control the electron lifetime time is to connect the superconductor to a lead L. As a result, one can define the lifetime matrix , where , and is the scattering (full reflection) matrix for the FI-N-S-L hybrid structure. is a Hermitian matrix constructed in the scattering-channel basis of lead L (electron, hole, and spin). When is diagonalized, its eigenvalues corresponds to the reflection time for an incident electron with energy . The bound states energies can be obtained by maximizing with respect to . For details, see F. T. Smith, Phys. Rev. 118, 1, (1960), or C. Texier, Physica E: Low-dimensional Systems and Nanostructures, 82, 16, (2016) .
- Kato et al. (2004) Y. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, “Coherent spin manipulation without magnetic fields in strained semiconductors,” Nature 427, 50 (2004).
- Hanson and Awschalom (2008) R. Hanson and D. D. Awschalom, “Coherent manipulation of single spins in semiconductors,” Nature 453, 1043 (2008).
- Stern et al. (2008) N. P. Stern, D. W. Steuerman, S. Mack, A. C. Gossard, and D. D. Awschalom, “Time-resolved dynamics of the spin hall effect,” Nature Physics 4, 843 (2008).
- Sala et al. (2021) G. Sala, V. Krizakova, E. Grimaldi, C.-H. Lambert, T. Devolder, and P. Gambardella, “Real-time hall-effect detection of current-induced magnetization dynamics in ferrimagnets,” Nature Communications 12, 656 (2021).
- Li et al. (2021) Sai Li, Jing Xue, Tao Chen, and Zheng-Yuan Xue, “High-fidelity geometric quantum gates with short paths on superconducting circuits,” Advanced Quantum Technologies 4, 2000140 (2021).
- Suhl (1955) H. Suhl, “Ferromagnetic resonance in nickel ferrite between one and two kilomegacycles,” Phys. Rev. 97, 555–557 (1955).
- Sovskii (2016) S. V. Von Sovskii, Ferromagnetic Resonance (Elsevier Science, 2016).
- Tserkovnyak and Ochoa (2017) Yaroslav Tserkovnyak and Hector Ochoa, “Generalized boundary conditions for spin transfer,” Phys. Rev. B 96, 100402 (2017).
- Tserkovnyak et al. (2002) Yaroslav Tserkovnyak, Arne Brataas, and Gerrit E. W. Bauer, “Enhanced gilbert damping in thin ferromagnetic films,” Phys. Rev. Lett. 88, 117601 (2002).
- Lifshitz (1995 - 1980) E. M. Lifshitz, Statistical physics. Part 2, Theory of the condensed state, 3rd ed., Course of Theoretical Physics; Volume 9 (Butterworth-Heinemann, Oxford, 1995 - 1980).
- Takahashi and Maekawa (2008) Saburo Takahashi and Sadamichi Maekawa, “Spin current, spin accumulation and spin hall effect,” Science and Technology of Advanced Materials 9, 014105 (2008), pMID: 27877931.
- Hahn (1950) E. L. Hahn, “Spin echoes,” Phys. Rev. 80, 580–594 (1950).
- Carr and Purcell (1954) H. Y. Carr and E. M. Purcell, “Effects of diffusion on free precession in nuclear magnetic resonance experiments,” Phys. Rev. 94, 630–638 (1954).
- Jones et al. (2000) Jonathan A. Jones, Vlatko Vedral, Artur Ekert, and Giuseppe Castagnoli, “Geometric quantum computation using nuclear magnetic resonance,” Nature 403, 869 (2000).
- Xiang-Bin and Keiji (2001b) Wang Xiang-Bin and Matsumoto Keiji, “Nonadiabatic conditional geometric phase shift with nmr,” Phys. Rev. Lett. 87, 097901 (2001b).
- Zhang et al. (2015) J. Zhang, Thi Ha Kyaw, D. M. Tong, Erik Sjöqvist, and Leong-Chuan Kwek, “Fast non-abelian geometric gates via transitionless quantum driving,” Scientific Reports 5, 18414 (2015).
- Senapati et al. (2011) K. Senapati, M. G. Blamire, and Z. H. Barber, “Spin-filter josephson junctions,” Nature Materials 10, 849 (2011).
- Kawabata et al. (2010) Shiro Kawabata, Yasuhiro Asano, Yukio Tanaka, Alexander A. Golubov, and Satoshi Kashiwaya, “Josephson state in a ferromagnetic insulator,” Phys. Rev. Lett. 104, 117002 (2010).
- Kawabata et al. (2006) Shiro Kawabata, Satoshi Kashiwaya, Yasuhiro Asano, Yukio Tanaka, and Alexander A. Golubov, “Macroscopic quantum dynamics of junctions with ferromagnetic insulators,” Phys. Rev. B 74, 180502 (2006).
- Tanaka and Kashiwaya (1997) Yukio Tanaka and Satoshi Kashiwaya, “Theory of josephson effect in superconductor-ferromagnetic-insulator-superconductor junction,” Physica C: Superconductivity 274, 357–363 (1997).
- Beenakker (1992) C. W. J. Beenakker, “Three “universal” mesoscopic josephson effects,” in Transport Phenomena in Mesoscopic Systems, edited by Hidetoshi Fukuyama and Tsuneya Ando (Springer Berlin Heidelberg, Berlin, Heidelberg, 1992) pp. 235–253.
- Wallraff et al. (2004) A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, “Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics,” Nature 431, 162 (2004).
- Hays et al. (2020) M. Hays, V. Fatemi, D. Bouman, S. Diamond, G. de Lange, P. Krogstrup, J. Nygård, A. Geresdi, and M. H. Devoret, “Continuous monitoring of a trapped superconducting spin,” Nature Physics 16, 1103 (2020).
- Kornich et al. (2019) Viktoriia Kornich, Hristo S. Barakov, and Yuli V. Nazarov, “Fine energy splitting of overlapping andreev bound states in multiterminal superconducting nanostructures,” Phys. Rev. Res. 1, 033004 (2019).
- Padurariu and Nazarov (2012) C. Padurariu and Yu. V. Nazarov, “Spin blockade qubit in a superconducting junction,” Europhysics Letters 100, 57006 (2012).
- Kafri et al. (2017) Dvir Kafri, Chris Quintana, Yu Chen, Alireza Shabani, John M. Martinis, and Hartmut Neven, “Tunable inductive coupling of superconducting qubits in the strongly nonlinear regime,” Phys. Rev. A 95, 052333 (2017).
- LaPierre (2021) R. LaPierre, Introduction to Quantum Computing, The Materials Research Society Series (Springer Cham, 2021).
- Mishra et al. (2021) Archana Mishra, Pascal Simon, Timo Hyart, and Mircea Trif, “Yu-shiba-rusinov qubit,” PRX Quantum 2, 040347 (2021).
- Fedele et al. (2021) Federico Fedele, Anasua Chatterjee, Saeed Fallahi, Geoffrey C. Gardner, Michael J. Manfra, and Ferdinand Kuemmeth, “Simultaneous operations in a two-dimensional array of singlet-triplet qubits,” PRX Quantum 2, 040306 (2021).
- Levy (2002) Jeremy Levy, “Universal quantum computation with spin- pairs and heisenberg exchange,” Phys. Rev. Lett. 89, 147902 (2002).