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

Spin-Phonon-Photon Strong Coupling in a Piezoelectric Nanocavity

Hamza Raniwala Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    Stefan Krastanov Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    Lisa Hackett Sandia National Laboratories, Albuqurque, NM, USA    Matt Eichenfield Sandia National Laboratories, Albuqurque, NM, USA    Dirk R. Englund Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    Matthew E. Trusheim Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA U.S. Army Research Laboratory, Sensors and Electron Devices Directorate, Adelphi, Maryland 20783, USA
Abstract

We introduce a hybrid tripartite quantum system for strong coupling between a semiconductor spin, a mechanical phonon, and a microwave photon. Consisting of a piezoelectric resonator with an integrated diamond strain concentrator, this system achieves microwave-acoustic and spin-acoustic coupling rates \simMHz or greater, allowing for simultaneous ultra-high cooperativities (103\sim 10^{3} and 102\sim 10^{2}, respectively). From finite-element modeling and master equation simulations, we estimate photon-to-spin quantum state transfer fidelities exceeding 0.97 based on separately demonstrated device parameters. We anticipate that this device will enable hybrid quantum architectures that leverage the advantages of both superconducting circuits and solid-state spins for information processing, memory, and networking.

Solid-state quantum systems based on superconductors and spins are leading platforms that offer complementary advantages in quantum computing and networking. Superconducting quantum processors enable fast and high-fidelity entangling gates [1, 2], but challenges remain in quantum memory time and long-distance networking. Conversely, atom-like emitters in solid-state have demonstrated long spin coherence time, efficient spin-photon interfaces for long-distance entanglement, and high readout fidelity [3, 4, 5, 6, 7, 8]. Coupling these modalities is therefore an exciting direction in quantum information science.

Previous studies using magnetic coupling between microwave (MW) photons and spins have been limited to multi-spin ensemble interactions [9, 10, 11, 12, 13, 14, 15] due to low spin-magnetic susceptibility and the low magnetic energy density of MW resonators [16, 17, 18]. Alternate experiments and proposals rely on coupling via intermediate acoustic modes [19, 20, 21], which have experimentally demonstrated large coupling to superconducting circuits [22, 23, 24, 25, 26] and are predicted to have large coupling to diamond quantum emitters [27, 28, 29, 30, 31, 32, 33]. However, designing a device that strongly couples one phonon to both one MW photon and to one spin – enabling an efficient MW photon-to-spin interface – remains an outstanding challenge.

Here we address this problem through the co-design of a scandium-doped aluminum nitride (ScAlN) Lamb wave resonator with a heterogeneously-integrated diamond thin film. This structure piezoelectrically couples a MW photon and acoustic phonon while concentrating strain at the location of a diamond quantum emitter. Through finite-element modeling, we predict photon-phonon coupling 10\sim 10 MHz concurrent with phonon-spin coupling 3\sim 3 MHz. These rates yield photon-phonon and phonon-spin cooperativities of order 10410^{4} assuming demonstrated lifetimes of spins, mechanical resonators, and superconducting circuits [34, 35]. We explore state transfer protocols via quantum master equation (QME) simulations and show that this device can achieve photon-to-spin transduction fidelity F>0.97F>0.97 with conservative hardware parameters. We find that performance of these schemes is likely limited by two-level system (TLS) loss in current piezoelectrics. An improvement in piezoelectric TLS loss rates to that of silicon will pave the way towards SC-spin state transduction with F>0.99F>0.99.

We consider a coupled tripartite system consisting of a superconducting circuit (SC), acoustic phonon, and Group-IV electron spin with the Hamiltonian (see [36] for detailed derivation)

H^=ωsc2σ^scz+ωpa^pa^p+ωe2σ^ez+gsc,p(σ^sc+a^p+σ^sca^p)+gp,e(σ^e+a^p+σ^ea^p).\displaystyle\begin{multlined}\frac{\hat{H}}{\hbar}=\frac{\omega_{sc}}{2}\hat{\sigma}^{z}_{sc}+\omega_{p}\hat{a}_{p}^{\dagger}\hat{a}_{p}+\frac{\omega_{e}}{2}\hat{\sigma}^{z}_{e}\\ +g_{sc,p}\left(\hat{\sigma}^{+}_{sc}\hat{a}_{p}+\hat{\sigma}^{-}_{sc}\hat{a}_{p}^{\dagger}\right)+g_{p,e}\left(\hat{\sigma}^{+}_{e}\hat{a}_{p}+\hat{\sigma}^{-}_{e}\hat{a}_{p}^{\dagger}\right).\end{multlined}\frac{\hat{H}}{\hbar}=\frac{\omega_{sc}}{2}\hat{\sigma}^{z}_{sc}+\omega_{p}\hat{a}_{p}^{\dagger}\hat{a}_{p}+\frac{\omega_{e}}{2}\hat{\sigma}^{z}_{e}\\ +g_{sc,p}\left(\hat{\sigma}^{+}_{sc}\hat{a}_{p}+\hat{\sigma}^{-}_{sc}\hat{a}_{p}^{\dagger}\right)+g_{p,e}\left(\hat{\sigma}^{+}_{e}\hat{a}_{p}+\hat{\sigma}^{-}_{e}\hat{a}_{p}^{\dagger}\right). (3)
Refer to caption
Figure 1: Coupled SC-phonon-spin quantum system. (a-c) depict the uncoupled modes of the (a) superconducting qubit with Josephson capacitance CJC_{J}, shunt capacitance CSC_{S}, and external flux bias ϕext\phi_{ext}; (b) acoustic mode capacitively coupled by CIDTC_{IDT}; and (c) diamond quantum emitter. (d) Piezoelectric interaction, where the color indicates the electric field profile under mechanical displacement. (e) Spin-strain coupling resulting from modulating the inter-atomic distance of the quantum emitter via mechanical strain under an external BB field B=Bxx^+Bzz^\textbf{B}=B_{x}\hat{\textbf{x}}+B_{z}\hat{\textbf{z}} with spin-gyromagnetic ratio γ\gamma.

Here, the SC frequency ωsc\omega_{sc} is defined by the transmon Josephson and shunt capacitances, the spin frequency ωe\omega_{e} is given by the Zeeman splitting of the electron spin states, and the acoustic frequency ωp\omega_{p} is defined by the acoustic resonator geometry. The first three terms of this equation describe the energies of the uncoupled modes of the devices (Fig. 1(a-c)) while the fourth and fifth terms describe the interaction dynamics. Generally, SCs feature ωsc46\omega_{sc}\sim 4-6 GHz [37]. Electron spin resonant frequencies can be arbitrarily set by an external magnetic field; to match this frequency range, fields 0.1\sim 0.1 T are required [38]. The coupling coefficient gsc,pg_{sc,p} is physically governed by the piezoelectric effect, whereby a strain field produces an electric response and vice versa (Fig. 1(d)). This interaction is described by the strain-charge equations

Sij\displaystyle S_{ij} =sijklTkl+dkijEk,\displaystyle=s_{ijkl}T_{kl}+d_{kij}E_{k}, (4)
Di\displaystyle D_{i} =dijkTij+ϵikEk,\displaystyle=d_{ijk}T_{ij}+\epsilon_{ik}E_{k}, (5)

where sijkls_{ijkl} and dijkd_{ijk} are the elastic and piezoelectric coefficient tensors of the resonator’s piezoelectric material, SijS_{ij} and TijT_{ij} are the stress and strain fields, and EiE_{i} and DiD_{i} are the electric and displacement fields. At single quantum levels, a MW photon in the SC will generate an electric field in the volume of the piezoelectric resonator described by

esc(r)=(ωsc(CS+CJ+CIDT)Vapp2/2)EIDT(r)eiωsct,\textbf{e}_{sc}(\textbf{r})=\sqrt{\left(\frac{\hbar\omega_{sc}}{\left(C_{S}+C_{J}+C_{IDT}\right)V_{app}^{2}/2}\right)}\textbf{E}_{IDT}(\textbf{r})e^{-i\omega_{sc}t}, (6)

where EIDT(r)\textbf{E}_{IDT}(\textbf{r}) is the electric field profile of the IDT for an arbitrary applied voltage VappV_{app}, and the capacitances are indicated in Fig. 1. Since CSC_{S} is typically much larger than CIDTC_{IDT} and CJC_{J} for transmon qubit configurations, the MW photon energy is largely contained in the shunt capacitor. Similarly, a phonon in the piezoelectric resonator will produce a strain field described by

tp(r)=(ωpV𝑑Vs(r)|Tp(r)|2/2)Tp(r)eiωpt,\textbf{t}_{p}(\textbf{r})=\sqrt{\left(\frac{\hbar\omega_{p}}{\int_{V}dV\;\textbf{s}(\textbf{r})\absolutevalue{\textbf{T}_{p}(\textbf{r})}^{2}/2}\right)}\textbf{T}_{p}(\textbf{r})e^{-i\omega_{p}t}, (7)

where Tp(r)\textbf{T}_{p}(\textbf{r}) is the strain profile of the acoustic mode for an arbitrary mechanical displacement. Following (5), tp(r)\textbf{t}_{p}(\textbf{r}) will produce an electric displacement field given by dtp(r)\textbf{d}\cdot\textbf{t}_{p}(\textbf{r}), where d is the piezoelectric coefficient tensor. Then the coupling gsc,pg_{sc,p} will be determined by the overlap integral between esc(r)\textbf{e}_{sc}(\textbf{r}) and dtp(r)\textbf{d}\cdot\textbf{t}_{p}(\textbf{r}) [39],

gsc,p=12V𝑑V(tp(r)dTesc(r)+esc(r)dtp(r)).g_{sc,p}=\frac{1}{2\hbar}\int_{V}dV\left(\textbf{t}^{*}_{p}(\textbf{r})\cdot\textbf{d}^{T}\cdot\textbf{e}_{sc}(\textbf{r})+\textbf{e}^{*}_{sc}(\textbf{r})\cdot\textbf{d}\cdot\textbf{t}_{p}(\textbf{r})\right). (8)
Refer to caption
Figure 2: Electromechanical transducer design. (a) Lamb wave resonator and relevant design parameters. In this Letter, the resonator geometry is parametrized by (λ,w,td,tAl,tAlN,tSi)=(1370,465,100,100,300,250)(\lambda,w,t_{d},t_{Al},t_{AlN},t_{Si})=(1370,465,100,100,300,250)[nm] (tit_{i} is the thickness of layer ii), with the diamond taper defined by (b,r,θ)=(40 nm,25 nm,50)(b,r,\theta)=(40\text{ nm},25\text{ nm},50^{\circ}). The phononic tethers are of two types: support tethers defined by (Ws,Ls,ws,ls)=(705,565,110150)(W_{s},L_{s},w_{s},l_{s})=(705,565,110150)[nm] and electrode tethers defined by (We,Le,we,le)=(685,565,110,150)(W_{e},L_{e},w_{e},l_{e})=(685,565,110,150)[nm]. (b,c) Phononic band structure of the support (c: electrode) tethers, with a 500 MHz band gap indicated in grey shading and the resonant frequency of the device indicated with the red line. (d) Normalized mechanical displacement of the resonator. (e) Induced piezoelectric displacement field at the central slice of the piezoelectric layer. (f) Spatial profile of gp,eg_{p,e} at the center slice of the diamond layer, assuming a magnetic field of 0.18 T.

The spin-phonon coupling gp,eg_{p,e} results from the spin-strain susceptibility χspin\mathbf{\chi}_{spin} of quantum emitters in a strain field [38, 40, 41]. For a single-phonon strain profile tp\textbf{t}_{p}, the resulting coupling is gspin(r)=χspintp(r)g_{spin}(\textbf{r})=\mathbf{\chi}_{spin}\cdot\textbf{t}_{p}(\textbf{r}). In Group IV emitters in diamond, χspin\chi_{spin} depends heavily on the spin-orbit mixing enabled by an off-axis magnetic field (see [36]) and primarily interacts with transverse strain in the emitter frame [38]. Therefore, for the rest of this analysis, we set this expression to be

gp,e(r)=χeff(txx(r)tyy(r)),g_{p,e}(\textbf{r})=\chi_{eff}(t^{\prime}_{xx}(\textbf{r})-t^{\prime}_{yy}(\textbf{r})), (9)

where t(r)\textbf{t}^{\prime}(\textbf{r}) is the single-phonon strain profile in the coordinate system of the emitter and χeff0.28\chi_{eff}\approx 0.28 PHz/strain [41].

To implement the device in Fig.1, we require a platform with (i) superconductivity, (ii) piezoelectricity, (iii) acoustic cavities, and (iv) strain transfer to diamond emitters. To address (i-ii), we propose a silicon-on-insulator (SOI) platform with a thin-film deposition of scandium-doped aluminum nitride (ScAlN). This material system allows for superconducting qubits and piezoelectrics to co-inhabit one chip [42, 43]. To answer (iii-iv), we co-design a Nb-on-Sc0.32Al0.68N-on-SOI piezoelectric resonator with a heterogeneously integrated diamond thin membrane. We propose Niobium (Nb) as a well-characterized superconductor with high Hc1=0.18H_{c1}=0.18 T and Hc2=2H_{c2}=2 T [44, 45, 46], as required for operation with the spin. SOI platforms have previously been used for piezoelectric resonators [47, 48], and diamond-AlN interfaces have been used to acoustically drive emitters in diamond [49, 50, 51]. ScAlN further boosts the piezoelectric coefficient of AlN, allowing us to achieve a stronger interaction [52, 53].

We present the resonator design in Fig. 2. Our device is based on Lamb wave resonators, which produce standing acoustic waves dependent on electrode periodicity λ\lambda and material thickness [54, 55, 56]. We localize the strain in the diamond thin film using a central “defect cell” (Fig. 2a inset) with a suspended taper. To maintain high quality factors, we tether the Lamb wave resonator via phononic crystal tethers placed at displacement nodes of the box. [57]. We further propose an angled ScAlN sidewall in the transducer (15 from normal) that allows the electrodes to ”climb” on top of the ScAlN film, rather than requiring a continuous piezoelectric layer over the phononic tethers. This both facilitates the design of wide-bandgap phononic tethers and is compatible with current fabrication techniques.

To calculate device performance, we simulate the architecture using the finite element method (FEM) in COMSOL to produce the phononic tether band structures and mode profiles (Fig. 2b-e). The tether band structure exhibits a 500 MHz bandgap around the device’s \approx 4.11 GHz resonant mode. This frequency is desirable as it falls near the central operating range of most superconducting qubits [37]. Additionally, the 4.11 GHz resonant mode is itself isolated from other acoustic modes of the system by \sim56 MHz, which is enough to neglect parastic couplings and treat the transducer in the single-mode approximation (see SI). Fig. 2(d-e) show the mechanical and electrical displacement fields of this mode, from which we derive 𝐞sc(𝐫)\mathbf{e}_{sc}(\mathbf{r}) and 𝐭p(𝐫)\mathbf{t}_{p}(\mathbf{r}), respectively. We calculate a gsc,p7.020.5g_{sc,p}\approx 7.0-20.5 MHz (for a shunt capacitance of 65-190 fF, corresponding to 100 MHz<EC/h<300 MHz100\text{ MHz}<E_{C}/h<300\text{ MHz}[37]) and a maximum gp,e3.2g_{p,e}\approx 3.2 MHz according to Equations (8) and (9). The strain maximum occurs at the edges of the central diamond taper, which maximizes gp,eg_{p,e} (Fig. 2f).

Refer to caption
Figure 3: Analysis of the coupled SC-phonon-spin system under different protocols: (a-c) uncontrolled time evolution, when all modes are on resonance and coupling rates are maximized; (d-f) time evolution detuned from the acoustic resonance, which allows for state transfer through virtual phonon excitation; and (g-h) time evolution under detuning control, which allows for controlled Rabi flops across the modes. Plots (b,e,h) depict the population dynamics of each mode for the above protocols. Plots (c,f,i) show the spin population over time for the variable parameter of the procedure, with operational points for plots (b,e,h) indicated with orange lines. (c) shows population for a given Δg\Delta g, (f) shows population for achievable phonon detuning Δp\Delta_{p}, and (i) shows performance for unused mode detuning Δi\Delta_{i} during each Rabi swap.

In Fig. 3, we explore different protocols for quantum transduction from an initialized SC to a spin. The time evolution of the system when initialized in the ρ0=|100100|\rho_{0}=\ket{100}\bra{100} state (where the indices consecutively refer to the state of the SC, the Fock state of the phonon, and the z-projection of the spin) is calculated using the Lindblad master equation,

ddtρ=i[ρ,H^(t)]+κp(a^pρa^pa^pa^pρρa^pa^p)+κsc(σ^scρσ^sc+σ^sc+σ^scρρσ^sc+σ^sc)+κe(σ^eρσ^e+σ^e+σ^eρρσ^e+σ^e).\begin{multlined}\frac{d}{dt}\rho=-\frac{i}{\hbar}[\rho,\hat{H}(t)]+\kappa_{p}\left(\hat{a}_{p}^{\dagger}\rho\hat{a}_{p}-\hat{a}_{p}^{\dagger}\hat{a}_{p}\rho-\rho\hat{a}_{p}^{\dagger}\hat{a}_{p}\right)\\ +\kappa_{sc}\left(\hat{\sigma}_{sc}^{-}\rho\hat{\sigma}_{sc}^{+}-\hat{\sigma}_{sc}^{+}\hat{\sigma}_{sc}^{-}\rho-\rho\hat{\sigma}_{sc}^{+}\hat{\sigma}_{sc}^{-}\right)\\ +\kappa_{e}\left(\hat{\sigma}_{e}^{-}\rho\hat{\sigma}_{e}^{+}-\hat{\sigma}_{e}^{+}\hat{\sigma}_{e}^{-}\rho-\rho\hat{\sigma}_{e}^{+}\hat{\sigma}_{e}^{-}\right).\end{multlined}\frac{d}{dt}\rho=-\frac{i}{\hbar}[\rho,\hat{H}(t)]+\kappa_{p}\left(\hat{a}_{p}^{\dagger}\rho\hat{a}_{p}-\hat{a}_{p}^{\dagger}\hat{a}_{p}\rho-\rho\hat{a}_{p}^{\dagger}\hat{a}_{p}\right)\\ +\kappa_{sc}\left(\hat{\sigma}_{sc}^{-}\rho\hat{\sigma}_{sc}^{+}-\hat{\sigma}_{sc}^{+}\hat{\sigma}_{sc}^{-}\rho-\rho\hat{\sigma}_{sc}^{+}\hat{\sigma}_{sc}^{-}\right)\\ +\kappa_{e}\left(\hat{\sigma}_{e}^{-}\rho\hat{\sigma}_{e}^{+}-\hat{\sigma}_{e}^{+}\hat{\sigma}_{e}^{-}\rho-\rho\hat{\sigma}_{e}^{+}\hat{\sigma}_{e}^{-}\right). (10)

where the Hamiltonian in a frame rotating at rate ωp\omega_{p} is

H^(t)=Δsc(t)2σ^scz+Δe(t)2σ^ez+gsc,p(σ^sc+a^p+σ^sca^p)+gp,e(σ^e+a^p+σ^ea^p).\displaystyle\begin{multlined}\frac{\hat{H}(t)}{\hbar}=\frac{\Delta_{sc}(t)}{2}\hat{\sigma}^{z}_{sc}+\frac{\Delta_{e}(t)}{2}\hat{\sigma}^{z}_{e}\\ +g_{sc,p}\left(\hat{\sigma}^{+}_{sc}\hat{a}_{p}+\hat{\sigma}^{-}_{sc}\hat{a}_{p}^{\dagger}\right)+g_{p,e}\left(\hat{\sigma}^{+}_{e}\hat{a}_{p}+\hat{\sigma}^{-}_{e}\hat{a}_{p}^{\dagger}\right).\end{multlined}\frac{\hat{H}(t)}{\hbar}=\frac{\Delta_{sc}(t)}{2}\hat{\sigma}^{z}_{sc}+\frac{\Delta_{e}(t)}{2}\hat{\sigma}^{z}_{e}\\ +g_{sc,p}\left(\hat{\sigma}^{+}_{sc}\hat{a}_{p}+\hat{\sigma}^{-}_{sc}\hat{a}_{p}^{\dagger}\right)+g_{p,e}\left(\hat{\sigma}^{+}_{e}\hat{a}_{p}+\hat{\sigma}^{-}_{e}\hat{a}_{p}^{\dagger}\right). (13)

Here, Δsc(t)ωsc(t)ωp\Delta_{sc}(t)\equiv\omega_{sc}(t)-\omega_{p} is the superconducting qubit detuning and Δe(t)ωe(t)ωp\Delta_{e}(t)\equiv\omega_{e}(t)-\omega_{p} is the spin detuning at time tt. The use of time-varying detuning can be easily implemented, e.g. via on-chip flux bias lines [58, 59, 60], unlike time-varying coupling rates explored in previous works [20]. We account for dephasing in each mode with conservative estimates on decoherence rates κsc2π=100\frac{\kappa_{sc}}{2\pi}=100 kHz, κp2π=ωp2πQ40\frac{\kappa_{p}}{2\pi}=\frac{\omega_{p}}{2\pi Q}\approx 40 kHz, and κe2π=1\frac{\kappa_{e}}{2\pi}=1 MHz [35, 34, 61, 62, 63]. As cryogenic operation of ScAlN-on-SOI acoustic resonators–as well as diamond hybrid intergration on said devices–has not been previously explored, we further discuss prospects for QmechQ_{mech} below.

Fig. 3(b,e,h) plot the state transfer fidelity Fiψi|ρ(t)|ψiF_{i}\equiv\bra{\psi_{i}}\rho(t)\ket{\psi_{i}} to the target state |ψi=|1i\ket{\psi_{i}}=\ket{1_{i}} under different conditions. In Fig. 3a where the modes are all resonant (ωsc=ωp=ωe=4.11\omega_{sc}=\omega_{p}=\omega_{e}=4.11 GHz), and gsc,p/2π=10g_{sc,p}/2\pi=10 MHz, FeF_{e} is poor due to the mismatch Δg(gp,e)=gsc,pgp,e\Delta g(g_{p,e})=g_{sc,p}-g_{p,e} (Fig. 3c). Assuming one reduces gsc,pg_{sc,p} or gp,eg_{p,e}, for example by increase the qubit shunt capacitance CSC_{S} or reducing the transverse magnetic field, FeF_{e} may increase at the cost of maximum coupling rates.

In Fig. 3b we detune the phonon mode by Δpωpωsc\Delta_{p}\equiv\omega_{p}-\omega_{sc} where ωsc=ωe\omega_{sc}=\omega_{e} and keep the coupling rates matched at 3.0 MHz. In this case, Fe0.95F_{e}\sim 0.95 via virtual excitation of the phonon mode, if the phonon mode is detuned by 3030 MHz. This protocol generates very low population in the phonon mode, primarily exchanging states between the superconducting qubit and spin. If the phonon mode is lossy, this transduction method is then preferred. However, while this protocol features wider efficiency peaks in time, which may require less stringent pulse control (see Fig. 3e), it does not overcome the issue of coupling imbalance and additionally suffers from decoherence of the superconducting qubit and spin modes over a longer protocol time (Fig. 3f).

Fig. 3g shows the optimal solution, assuming control over Δsc(t)\Delta_{sc}(t) and Δe(t)\Delta_{e}(t), in a double Rabi-flop protocol. During this protocol, it is assumed that gsc,p/2π=10g_{sc,p}/2\pi=10 MHz (which overcomes losses during the Rabi flop while still allowing mode isolation during the next flop) and gp,e/2π=3.0g_{p,e}/2\pi=3.0 MHz. We also assume Δsc(t)=0\Delta_{sc}(t)=0 and 0 MHzΔe(t)10\text{ MHz}\leq\Delta_{e}(t)\leq 1 GHz for t{0,π/(2gsc,p)}t\in\{0,\pi/(2g_{sc,p})\}–the duration of a Rabi flop between the SC and phonon. Then, Δsc(t)=Δe(t=0)\Delta_{sc}(t)=\Delta_{e}(t=0) MHz and Δe(t)=0\Delta_{e}(t)=0 for t{π/(2gsc,p),π/(2gsc,p)+π/(2gp,e)}t\in\{\pi/(2g_{sc,p}),\pi/(2g_{sc,p})+\pi/(2g_{p,e})\}–the duration of a Rabi flop between the phonon and spin. This sequentially transfers states between the modes (Fig. 3h), and for Δe(t=0)>500\Delta_{e}(t=0)>500 MHz, can achieve Fe>0.97F_{e}>0.97 (Fig. 3i; for Δi=1.0\Delta_{i}=1.0 GHz, Fe=0.971F_{e}=0.971). In this protocol, we have neglected the losses that can occur when varying Δsc\Delta_{sc} and Δe\Delta_{e}. In reality, one has to select a pair of Δsc\Delta_{sc} and Δe\Delta_{e} that do not fall on resonance with another acoustic mode of the system to prevent Rabi oscillations between the SC qubit or electron spin and an undesired acoustic mode (see SI for more details).

Each of these scenarios achieves transduction to the spin with high fidelity. The third scenario allows the quantum state to persist in the spin without continued interaction with the acoustic or SC modes. While in this state, the electron spin can access other degrees of freedom (e.g. C13{}^{13}C spins [64, 65]).

Since acoustic losses and therefore the total mechanical quality factor QmechQ_{mech} are difficult to predict from first principles, we evaluate the transduction fidelity FeF_{e} of each protocol in different regimes of QmechQ_{mech} in Fig. 4. Here, protocol 1 is the resonant protocol with gsc,p=gp,e=3g_{sc,p}=g_{p,e}=3 MHz; protocol 2 is the virtual excitation protocol with identical gsc,pg_{sc,p} at a detuning of 3030 MHz; and protocol 3 is the Rabi protocol with gsc,p=10g_{sc,p}=10 MHz, gp,e=3g_{p,e}=3 MHz, and Δe(t=0)=1\Delta_{e}(t=0)=1 GHz. QmechQ_{mech} is the inverse sum of three components,

Qmech=(Qc1+ipi(QTLS,i)1+QA1)1.Q_{mech}=\left(Q_{c}^{-1}+\sum_{i}p_{i}\left(Q_{TLS,i}\right)^{-1}+Q_{A}^{-1}\right)^{-1}. (14)

Here, QcQ_{c} is the mechanical clamping loss, which we can engineer to be non-limiting (see [36]), and QAQ_{A} is the Akhieser loss-related QQ, which at millikelvin temperatures is negligible [66]. These two losses are well-described for analogous systems; in contrast, QTLS,iQ_{TLS,i}–the dielectric loss-related QQ–is harder to predict. These QsQs depends on the number of quasi-particles or TLSs trapped in each of the device’s material interfaces and are weighted by the electric field participation pip_{i} in each interface. Given this uncertainty in QTLSiQ_{TLS_{i}}, we lay out the protocol hierarchy as a function of the overall QmechQ_{mech}:

  • If Qmech2×103Q_{mech}\lesssim 2\times 10^{3}, protocol 2 is superior.

  • If 2×103Qmech5×1052\times 10^{3}\lesssim Q_{mech}\lesssim 5\times 10^{5}, protocol 1 is superior.

  • If 5×105Qmech5\times 10^{5}\lesssim Q_{mech}, protocol 3 is superior.

In existing hardware, the largest challenge to reach the high-fidelity regime (F0.99F\gtrsim 0.99) is reducing dielectric loss in the thin-film piezoelectric, as indicated by published intrinsic quality factors of, e.g., monolithic aluminum nitride or lithium niobate resonators [67, 68]. So, while current hardware may encourage us to utilize the virtual coupling protocol for coupling through a lossy intermediary phononic mode, future iterations of this scheme with improved materials and interfaces can expect to break the 0.99 transduction fidelity barrier using a resonant protocol. At this fidelity, SC-spin transduction would surpass the 1% error correction thresholds of common codes and thus be compatible with scalable quantum information processing schemes [69, 70, 71].

An open question remains in the bonding strength between the diamond thin film and underlying resonator, which, if poor, can incur additional losses. However, for single-phonon occupation, the Van der Waals static frictional force exceeds the strain-generated force on the resonator.

Refer to caption
Figure 4: Sweep of protocol performance as a function of the total quality factor of the mechanical resonator. TLS-limited QQs for Si [72, 73], AlN [67], Nb [74], and alternatives in GaAs [74] and LN [68] are in cyan. This device’s clamping-limited QQs as a function of tether number are listed in blue. Akhiezer losses (grey) are non-dominant at T=0.015T=0.015 K. Finally, our assumed Q105Q\approx 10^{5} is in red. The F>0.995F>0.995 regime (dark grey) requires better SCs and spins to achieve.

Ultimately, we have proposed a resonator architecture capable of simultaneously coupling a microwave photonic mode from a superconducting circuit and an electronic spin from a solid state color center to a single phonon. For our calculated coupling parameters and conservatively assumed QQs across the three modes, we expect SC-phonon cooperativity Csc,p=4gsc,p2κscκp4×103C_{sc,p}=\frac{4g_{sc,p}^{2}}{\kappa_{sc}\kappa_{p}}\sim 4\times 10^{3} and similarly, spin-phonon cooperativity Cp,e=4gp,e2κpκe102C_{p,e}=\frac{4g_{p,e}^{2}}{\kappa_{p}\kappa_{e}}\sim 10^{2}. This doubly strongly-coupled architecture has a number of uses. Firstly, it can provide superconducting circuit qubits access to a long-lived quantum memory in the form of a nuclear spin register surrounding the electron spin. Secondly, this resonator can grant superconducting circuit qubits a spin-photon interface for efficient coupling to fiber optical quantum networks. Finally, by multiplexing each SC with several acoustic resonators and each acoustic resonator with several spins, this architecture can yield a memory bank of quantum memories for computational superconducting circuits (see [36] for a more detailed discussion). We believe introducing this quantum transducer into existing superconducting circuits is a large step towards developing a specialized hybrid quantum computer with fast superconducting qubits for processing and slow, long-lived memory qubits in the solid state for storage and communication.

Acknowledgements.
The authors would like to thank Ian Christen for insightful comments and discussions pertaining to this research. HR acknowledges funding from the NDSEG fellowship. HR, SK, ME, and DRE acknowledge funding from the MITRE Corporation and the NSF Center for Ultracold Atoms. MET acknowledges support from the Army Research Laboratory ENIAC Distinguished Postdoctoral Fellowship. ME and LH acknowledge support from Sandia National Laboratories.

References

  • Jurcevic et al. [2021] P. Jurcevic, A. Javadi-Abhari, L. S. Bishop, I. Lauer, D. F. Bogorin, M. Brink, L. Capelluto, O. Günlük, T. Itoko, N. Kanazawa, et al., Demonstration of quantum volume 64 on a superconducting quantum computing system, Quantum Science and Technology 6, 025020 (2021).
  • Arute et al. [2020] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, S. Boixo, M. Broughton, B. B. Buckley, D. A. Buell, et al., Hartree-fock on a superconducting qubit quantum computer, Science 369, 1084 (2020).
  • Steiner et al. [2010] M. Steiner, P. Neumann, J. Beck, F. Jelezko, and J. Wrachtrup, Universal enhancement of the optical readout fidelity of single electron spins at nitrogen-vacancy centers in diamond, Phys. Rev. B 81, 035205 (2010).
  • Rogers et al. [2014] L. J. Rogers, K. D. Jahnke, M. H. Metsch, A. Sipahigil, J. M. Binder, T. Teraji, H. Sumiya, J. Isoya, M. D. Lukin, P. Hemmer, and F. Jelezko, All-optical initialization, readout, and coherent preparation of single silicon-vacancy spins in diamond, Phys. Rev. Lett. 113, 263602 (2014).
  • Trusheim et al. [2020] M. E. Trusheim, B. Pingault, N. H. Wan, M. Gündoğan, L. De Santis, R. Debroux, D. Gangloff, C. Purser, K. C. Chen, M. Walsh, et al., Transform-limited photons from a coherent tin-vacancy spin in diamond, Physical review letters 124, 023602 (2020).
  • Bhaskar et al. [2020] M. K. Bhaskar, R. Riedinger, B. Machielse, D. S. Levonian, C. T. Nguyen, E. N. Knall, H. Park, D. Englund, M. Lončar, D. D. Sukachev, et al., Experimental demonstration of memory-enhanced quantum communication, Nature 580, 60 (2020).
  • Pla et al. [2013] J. J. Pla, K. Y. Tan, J. P. Dehollain, W. H. Lim, J. J. Morton, F. A. Zwanenburg, D. N. Jamieson, A. S. Dzurak, and A. Morello, High-fidelity readout and control of a nuclear spin qubit in silicon, Nature 496, 334 (2013).
  • Nagy et al. [2019] R. Nagy, M. Niethammer, M. Widmann, Y.-C. Chen, P. Udvarhelyi, C. Bonato, J. U. Hassan, R. Karhu, I. G. Ivanov, N. T. Son, et al., High-fidelity spin and optical control of single silicon-vacancy centres in silicon carbide, Nature communications 10, 1 (2019).
  • Ranjan et al. [2013] V. Ranjan, G. De Lange, R. Schutjens, T. Debelhoir, J. Groen, D. Szombati, D. Thoen, T. Klapwijk, R. Hanson, and L. DiCarlo, Probing dynamics of an electron-spin ensemble via a superconducting resonator, Physical review letters 110, 067004 (2013).
  • Xiang et al. [2013] Z.-L. Xiang, X.-Y. Lü, T.-F. Li, J. Q. You, and F. Nori, Hybrid quantum circuit consisting of a superconducting flux qubit coupled to a spin ensemble and a transmission-line resonator, Phys. Rev. B 87, 144516 (2013).
  • Kubo et al. [2011] Y. Kubo, C. Grezes, A. Dewes, T. Umeda, J. Isoya, H. Sumiya, N. Morishita, H. Abe, S. Onoda, T. Ohshima, et al., Hybrid quantum circuit with a superconducting qubit coupled to a spin ensemble, Physical review letters 107, 220501 (2011).
  • Zhu et al. [2011] X. Zhu, S. Saito, A. Kemp, K. Kakuyanagi, S.-i. Karimoto, H. Nakano, W. J. Munro, Y. Tokura, M. S. Everitt, K. Nemoto, et al., Coherent coupling of a superconducting flux qubit to an electron spin ensemble in diamond, Nature 478, 221 (2011).
  • Sigillito et al. [2014] A. J. Sigillito, H. Malissa, A. M. Tyryshkin, H. Riemann, N. V. Abrosimov, P. Becker, H.-J. Pohl, M. L. Thewalt, K. M. Itoh, J. J. Morton, et al., Fast, low-power manipulation of spin ensembles in superconducting microresonators, Applied Physics Letters 104, 222407 (2014).
  • Grezes et al. [2016] C. Grezes, Y. Kubo, B. Julsgaard, T. Umeda, J. Isoya, H. Sumiya, H. Abe, S. Onoda, T. Ohshima, K. Nakamura, et al., Towards a spin-ensemble quantum memory for superconducting qubits, Comptes Rendus Physique 17, 693 (2016).
  • Dold et al. [2019] G. Dold, C. W. Zollitsch, J. O’sullivan, S. Welinski, A. Ferrier, P. Goldner, S. de Graaf, T. Lindström, and J. J. Morton, High-cooperativity coupling of a rare-earth spin ensemble to a superconducting resonator using yttrium orthosilicate as a substrate, Physical Review Applied 11, 054082 (2019).
  • Carter et al. [2015] S. G. Carter, O. O. Soykal, P. Dev, S. E. Economou, and E. R. Glaser, Spin coherence and echo modulation of the silicon vacancy in 4hSiC4h-\mathrm{SiC} at room temperature, Phys. Rev. B 92, 161202 (2015).
  • Rabl et al. [2009] P. Rabl, P. Cappellaro, M. G. Dutt, L. Jiang, J. Maze, and M. D. Lukin, Strong magnetic coupling between an electronic spin qubit and a mechanical resonator, Physical Review B 79, 041302 (2009).
  • Angerer et al. [2018] A. Angerer, K. Streltsov, T. Astner, S. Putz, H. Sumiya, S. Onoda, J. Isoya, W. J. Munro, K. Nemoto, J. Schmiedmayer, et al., Superradiant emission from colour centres in diamond, Nature Physics 14, 1168 (2018).
  • Schuetz et al. [2015] M. J. A. Schuetz, E. M. Kessler, G. Giedke, L. M. K. Vandersypen, M. D. Lukin, and J. I. Cirac, Universal quantum transducers based on surface acoustic waves, Phys. Rev. X 5, 031031 (2015).
  • Neuman et al. [2020] T. Neuman, M. Eichenfield, M. Trusheim, L. Hackett, P. Narang, and D. Englund, A phononic bus for coherent interfaces between a superconducting quantum processor, spin memory, and photonic quantum networks, arXiv preprint arXiv:2003.08383  (2020).
  • Maity et al. [2020] S. Maity, L. Shao, S. Bogdanović, S. Meesala, Y.-I. Sohn, N. Sinclair, B. Pingault, M. Chalupnik, C. Chia, L. Zheng, et al., Coherent acoustic control of a single silicon vacancy spin in diamond, Nature communications 11, 1 (2020).
  • O’Connell et al. [2010] A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, M. Weides, et al., Quantum ground state and single-phonon control of a mechanical resonator, Nature 464, 697 (2010).
  • Arrangoiz-Arriola and Safavi-Naeini [2016] P. Arrangoiz-Arriola and A. H. Safavi-Naeini, Engineering interactions between superconducting qubits and phononic nanostructures, Physical Review A 94, 063864 (2016).
  • Arrangoiz-Arriola et al. [2018] P. Arrangoiz-Arriola, E. A. Wollack, M. Pechal, J. D. Witmer, J. T. Hill, and A. H. Safavi-Naeini, Coupling a superconducting quantum circuit to a phononic crystal defect cavity, Physical Review X 8, 031007 (2018).
  • Arrangoiz-Arriola et al. [2019] P. Arrangoiz-Arriola, A. Wollack, M. Pechal, Z. Wang, W. Jiang, T. McKenna, and A. Safavi-Naeini, Strong coupling of a transmon qubit and a phononic crystal cavity array, in APS March Meeting Abstracts, Vol. 2019 (2019) pp. S29–011.
  • Peterson et al. [2019] G. Peterson, S. Kotler, F. Lecocq, K. Cicak, X. Jin, R. Simmonds, J. Aumentado, and J. Teufel, Ultrastrong parametric coupling between a superconducting cavity and a mechanical resonator, Physical review letters 123, 247701 (2019).
  • Kuzyk and Wang [2018] M. C. Kuzyk and H. Wang, Scaling phononic quantum networks of solid-state spins with closed mechanical subsystems, Physical Review X 8, 041027 (2018).
  • Li et al. [2019] X. Li, M. C. Kuzyk, and H. Wang, Honeycomblike phononic networks of spins with closed mechanical subsystems, Physical Review Applied 11, 064037 (2019).
  • Wang and Lekavicius [2020] H. Wang and I. Lekavicius, Coupling spins to nanomechanical resonators: Toward quantum spin-mechanics, Applied Physics Letters 117, 230501 (2020).
  • Rabl et al. [2010] P. Rabl, S. J. Kolkowitz, F. Koppens, J. Harris, P. Zoller, and M. D. Lukin, A quantum spin transducer based on nanoelectromechanical resonator arrays, Nature Physics 6, 602 (2010).
  • Lekavicius et al. [2019] I. Lekavicius, T. Oo, and H. Wang, Diamond lamb wave spin-mechanical resonators with optically coherent nitrogen vacancy centers, Journal of Applied Physics 126, 214301 (2019).
  • Joe et al. [2021] G. Joe, C. Chia, M. Chalupnik, B. Pingault, S. Meesala, E. Cornell, D. Assumpcao, B. Machielse, and M. Lončar, Diamond phononic crystals with silicon-vacancy centers at cryogenic temperatures, in CLEO: QELS_Fundamental Science (Optical Society of America, 2021) pp. FTh4M–1.
  • Raniwala et al. [2022] H. Raniwala, S. Krastanov, M. Eichenfield, and D. Englund, A spin-optomechanical quantum interface enabled by an ultrasmall mechanical and optical mode volume cavity, arXiv preprint arXiv:2202.06999  (2022).
  • Kjaergaard et al. [2020] M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Superconducting qubits: Current state of play, Annual Review of Condensed Matter Physics 11, 369–395 (2020).
  • Devoret and Schoelkopf [2013] M. H. Devoret and R. J. Schoelkopf, Superconducting circuits for quantum information: an outlook, Science 339, 1169 (2013).
  • [36] See supplemental material including a theoretical analysis of the coupling rates of the tripartite system, a discussion of fem and qme simulations, and a discussion of a spin memory register enabled by this sc-spin transducer architecture.
  • Krantz et al. [2019] P. Krantz, M. Kjaergaard, F. Yan, T. P. Orlando, S. Gustavsson, and W. D. Oliver, A quantum engineer’s guide to superconducting qubits, Applied Physics Reviews 6, 021318 (2019).
  • Hepp et al. [2014] C. Hepp, T. Müller, V. Waselowski, J. N. Becker, B. Pingault, H. Sternschulte, D. Steinmüller-Nethl, A. Gali, J. R. Maze, M. Atatüre, et al., Electronic structure of the silicon vacancy color center in diamond, Physical Review Letters 112, 036405 (2014).
  • Zou et al. [2016] C.-L. Zou, X. Han, L. Jiang, and H. X. Tang, Cavity piezomechanical strong coupling and frequency conversion on an aluminum nitride chip, Phys. Rev. A 94, 013812 (2016).
  • Meesala et al. [2016] S. Meesala, Y.-I. Sohn, H. A. Atikian, S. Kim, M. J. Burek, J. T. Choy, and M. Lončar, Enhanced strain coupling of nitrogen-vacancy spins to nanoscale diamond cantilevers, Physical Review Applied 5, 034010 (2016).
  • Meesala et al. [2018] S. Meesala, Y.-I. Sohn, B. Pingault, L. Shao, H. A. Atikian, J. Holzgrafe, M. Gündoğan, C. Stavrakas, A. Sipahigil, C. Chia, et al., Strain engineering of the silicon-vacancy center in diamond, Physical Review B 97, 205444 (2018).
  • Keller et al. [2017a] A. J. Keller, P. B. Dieterle, M. Fang, B. Berger, J. M. Fink, and O. Painter, Al transmon qubits on silicon-on-insulator for quantum device integration, Applied Physics Letters 111, 042603 (2017a).
  • Keller et al. [2017b] A. J. Keller, P. B. Dieterle, M. Fang, B. Berger, J. M. Fink, and O. Painter, Superconducting qubits on silicon substrates for quantum device integration, arXiv preprint arXiv:1703.10195  (2017b).
  • Saito et al. [2001] K. Saito et al., Critical field limitation of the niobium superconducting rf cavity, in Proceedings of the 10th International Conference on RF Superconductivity, Tsukuba, Japan (2001).
  • Kerchner et al. [1981] H. R. Kerchner, D. K. Christen, and S. T. Sekula, Critical fields Hc{H}_{c} and Hc2{H}_{c2} of superconducting niobium, Phys. Rev. B 24, 1200 (1981).
  • Finnemore et al. [1966] D. K. Finnemore, T. F. Stromberg, and C. A. Swenson, Superconducting properties of high-purity niobium, Phys. Rev. 149, 231 (1966).
  • Löbl et al. [2001] H. Löbl, M. Klee, R. Milsom, R. Dekker, C. Metzmacher, W. Brand, and P. Lok, Materials for bulk acoustic wave (baw) resonators and filters, Journal of the European Ceramic Society 21, 2633 (2001).
  • Loebl et al. [2003] H. Loebl, M. Klee, C. Metzmacher, W. Brand, R. Milsom, and P. Lok, Piezoelectric thin aln films for bulk acoustic wave (baw) resonators, Materials Chemistry and Physics 79, 143 (2003).
  • Golter and Wang [2014] D. A. Golter and H. Wang, Optically driven rabi oscillations and adiabatic passage of single electron spins in diamond, Physical review letters 112, 116403 (2014).
  • Golter et al. [2016a] D. A. Golter, T. Oo, M. Amezcua, I. Lekavicius, K. A. Stewart, and H. Wang, Coupling a surface acoustic wave to an electron spin in diamond via a dark state, Physical Review X 6, 041060 (2016a).
  • Golter et al. [2016b] D. A. Golter, T. Oo, M. Amezcua, K. A. Stewart, and H. Wang, Optomechanical quantum control of a nitrogen-vacancy center in diamond, Physical review letters 116, 143602 (2016b).
  • Akiyama et al. [2009] M. Akiyama, K. Kano, and A. Teshigahara, Influence of growth temperature and scandium concentration on piezoelectric response of scandium aluminum nitride alloy thin films, Applied Physics Letters 95, 162107 (2009).
  • Kurz et al. [2019] N. Kurz, A. Ding, D. F. Urban, Y. Lu, L. Kirste, N. M. Feil, A. Žukauskaitė, and O. Ambacher, Experimental determination of the electro-acoustic properties of thin film alscn using surface acoustic wave resonators, Journal of Applied Physics 126, 075106 (2019).
  • Bjurström et al. [2005] J. Bjurström, I. Katardjiev, and V. Yantchev, Lateral-field-excited thin-film lamb wave resonator, Applied Physics Letters 86, 154103 (2005).
  • Lin et al. [2010] C.-M. Lin, T.-T. Yen, Y.-J. Lai, V. V. Felmetsger, M. A. Hopcroft, J. H. Kuypers, and A. P. Pisano, Temperature-compensated aluminum nitride lamb wave resonators, IEEE transactions on ultrasonics, ferroelectrics, and frequency control 57, 524 (2010).
  • Konno et al. [2013] A. Konno, M. Sumisaka, A. Teshigahara, K. Kano, K.-y. Hashimo, H. Hirano, M. Esashi, M. Kadota, and S. Tanaka, Scaln lamb wave resonator in ghz range released by xef 2 etching, in 2013 IEEE International Ultrasonics Symposium (IUS) (IEEE, 2013) pp. 1378–1381.
  • Mirhosseini et al. [2020] M. Mirhosseini, A. Sipahigil, M. Kalaee, and O. Painter, Superconducting qubit to optical photon transduction, Nature 588, 599 (2020).
  • Sillanpää et al. [2007] M. A. Sillanpää, J. I. Park, and R. W. Simmonds, Coherent quantum state storage and transfer between two phase qubits via a resonant cavity, Nature 449, 438 (2007).
  • Strand et al. [2013] J. D. Strand, M. Ware, F. Beaudoin, T. A. Ohki, B. R. Johnson, A. Blais, and B. L. T. Plourde, First-order sideband transitions with flux-driven asymmetric transmon qubits, Phys. Rev. B 87, 220505 (2013).
  • McKay et al. [2016] D. C. McKay, S. Filipp, A. Mezzacapo, E. Magesan, J. M. Chow, and J. M. Gambetta, Universal gate for fixed-frequency qubits via a tunable bus, Phys. Rev. Applied 6, 064007 (2016).
  • Premkumar et al. [2021] A. Premkumar, C. Weiland, S. Hwang, B. Jäck, A. P. Place, I. Waluyo, A. Hunt, V. Bisogni, J. Pelliciari, A. Barbour, et al., Microscopic relaxation channels in materials for superconducting qubits, Communications Materials 2, 1 (2021).
  • Sukachev et al. [2017] D. D. Sukachev, A. Sipahigil, C. T. Nguyen, M. K. Bhaskar, R. E. Evans, F. Jelezko, and M. D. Lukin, Silicon-vacancy spin qubit in diamond: A quantum memory exceeding 10 ms with single-shot state readout, Phys. Rev. Lett. 119, 223602 (2017).
  • Pingault et al. [2017] B. Pingault, D.-D. Jarausch, C. Hepp, L. Klintberg, J. N. Becker, M. Markham, C. Becher, and M. Atatüre, Coherent control of the silicon-vacancy spin in diamond, Nature communications 8, 1 (2017).
  • Metsch et al. [2019] M. H. Metsch, K. Senkalla, B. Tratzmiller, J. Scheuer, M. Kern, J. Achard, A. Tallaire, M. B. Plenio, P. Siyushev, and F. Jelezko, Initialization and readout of nuclear spins via a negatively charged silicon-vacancy center in diamond, Phys. Rev. Lett. 122, 190503 (2019).
  • Maity et al. [2021] S. Maity, B. Pingault, G. Joe, M. Chalupnik, D. Assumpção, E. Cornell, L. Shao, and M. Lončar, Coherent coupling of mechanics to a single nuclear spin, arXiv preprint arXiv:2107.10961  (2021).
  • Chan [2012] J. Chan, Laser cooling of an optomechanical crystal resonator to its quantum ground state of motion (California Institute of Technology, 2012).
  • Fan et al. [2013] L. Fan, X. Sun, C. Xiong, C. Schuck, and H. X. Tang, Aluminum nitride piezo-acousto-photonic crystal nanocavity with high quality factors, Applied Physics Letters 102, 153507 (2013).
  • Wollack et al. [2021] E. A. Wollack, A. Y. Cleland, P. Arrangoiz-Arriola, T. P. McKenna, R. G. Gruenke, R. N. Patel, W. Jiang, C. J. Sarabalis, and A. H. Safavi-Naeini, Loss channels affecting lithium niobate phononic crystal resonators at cryogenic temperature, Applied Physics Letters 118, 123501 (2021).
  • Kitaev [2003] A. Kitaev, Fault-tolerant quantum computation by anyons, Annals of Physics 303, 2 (2003).
  • Raussendorf et al. [2007] R. Raussendorf, J. Harrington, and K. Goyal, Topological fault-tolerance in cluster state quantum computation, New Journal of Physics 9, 199 (2007).
  • Wang et al. [2009] D. S. Wang, A. G. Fowler, A. M. Stephens, and L. C. L. Hollenberg, Threshold error rates for the toric and surface codes (2009), arXiv:0905.0531 [quant-ph] .
  • Melville et al. [2020] A. Melville, G. Calusine, W. Woods, K. Serniak, E. Golden, B. M. Niedzielski, D. K. Kim, A. Sevi, J. L. Yoder, E. A. Dauler, et al., Comparison of dielectric loss in titanium nitride and aluminum superconducting resonators, Applied Physics Letters 117, 124004 (2020).
  • Woods et al. [2019] W. Woods, G. Calusine, A. Melville, A. Sevi, E. Golden, D. Kim, D. Rosenberg, J. Yoder, and W. Oliver, Determining interface dielectric losses in superconducting coplanar-waveguide resonators, Phys. Rev. Applied 12, 014012 (2019).
  • McRae et al. [2020] C. R. H. McRae, H. Wang, J. Gao, M. R. Vissers, T. Brecht, A. Dunsworth, D. P. Pappas, and J. Mutus, Materials loss measurements using superconducting microwave resonators, Review of Scientific Instruments 91, 091101 (2020).
  • Brown et al. [2008] C. Brown, A. S. Morris, A. I. Kingon, and J. Krim, Cryogenic performance of rf mems switch contacts, Journal of Microelectromechanical Systems 17, 1460 (2008).

Appendix A Theoretical Analysis of Electromechanical Coupling

This section will review the basic theory surrounding a superconducting transmon coupled to a diamond defect spin via an intermediary mechanical mode.

We begin by considering a transmon architecture, which consists of a SQUID loop with combined Josephson energy EJE_{J} and capacitance CJC_{J} in parallel with a shunt capacitor CSC_{S}. For the sake of constructing only the coupled system, we omit the transmon readout resonator, which typically consists of a quarter wave resonator coupled in parallel to the transmon. The transmon’s Josephson and charging energies are EJ(ϕ)=ICΦ0πcos(ϕext)=EJcos(ϕext)E_{J}(\phi)=\frac{I_{C}\Phi_{0}}{\pi}\cos(\phi_{ext})=E_{J}\cos(\phi_{ext}) and EC=e22(CS+CJ)E_{C}=\frac{e^{2}}{2(C_{S}+C_{J})} (ICI_{C} is the Josephson junction critical current). Note here that the total charging energy for n^\hat{n} Cooper pairs will be 4ECn^4E_{C}\hat{n}, where ϕ^\hat{\phi} is the conjugate variable of n^\hat{n}. Then the transmon Hamiltonian is given by

H^transmon\displaystyle\hat{H}_{transmon} =4ECn^+EJ(ϕ^)\displaystyle=4E_{C}\hat{n}+E_{J}(\hat{\phi}) (15)
=4ECn^+EJ(ϕ^+12ϕ^2+16ϕ^3+112ϕ^4+)\displaystyle=4E_{C}\hat{n}+E_{J}\left(\hat{\phi}+\frac{1}{2}\hat{\phi}^{2}+\frac{1}{6}\hat{\phi}^{3}+\frac{1}{12}\hat{\phi}^{4}+...\right) (16)
(8EJECEC)a^a^EC(a^a^a^a^).\displaystyle\approx\left(\sqrt{8E_{J}E_{C}}-E_{C}\right)\hat{a}^{\dagger}\hat{a}-E_{C}(\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a}). (17)

In the last step, we have rewritten in terms of the ladder operators. If we approximate the transmon as a two-level system, then we can simply write H^transmon\hat{H}_{transmon} as

H^transmon/=ωsc2σ^scz.\hat{H}_{transmon}/\hbar=\frac{\omega_{sc}}{2}\hat{\sigma}^{z}_{sc}. (18)

Next, we make note of the Hamiltonian of the electromechanical resonator. Sans coupling, the resonator modes can each be approximated as harmonic oscillators with energy ωp,k\hbar\omega_{p,k}, where ωp,k\omega_{p,k} is the resonant frequency of the kkth resonator mode, plus some vacuum energy terms. Ignoring these terms, the Hamiltonian HresH_{res} is

H^res/=kωp,kb^kb^k.\hat{H}_{res}/\hbar=\sum_{k}\omega_{p,k}\hat{b}_{k}^{\dagger}\hat{b}_{k}. (19)

Finally, we consider the Hamiltonian of the Group IV electron spin. The full Hamiltonian of Group IV color centers has been discussed at length in [38], but for the purposes of this paper we consider the system under an off-axis (transverse and longitudinal) magnetic field (discussed in [20]). In these conditions, the Group IV Hamiltonian can be written as a sum of the spin-orbit Hamiltonian and a Zeeman perturbation (in the {|ex,|ey,|ex,|ey}\{\ket{e_{x}\uparrow},\ket{e_{y}\uparrow},\ket{e_{x}\downarrow},\ket{e_{y}\downarrow}\} basis),

H^spin\displaystyle\hat{H}_{\textrm{spin}} =H^SO+H^Z\displaystyle=\hat{H}^{\text{SO}}+\hat{H}^{\text{Z}} (20)
=[00iλg0000iλgiλg0000iλg00]+[γsBzγsBxiqγLBz0γsBxγsBz0iqγLBziqγLBz0γsBzγsBx0iqγLBzγsBxγsBz]\displaystyle=\begin{bmatrix}0&0&-i\lambda_{g}&0\\ 0&0&0&i\lambda_{g}\\ i\lambda_{g}&0&0&0\\ 0&-i\lambda_{g}&0&0\end{bmatrix}+\begin{bmatrix}\gamma_{s}B_{z}&\gamma_{s}B_{x}&iq\gamma_{L}B_{z}&0\\ \gamma_{s}B_{x}&-\gamma_{s}B_{z}&0&-iq\gamma_{L}B_{z}\\ -iq\gamma_{L}B_{z}&0&\gamma_{s}B_{z}&\gamma_{s}B_{x}\\ 0&iq\gamma_{L}B_{z}&\gamma_{s}B_{x}&\gamma_{s}B_{z}\end{bmatrix} (21)
=[γsBzγsBxiλ0γsBxγsBz0iλiλ0γsBzγsBx0iλγsBxγsBz].\displaystyle=\begin{bmatrix}\gamma_{s}B_{z}&\gamma_{s}B_{x}&-i\lambda&0\\ \gamma_{s}B_{x}&-\gamma_{s}B_{z}&0&i\lambda\\ i\lambda&0&\gamma_{s}B_{z}&\gamma_{s}B_{x}\\ 0&-i\lambda&\gamma_{s}B_{x}&\gamma_{s}B_{z}\end{bmatrix}. (22)

Here, we use λλgqγLBz\lambda\equiv\lambda_{g}-q\gamma_{L}B_{z} [38]. Solving the eigensystem of this Hamiltonian gives us the eigenvectors

|ψ1=(12γs2Bx2+(λ)(λ+γs2Bx2+(λ)2))[(i(λ+γs2Bx2+(λ)2))|ex+i|ex(λ+γs2Bx2+(λ)2)|ey+|ey],\ket{\psi_{1}}=\left(\frac{1}{2\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})\left(\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}\right)}}\right)\Bigg{[}\left(-i\left(\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}\right)\right)\ket{e_{x}\uparrow}+i\ket{e_{x}\downarrow}\\ -\left(\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}\right)\ket{e_{y}\uparrow}+\ket{e_{y}\downarrow}\Bigg{]}, (24)
|ψ2=(12γs2Bx2+(λ)(λ+γs2Bx2+(λ)2))[i(λγs2Bx2+(λ)2λ+γs2Bx2+(λ)2)|ex+i|ex(λγs2Bx2+(λ)2λ+γs2Bx2+(λ)2)|ey+|ey],\ket{\psi_{2}}=\left(\frac{1}{2\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})\left(\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}\right)}}\right)\Bigg{[}-i\left(\frac{\lambda_{-}-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}}{\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}}\right)\ket{e_{x}\uparrow}+i\ket{e_{x}\downarrow}\\ -\left(\frac{\lambda_{-}-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}}{\lambda_{-}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}}\right)\ket{e_{y}\uparrow}+\ket{e_{y}\downarrow}\Bigg{]}, (25)
|ψ3=(12γs2Bx2+(λ+)(λ++γs2Bx2+(λ+)2))[i(λ+γs2Bx2+(λ+)2λ++γs2Bx2+(λ+)2)|ex+i|ex(λ+γs2Bx2+(λ+)2λ++γs2Bx2+(λ+)2)|ey+|ey],\ket{\psi_{3}}=\left(\frac{1}{2\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})\left(\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}\right)}}\right)\Bigg{[}-i\left(\frac{\lambda_{+}-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}}{\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}}\right)\ket{e_{x}\uparrow}+i\ket{e_{x}\downarrow}\\ -\left(\frac{\lambda_{+}-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}}{\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}}\right)\ket{e_{y}\uparrow}+\ket{e_{y}\downarrow}\Bigg{]}, (26)
|ψ4=(12γs2Bx2+(λ+)(λ++γs2Bx2+(λ+)2))[(i(λ++γs2Bx2+(λ+)2))|exi|ex+(λ++γs2Bx2+(λ+)2)|ey+|ey].\ket{\psi_{4}}=\left(\frac{1}{2\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})\left(\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}\right)}}\right)\Bigg{[}\left(-i\left(\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}\right)\right)\ket{e_{x}\uparrow}-i\ket{e_{x}\downarrow}\\ +\left(\lambda_{+}+\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}\right)\ket{e_{y}\uparrow}+\ket{e_{y}\downarrow}\Bigg{]}. (27)

These eigenvectors are associated with the eigenvalues

ν1\displaystyle\nu_{1} =γs2Bx2+(λ)2,\displaystyle=-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}, (28)
ν2\displaystyle\nu_{2} =γs2Bx2+(λ)2,\displaystyle=\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{-})^{2}}, (29)
ν3\displaystyle\nu_{3} =γs2Bx2+(λ+)2,\displaystyle=-\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}, (30)
ν4\displaystyle\nu_{4} =γs2Bx2+(λ+)2.\displaystyle=\sqrt{\gamma_{s}^{2}B_{x}^{2}+(\lambda_{+})^{2}}. (31)

Here, we use λ=λγsBz\lambda_{-}=\lambda-\gamma_{s}B_{z} and λ+=λ+γsBz\lambda_{+}=\lambda+\gamma_{s}B_{z}. (Note that, in the limit where Bx0B_{x}\rightarrow 0, these eigenvectors and eigenvalues simplify as {|ψ1,|ψ2,|ψ3,|ψ4}{|e+,|e+,|e,|e}\left\{\ket{\psi_{1}},\ket{\psi_{2}},\ket{\psi_{3}},\ket{\psi_{4}}\right\}\rightarrow\left\{\ket{e_{+}\uparrow},\ket{e_{+}\downarrow},\ket{e_{-}\downarrow},\ket{e_{-}\uparrow}\right\} from [38]).

Finally, the coupling rate gp,eg_{p,e} between the lowest lying states |ψ1\ket{\psi_{1}} and |ψ3\ket{\psi_{3}} can be calculated as

gp,e2π=|ψ3|M1HstrainM|ψ1|,\frac{g_{p,e}}{2\pi}=\absolutevalue{\bra{\psi_{3}}M^{-1}H_{strain}M\ket{\psi_{1}}}, (32)

where

Hstrain=[α0β00α0ββ0α00β0α]H_{strain}=\begin{bmatrix}\alpha&0&\beta&0\\ 0&\alpha&0&\beta\\ \beta&0&-\alpha&0\\ 0&\beta&0&-\alpha\end{bmatrix} (33)

and MM is the matrix that transforms the eigenvectors ψi\psi_{i} to the strain basis, such that

MH^spin=M[ν10000ν20000ν30000ν4].M\hat{H}_{\textrm{spin}}=M\begin{bmatrix}\nu_{1}&0&0&0\\ 0&\nu_{2}&0&0\\ 0&0&\nu_{3}&0\\ 0&0&0&\nu_{4}\end{bmatrix}. (34)

In SiV- centers in diamond, β\beta is more than ten times smaller than α\alpha [41], so we can simplify HstrainH_{strain} to the case where β0\beta\rightarrow 0 and αχeff(ϵxxϵyy)\alpha\rightarrow\chi_{eff}(\epsilon_{xx}-\epsilon_{yy}) as discussed in the main text (Equation (9)). Then for a known gorbg_{orb} and a maximum magnetic field magnitude |B|\absolutevalue{B}, we can plot out the required BzB_{z} and BxB_{x} alongside the projected gp,eg_{p,e} (Fig. 5). We are mostly interested in the regime 0<|B|0.20<\absolutevalue{B}\leq 0.2 T, as this regime lies below the Hc1H_{c_{1}} of Nb. Above this critical field, we would incur additional losses in the coupled system due to the presence of normal currents in the superconducting circuit. As higher Hc1H_{c_{1}} superconductors are explored as SC qubit materials, higher |B|\absolutevalue{B} regimes will become accessible to this scheme.

Refer to caption
Figure 5: Effect of the maximum applicable magnetic field on various parameters of the system. (a) Evolution of the BxB_{x} and BzB_{z} required to maintain 4.31 GHz spin splitting as a function of |B|\absolutevalue{B}. (b) Change in eigenfrequencies as a function of |B|\absolutevalue{B}, where ν1\nu_{1} and ν3\nu_{3} are the eigenfrequencies of |ψ1\ket{\psi_{1}} and |ψ3\ket{\psi_{3}} are the ground state qubit levels of interest. (c) Change in the components of |ψ1\ket{\psi_{1}} and |ψ3\ket{\psi_{3}} with |B|\absolutevalue{B}, indicating greater spin-orbit mixing as the maximum applicable magnetic field increases. (d) projected gp,eg_{p,e} vs |B|\absolutevalue{B} as determined by Eq. (32).

Now, we must consider the coupling between the superconducting circuit and the electron spin to all acoustic modes supported by the piezoelectric resonator. The Hamiltonian of describing this interaction can be written as

H^=ωsc2σ^scz+kωp,kb^kb^k+ωe2σ^ez+kgsc,p;k(σ^sc++σ^sc)(b^k+b^k)+kgp,e;k(σ^e++σ^e)(b^k+b^k).\frac{\hat{H}}{\hbar}=\frac{\omega_{sc}}{2}\hat{\sigma}^{z}_{sc}+\sum_{k}\omega_{p,k}\hat{b}_{k}^{\dagger}\hat{b}_{k}+\frac{\omega_{e}}{2}\hat{\sigma}^{z}_{e}+\sum_{k}g_{sc,p;k}\left(\hat{\sigma}^{+}_{sc}+\hat{\sigma}^{-}_{sc}\right)\left(\hat{b}_{k}+\hat{b}_{k}^{\dagger}\right)+\sum_{k}g_{p,e;k}\left(\hat{\sigma}^{+}_{e}+\hat{\sigma}^{-}_{e}\right)\left(\hat{b}_{k}+\hat{b}_{k}^{\dagger}\right). (35)

where the index kk labels each acoustic mode and ωp,0\omega_{p,0} is the frequency of the resonator mode of interest. We can shift into a interaction picture by applying the transformation H^=U^H^U^+iU^˙U^\hat{H}^{\prime}=\hat{U}\hat{H}\hat{U}^{\dagger}+i\dot{\hat{U}}\hat{U}^{\dagger}, where U^=exp[i(ωsc2σ^scz+kωp,kb^kb^k+ωe2σ^ez)t]\hat{U}=\exp\left[i\left(\frac{\omega_{sc}}{2}\hat{\sigma}_{sc}^{z}+\sum_{k}\omega_{p,k}\hat{b}_{k}^{\dagger}\hat{b}_{k}+\frac{\omega_{e}}{2}\hat{\sigma}_{e}^{z}\right)t\right]. This transformation gives

H^=kgsc,p;k(ei(ωscωp,k)tσ^sc+b^k+ei(ωscωp,k)tσ^scb^k)+kgp,e;k(ei(ωeωp,k)tσ^e+b^k+ei(ωeωp,k)tσ^eb^k),\frac{\hat{H}^{\prime}}{\hbar}=\sum_{k}g_{sc,p;k}\left(e^{i(\omega_{sc}-\omega_{p,k})t}\hat{\sigma}^{+}_{sc}\hat{b}_{k}+e^{-i(\omega_{sc}-\omega_{p,k})t}\hat{\sigma}^{-}_{sc}\hat{b}_{k}^{\dagger}\right)+\sum_{k}g_{p,e;k}\left(e^{i(\omega_{e}-\omega_{p,k})t}\hat{\sigma}^{+}_{e}\hat{b}_{k}+e^{-i(\omega_{e}-\omega_{p,k})t}\hat{\sigma}^{-}_{e}\hat{b}_{k}^{\dagger}\right), (36)

We would like to determine the conditions in which we can neglect all resonator modes except the mode of interest, which we will call k0k_{0} with frequency ωp,k0\omega_{p,k_{0}}. Let us first ignore the spin-phonon coupling and focus on the superconducting circuit-phonon coupling. In the interaction Hamiltonian in Eq. 36, we can see that when ωsc=ωp,k0\omega_{sc}=\omega_{p,k_{0}} (the frequency of the acoustic resonator mode of interest), Rabi oscillations will be induced between the two modes. We would also, however, like to consider the oscillations induced between the superconducting circuit and the other resonator modes. Let us select a different transformation H^2=U^2H^U^2+iU^˙2U^2\hat{H}^{\prime}_{2}=\hat{U}_{2}\hat{H}\hat{U}_{2}^{\dagger}+i\dot{\hat{U}}_{2}\hat{U}_{2}^{\dagger}, where U^2=exp[i(ωsc2σ^scz+k(ωp,k+Δp,k)b^kb^k)t]\hat{U}_{2}=\exp\left[i\left(\frac{\omega_{sc}}{2}\hat{\sigma}_{sc}^{z}+\sum_{k}\left(\omega_{p,k}+\Delta_{p,k}\right)\hat{b}_{k}^{\dagger}\hat{b}_{k}\right)t\right], where Δp,kωscωp,k\Delta_{p,k}\equiv\omega_{sc}-\omega_{p,k}, and ignore the electron spin-related terms. The resulting interaction Hamiltonian is

H^2=kΔp,kb^kb^k+kgsc,p;k(σ^sc+b^k+σ^scb^k).\frac{\hat{H}^{\prime}_{2}}{\hbar}=-\sum_{k}\Delta_{p,k}\hat{b}_{k}^{\dagger}\hat{b}_{k}+\sum_{k}g_{sc,p;k}\left(\hat{\sigma}^{+}_{sc}\hat{b}_{k}+\hat{\sigma}^{-}_{sc}\hat{b}_{k}^{\dagger}\right). (37)

The Heisenberg equations of motion for σ^sc\hat{\sigma}_{sc} and b^k\hat{b}_{k} are

σ^˙sc\displaystyle\dot{\hat{\sigma}}^{-}_{sc} =i[H^2,σ^sc]\displaystyle=-\frac{i}{\hbar}\left[\hat{H}_{2}^{\prime},\hat{\sigma}^{-}_{sc}\right] (38)
=κsc2σ^scigsc,p;k0b^k0ikk0gsc,p;kb^k,\displaystyle=-\frac{\kappa_{sc}}{2}\hat{\sigma}^{-}_{sc}-ig_{sc,p;k_{0}}\hat{b}_{k_{0}}-i\sum_{k\neq k_{0}}g_{sc,p;k}\hat{b}_{k}, (39)
b^˙k\displaystyle\dot{\hat{b}}_{k} =i[H^2,b^k]\displaystyle=-\frac{i}{\hbar}\left[\hat{H}_{2}^{\prime},\hat{b}_{k}\right] (40)
=(iΔp,kκp,k2)b^k+igsc,p;kσ^sc\displaystyle=\left(-i\Delta_{p,k}-\frac{\kappa_{p,k}}{2}\right)\hat{b}_{k}+ig_{sc,p;k}\hat{\sigma}^{-}_{sc} (41)

where gsc,p;k0g_{sc,p;k_{0}} is the desired acoustic mode’s electromechanical coupling. In matrix form, this becomes

[σsc^b1^b2^bN^]˙=[κsc2igsc,p;1igsc,p;2igsc,p;Nigsc,p;1(iΔp,1κp,12)00igsc,p;20(iΔp,1κp,12)0igsc,p;N(iΔp,Nκp,N2)][σsc^b1^b2^bN^].\dot{\begin{bmatrix}\hat{\sigma_{sc}}\\ \hat{b_{1}}\\ \hat{b_{2}}\\ \vdots\\ \hat{b_{N}}\\ \end{bmatrix}}=\begin{bmatrix}-\frac{\kappa_{sc}}{2}&-ig_{sc,p;1}&-ig_{sc,p;2}&\ldots&-ig_{sc,p;N}\\ ig_{sc,p;1}&(-i\Delta_{p,1}-\frac{\kappa_{p,1}}{2})&0&\ddots&0\\ ig_{sc,p;2}&0&(-i\Delta_{p,1}-\frac{\kappa_{p,1}}{2})&\ddots&0\\ \vdots&\ddots&\ddots&\ddots&\vdots\\ ig_{sc,p;N}&\ldots&\ldots&\ldots&(-i\Delta_{p,N}-\frac{\kappa_{p,N}}{2})\end{bmatrix}\begin{bmatrix}\hat{\sigma_{sc}}\\ \hat{b_{1}}\\ \hat{b_{2}}\\ \vdots\\ \hat{b_{N}}\\ \end{bmatrix}. (42)

This is equivalent to inducing Rabi oscillations of various frequencies and suppressions between the SC qubit and acoustic modes. The probability amplitude of population transfer to each acoustic mode from an excited SC state becomes

σsc,k=4(gsc,p;k)24(gsc,p;k2)+|Δp,k+i(κsc+κp,k2)|2sin2(4(gsc,p;k)2+|Δp,k+i(κsc+κp,k2)|22t).\langle\sigma_{sc,k}\rangle=\frac{4(g_{sc,p;k})^{2}}{4(g_{sc,p;k}^{2})+\absolutevalue{\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)}^{2}}\sin^{2}\left(\frac{\sqrt{4\left(g_{sc,p;k}\right)^{2}+\absolutevalue{\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)}^{2}}}{2}t\right). (43)

This gives us a SC qubit probability of being in the excited state as a function of time is then

σ^sc=kσsc,k=k4(gsc,p;k)24(gsc,p;k)2+(Δp,k+i(κsc+κp,k2))2sin2(4(gsc,p;k)2+|Δp,k+i(κsc+κp,k2)|22t).\hat{\sigma}_{sc}=\sum_{k}\langle\sigma_{sc,k}\rangle=\sum_{k}\frac{4(g_{sc,p;k})^{2}}{4\left(g_{sc,p;k}\right)^{2}+\left(\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)\right)^{2}}\sin^{2}\left(\frac{\sqrt{4\left(g_{sc,p;k}\right)^{2}+\absolutevalue{\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)}^{2}}}{2}t\right). (44)

The sum over all σsc,k\langle\sigma_{sc,k}\rangle with kk0k\neq k_{0} is a worst-case bound on the probability amplitude that could escape the computational basis into undesired acoustic modes, limiting state fidelity. If the ratio of σsc,k0/kk0σsc,k1\langle\sigma_{sc,k_{0}}\rangle/\sum_{k\neq k_{0}}\langle\sigma_{sc,k}\rangle\gg 1, then we can effectively treat our system as having only one acoustic mode coupled to a SC qubit. The same physics governs the spin-phonon dynamics, replacing the appropriate couplings in equation (43) and (44).

Appendix B Details of Numerical Simulations

Simulations were completed using the finite element method (FEM) in COMSOL Multiphysics, utilizing the Electrostatics and Structural Mechanics modules. Stationary simulations were conducted to determine the electrostatic field applied to the piezoelectric transducer from a microwave source, and eigenfrequency simulations were conducted to determine the transducer’s resonant acoustic modes. Finally, gsc,pg_{sc,p} was calculated for each mode by determining the overlap between the piezoelectrically induced field and electrostatic field (see Eq.

Refer to caption
Figure 6: Electromechanical and spin-mechanical couplings and population transfer to each acoustic mode. (a) demonstrates a 56\sim 56 MHz frequency window (grey shaded region) in which our mode of interest (4.115\sim 4.115 GHz) lies. The couplings gsc,pig_{sc,p_{i}} and gp,eig_{p,e_{i}} are plotted for each mode, assuming a shunt capacitance CS130C_{S}\sim 130 fF and a magnetic field of 0.180.18 T. (b) displays the Rabi population transfer probability from the superconducting circuit and electron spin to each acoustic mode (see Eq. (43)), showing a combined mode suppression (diamond markers) of at least three orders of magnitude

The coupling parameters gsc,pg_{sc,p} and gp,eg_{p,e} were then calculated using the combination of these two simulations (see Eqs.  (8) and  (9)). The parameter set with the best mode isolation (see Fig. 6 featured λ=1370\lambda=1370 nm and wres=465w_{res}=465 nm. This device was then tethered using the phononic tethers shown in Fig. 2 and the number of tethers were varied to calculate mechanical quality factor as a function of number of tether periods, shown in Fig. 7(b).

Refer to caption
Figure 7: FEM simulation of the piezoelectric transducer with phononic tethers and surrounding bulk treated as perfectly matched layers (PML) to simulate clamping quality factor QcQ_{c}. (a) Diagram of the setup with free variable Ntethers,yN_{tethers,y} identified and (b) plot of QcQ_{c} vs Ntethers,yN_{tethers,y} for the 4.11 GHz resonator mode of interest.

Appendix C Analysis of Spin Register System

In Figure 8, we present a roadmap to scaling this architecture to form a memory register for superconducting circuits. Since the shunt capacitance far exceeds the capacitance of a single IDT, additional electromechanical resonators in parallel to a single transmon qubit do not significantly change the coupling rates to each resonator. Individual control over each resonator can be obtained with (i.e. cryo-MEMS) electrical switching of contacts to each resonator [75]. If this is not possible, controls can still be obtained in the frequency domain if each resonator frequency is sufficiently detuned from all others and within the tunability range of the transmon. This gives two constraints on the number of parallel resonators we can add: the maximum number of resonators before gsc,pg_{sc,p} for each resonator drops below a desired value, and the maximum number of resonators before the frequency spectrum becomes overcrowded.

From electrostatic simulations in COMSOL, CS70CIDTC_{S}\approx 70C_{IDT}, allowing us to add around 10 resonators in parallel without decreasing the coupling to each resonator by more than 15%. Additionally, each resonator can house several quantum emitters, which themselves will be operating at different frequencies ωe,ij\omega_{e,ij} due to differing magnetic field and strain environments creating a unique Zeeman effect for each color center. Assuming one implants nn emitters in each resonator, this creates an easily accessible m×nm\times n register of ancillas for a single transmon.

We would like to evaluate overcrowding of the frequency spectrum in this picture. In an ideal case, when we tune the superconducting circuit on resonance with a mechanical mode ωm\omega_{m}, we would like the circuit to be approximately coupled only to that acoustic mode. This is the same condition as we presented in Appendix A to assume that we can simplify the dynamics of the SC-phonon-spin system to that of coupling via a single acoustic mode. Thus, when the condition for every mode jj, then we can suppose that we individually couple to one piezoelectric resonator out of a number of resonators (see Fig. 8). Similarly, we would like to determine the condition where we can assume each piezoelectric resonator can individually couple to a single spin. This complicates the second stage of the system in Appendix A. Assuming that the conditions in Appendix A already holds for each of mm resonators coupled to the SC qubit, the full Hamiltonian describing the mm resonator, m×nm\times n spin system is

H=ωsc2σ^scz+i=1mj=1n[ωp,ia^p,ia^p,i+ωeij2σ^eijz+gsc,pi(σ^sc+a^p,i+σ^sca^p,i)+gp,eij(σ^eij+a^p,i+σ^eija^p,i)].H_{\sum}=\frac{\omega_{sc}}{2}\hat{\sigma}^{z}_{sc}+\sum_{i=1}^{m}\sum_{j=1}^{n}\Bigg{[}\omega_{p,i}\hat{a}_{p,i}^{\dagger}\hat{a}_{p,i}+\frac{\omega_{e_{ij}}}{2}\hat{\sigma}^{z}_{e_{ij}}+g_{sc,p_{i}}\left(\hat{\sigma}^{+}_{sc}\hat{a}_{p,i}+\hat{\sigma}^{-}_{sc}\hat{a}_{p,i}^{\dagger}\right)+g_{p,e_{ij}}\left(\hat{\sigma}^{+}_{e_{ij}}\hat{a}_{p,i}+\hat{\sigma}^{-}_{e_{ij}}\hat{a}_{p,i}^{\dagger}\right)\Bigg{]}. (45)

Following exactly from Eqs. 43 and 44 in Appendix A, the required condition for assuming electromechanical coupling to just the k0k_{0}th of mm resonators is that

kk0mσsc,k=kk0m4(gsc,p;k)24(gsc,p;k2)+|Δp,k+i(κsc+κp,k2)|2sin2(4(gsc,p;k)2+|Δp,k+i(κsc+κp,k2)|22t)σsc,k0.\sum_{k\neq k_{0}}^{m}\langle\sigma_{sc,k}\rangle=\sum_{k\neq k_{0}}^{m}\frac{4(g_{sc,p;k})^{2}}{4(g_{sc,p;k}^{2})+\absolutevalue{\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)}^{2}}\sin^{2}\left(\frac{\sqrt{4\left(g_{sc,p;k}\right)^{2}+\absolutevalue{\Delta_{p,k}+i\left(\frac{\kappa_{sc}+\kappa_{p,k}}{2}\right)}^{2}}}{2}t\right)\ll\langle\sigma_{sc,k_{0}}\rangle. (46)

Similarly, after swapping population into one of the resonator modes, the condition for assuming spin-mechanical coupling to just the j0j_{0}th of nn electron spins is that

jj0nσe,j=jj0n4(gp,e;j)24(gp,e;j2)+|Δp,j+i(κe+κp,j2)|2sin2(4(gp,e;j)2+|Δp,j+i(κe+κp,j2)|22t)σe,j0.\sum_{j\neq j_{0}}^{n}\langle\sigma_{e,j}\rangle=\sum_{j\neq j_{0}}^{n}\frac{4(g_{p,e;j})^{2}}{4(g_{p,e;j}^{2})+\absolutevalue{\Delta_{p,j}+i\left(\frac{\kappa_{e}+\kappa_{p,j}}{2}\right)}^{2}}\sin^{2}\left(\frac{\sqrt{4\left(g_{p,e;j}\right)^{2}+\absolutevalue{\Delta_{p,j}+i\left(\frac{\kappa_{e}+\kappa_{p,j}}{2}\right)}^{2}}}{2}t\right)\ll\langle\sigma_{e,j_{0}}\rangle. (47)
Refer to caption
Figure 8: Scaling the schematic to a quantum memory register. By implanting nn emitters in each of mm detuned mechanical resonators in parallel with the supercondcuting qubit of interest, one can create an efficient interface to an m×nm\times n optically addressable ancilla register.

We can see from the spin-phonon coupling points in Fig. 6 that frequency crowding can begin to promote Rabi oscillations with populations on the order of 10310^{-}3 of the desired mode when within a 100 MHz frequency window. So parallelization of spins in one resonator would require changing the local magnetic field for each resonator and intelligent spacing of the emitters to promote a wide distribution of resonant frequencies, or sacrificing state tansfer fidelity to a single spin by overcrowding the simulated frequency window of operation. This is not as much of a problem given the order-of-magnitude superior mode suppression on the electromechanical side of the system. Thus, we can comfortably parallelize around 1010 piezoelectric resonators to a single SC qubit and 1-3 emitters per resonator. When accounting for the surrounding C13{}^{13}C nuclear spins, we envision that this scaling method can provide a SC qubit with a 10+ nuclear spin memory register.