21
An integrated nanophotonic quantum register based on silicon-vacancy spins in diamond C. T. Nguyen, 1, * D. D. Sukachev, 1, * M. K. Bhaskar, 1, * B. Machielse, 1, 2, * D. S. Levonian, 1, * E. N. Knall, 2 P. Stroganov, 1 C. Chia, 2 M. J. Burek, 2 R. Riedinger, 1 H. Park, 1, 3 M. Lonˇ car, 2 and M. D. Lukin 1, 1 Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA 2 John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, USA 3 Department of Chemistry and Chemical Biology, Harvard University, Cambridge, Massachusetts 02138, USA We realize an elementary quantum network node consisting of a silicon-vacancy (SiV) color center inside a diamond nanocavity coupled to a nearby nuclear spin with 100 ms long coherence times. Specifically, we describe experimental techniques and discuss effects of strain, magnetic field, mi- crowave driving, and spin bath on the properties of this 2-qubit register. We then employ these techniques to generate Bell-states between the SiV spin and an incident photon as well as between the SiV spin and a nearby nuclear spin. We also discuss control techniques and parameter regimes for utilizing the SiV-nanocavity system as an integrated quantum network node. I. INTRODUCTION Quantum networks have the potential to enable a plethora of new technologies including secure commu- nication, enhanced metrology, and distributed quantum computing [1–5]. Such networks require nodes which per- form quantum processing on a small register of intercon- nected qubits with long coherence times. Distant nodes are connected by efficiently interfacing qubits with op- tical photons that can be coupled into an optical fiber [Fig. 1(a)]. Si C 13 C 12 (a) (b) Node Quantum Network MW FIG. 1. (a) Schematic of a quantum network. Nodes con- sisting of several qubits are coupled together via an optical interface. (b) A quantum network node based on the SiV. SiV centers and ancilla 13 C are incorporated into a nanophotonic device and addressed with a coupled fiber and microwave coplanar waveguide. * These authors contributed equally to this work [email protected] [email protected] The prevailing strategy for engineering an efficient, co- herent optical interface is that of cavity quantum electro- dynamics (QED), which enhances the interactions be- tween atomic quantum memories and photons [6–10]. Nanophotonic cavity QED systems are particularly ap- pealing, as the tight confinement of light inside opti- cal nanostructures enables strong, high-bandwidth qubit- photon interactions [11–13]. In practice, nanophotonic devices also have a number of technological advantages over macroscopic optical cavities, as they can be fab- ricated en-masse and interfaced with on-chip electron- ics and photonics, making them suitable for scaling up to large-scale networks [9, 14]. While strong interac- tions between single qubits and optical photons have been demonstrated in a number of cavity QED platforms [9, 10, 15–18], no single realization currently meets all of the requirements of a quantum network node. Simultane- ously achieving high-fidelity, coherent control of multiple long-lived qubits inside of a photonic structure is a major outstanding challenge. Recent work has established the silicon-vacancy color- center in diamond (SiV) as a promising candidate for quantum networking applications [19–24]. The SiV is an optically active point defect in the diamond lattice [25, 26]. Its D 3d inversion symmetry results in a vanish- ing permanent electric dipole moment of the ground and excited states, rendering the transition insensitive to elec- tric field noise typically present in nanostructures [27]. Recent work has independently shown that SiV centers in nanostructures display strong interactions with single photons [22] and that SiV centers at temperatures below 100 mK (achievable in dilution refrigerators) exhibit long coherence times [20, 28]. While these results indicate the promising potential of the SiV center for future quantum network nodes, significant technical challenges must be overcome in order to combine these ingredients. In this paper, we outline the practical considerations and approaches needed to build a quantum network node with SiV centers in nanophotonic diamond cavities cou- arXiv:1907.13200v2 [quant-ph] 1 Aug 2019

diamond - arXiv · 2019. 8. 2. · computing [1{5]. Such networks require nodes which per-form quantum processing on a small register of intercon-nected qubits with long coherence

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • An integrated nanophotonic quantum register based on silicon-vacancy spins indiamond

    C. T. Nguyen,1, ∗ D. D. Sukachev,1, ∗ M. K. Bhaskar,1, ∗ B. Machielse,1, 2, ∗ D. S. Levonian,1, ∗ E. N. Knall,2

    P. Stroganov,1 C. Chia,2 M. J. Burek,2 R. Riedinger,1 H. Park,1, 3 M. Lončar,2 and M. D. Lukin1, †

    1Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA2John A. Paulson School of Engineering and Applied Sciences,Harvard University, Cambridge, Massachusetts 02138, USA

    3Department of Chemistry and Chemical Biology,Harvard University, Cambridge, Massachusetts 02138, USA

    We realize an elementary quantum network node consisting of a silicon-vacancy (SiV) color centerinside a diamond nanocavity coupled to a nearby nuclear spin with 100 ms long coherence times.Specifically, we describe experimental techniques and discuss effects of strain, magnetic field, mi-crowave driving, and spin bath on the properties of this 2-qubit register. We then employ thesetechniques to generate Bell-states between the SiV spin and an incident photon as well as betweenthe SiV spin and a nearby nuclear spin. We also discuss control techniques and parameter regimesfor utilizing the SiV-nanocavity system as an integrated quantum network node.

    I. INTRODUCTION

    Quantum networks have the potential to enable aplethora of new technologies including secure commu-nication, enhanced metrology, and distributed quantumcomputing [1–5]. Such networks require nodes which per-form quantum processing on a small register of intercon-nected qubits with long coherence times. Distant nodesare connected by efficiently interfacing qubits with op-tical photons that can be coupled into an optical fiber[Fig. 1(a)].

    Si

    C13 C12

    (a)

    (b)

    Node

    ↑〉

    ↓〉 ↑〉

    ↓〉 ↑〉

    ↑〉Quantum

    Network

    MW

    FIG. 1. (a) Schematic of a quantum network. Nodes con-sisting of several qubits are coupled together via an opticalinterface. (b) A quantum network node based on the SiV. SiVcenters and ancilla 13C are incorporated into a nanophotonicdevice and addressed with a coupled fiber and microwavecoplanar waveguide.

    ∗ These authors contributed equally to this [email protected][email protected]

    The prevailing strategy for engineering an efficient, co-herent optical interface is that of cavity quantum electro-dynamics (QED), which enhances the interactions be-tween atomic quantum memories and photons [6–10].Nanophotonic cavity QED systems are particularly ap-pealing, as the tight confinement of light inside opti-cal nanostructures enables strong, high-bandwidth qubit-photon interactions [11–13]. In practice, nanophotonicdevices also have a number of technological advantagesover macroscopic optical cavities, as they can be fab-ricated en-masse and interfaced with on-chip electron-ics and photonics, making them suitable for scaling upto large-scale networks [9, 14]. While strong interac-tions between single qubits and optical photons havebeen demonstrated in a number of cavity QED platforms[9, 10, 15–18], no single realization currently meets all ofthe requirements of a quantum network node. Simultane-ously achieving high-fidelity, coherent control of multiplelong-lived qubits inside of a photonic structure is a majoroutstanding challenge.

    Recent work has established the silicon-vacancy color-center in diamond (SiV) as a promising candidate forquantum networking applications [19–24]. The SiV isan optically active point defect in the diamond lattice[25, 26]. Its D3d inversion symmetry results in a vanish-ing permanent electric dipole moment of the ground andexcited states, rendering the transition insensitive to elec-tric field noise typically present in nanostructures [27].Recent work has independently shown that SiV centersin nanostructures display strong interactions with singlephotons [22] and that SiV centers at temperatures below100 mK (achievable in dilution refrigerators) exhibit longcoherence times [20, 28]. While these results indicate thepromising potential of the SiV center for future quantumnetwork nodes, significant technical challenges must beovercome in order to combine these ingredients.

    In this paper, we outline the practical considerationsand approaches needed to build a quantum network nodewith SiV centers in nanophotonic diamond cavities cou-

    arX

    iv:1

    907.

    1320

    0v2

    [qu

    ant-

    ph]

    1 A

    ug 2

    019

    mailto:[email protected]:[email protected]

  • 2

    pled to ancillary nuclear spins [Fig. 1(b)] [29]. SectionII describes recent improvements to the fabrication tech-niques used to create and incorporate SiV centers intohigh-quality factor, critically-coupled nanophotonic cav-ities with an efficient fiber-optical interface. Section IIIdescribes the millikelvin experimental apparatus and sev-eral common experimental protocols. Section IV de-scribes the SiV level structure and electronic transitions,illusutrating the interplay of strain and magnetic field inenabling both coherent control of– and a photonic inter-face for– SiV spins. Sections V, VI and VII outline exper-imental implementations of optical and microwave con-trol of SiV centers, and use this control to create electron-photon Bell states with high fidelity in section VIII. Sec-tion IX introduces techniques for coupling to additionalqubits consisting of naturally occuring 13C in diamond.We describe our method for initializing and reading outthese nuclear spins via the SiV, coherent control of 13Cwith microwave and radio-frequency driving, probe thecoherence of these nuclei, and finally entangle the SiVwith a nearby 13C and demonstrate electron-nuclear Bellstates.

    (a)

    II III VI

    (b)

    FIG. 2. (a) Schematic of the nanofabricataion process usedto produce devices. I: Titanium-HSQ mask is patterned us-ing EBL. II: Pattern is transferred onto diamond using topdown O2 RIE. III: Angled IBE is used to separate structuresfrom substrate. IV: Devices are covered in PMMA and im-plantation aperatures are formed using EBL. Device are thencleaned, implanted, and annealed. V: PMMA is used in a litoffprocedure to pattern gold microwave striplines. VI: Final de-vices are cleaned and prepared for experiment. (b) Scanningelectron micrographs corresponding to steps II, III, and VI inthe fabricaton procedure.

    II. NANOPHOTONIC DEVICE FABRICATION

    A. Device design

    The devices used in these experiments integratenanophotonic cavities, implanted SiV centers, and mi-crowave coplanar waveguides onto a single diamond chip.Here we present the fabrication process used to realizesuch devices.

    Typically, high-quality photonic crystal resonators arefabricated from 2-D membranes, which tightly confinelight due to total internal reflection off of material bound-aries. Difficulties in growing high-purity, single-crystaldiamond films on non-diamond substrates are one ofthe key challenges to fabricating such resonators in dia-mond [30]. As a result, nanophotonic diamond structuresmust be etched out of bulk diamond, which requires non-traditional etching techniques [31, 32]. In particular, twomethods have emerged for creating freestanding diamondnanostructures: Isotropic undercutting [32, 33] and an-gled ion-beam etching (IBE) [34]. In this work, we usethe latter technique, resulting in freestanding, triangular-cross-section waveguides.

    Preliminary design of the nanophotonic structures aredescribed in appendix A, and are optimized to maximizeatom-photon interaction while maintaining high waveg-uide coupling. To take advantage of the scalable natureof nanofabrication, these optimized devices are patternedin sets of roughly 100 with slightly modified fabricationparameters. The overall scale of all photonic crystal cav-ity parameters are varied between different devices on thesame diamond chip to compensate for fabrication errors(which lead to unexpected variations in the resonator fre-quency and quality-factor). Due to these errors, roughlyone in six cavities are suitable for SiV experiments. For-tunately, hundreds of devices are made in a single fabri-cation run, ensuring that every run yields many usabledevices.

    The diamond waveguide region (as opposed to the pho-tonic crystal cavity region [Appendix. A]) has two dis-tinguishing features. First, thin support structures areplaced periodically along the waveguide and are used tosuspend the structures above the substrate. These sup-ports are portions of the waveguide which are adiabat-ically tapered to be ∼ 30% wider than the rest of thewaveguide, and take longer to etch away during the an-gled etch process. By terminating the etch after normalwaveguide regions are fully etched through, these widesections become ∼ 10 nm thick supports which tetherthe waveguide structures to the substrate while minimiz-ing scattered loss from guided modes. Second, one endof the waveguide structure is adiabatically tapered intofree-space [35]. These tapers are formed by a linear ta-per of the waveguide down to less than 50 nm wide overa 10 µm length. This tapered region can be coupled toa similarly tapered optical fiber, allowing structures toefficiently interface with a fiber network [Sec. III]. Thistapered end of the waveguide is the most fragile por-

  • 3

    tion of the structure, and can break after repeated fibercoupling attempts. This is often what limits the totalmeasurement lifetime of a device.

    The number of devices (and thus the relative yield ofthe fabrication process) is limited by the maximum pack-ing density on the diamond chip. This is primarily lim-ited by the need to accommodate 10 µm wide microwavecoplanar waveguides (CPWs) between devices, which arepatterned directly onto the diamond surface to efficientlycontrol SiV spins using microwaves. Simulations (Son-net Inc) of prospective design geometries ensure that theCPW is impedance matched with our 50 Ω feed lines,which minimizes scattered power from the waveguides.Tapers in the CPW near the center of the cavity regionsconcentrate current and increase the amplitude of the mi-crowave field near the SiVs, and CPWs are terminatedwith a short in order to ensure a magnetic field maximumalong the device.

    B. Device fabrication

    Fabrication of the diamond structures proceeds as de-scribed in ref. [35] with the notable modification that theangled etch is conducted not with a Faraday cage loadedinside a reactive ion etching chamber, but instead with anIBE. The Faraday cage technique [31, 36] offered the ben-efit of simplicity and accessibility—requiring only thatthe reactive ion etching chamber in question was largeenough to accommodate the cage structure—but sufferedfrom large fluctuations in etch rate across the surface ofthe sample, as well as between different fabrication runs,due to imperfections in the Faraday cage mesh. Theseirregularities could be partially compensated for by re-peatedly repositioning and rotating the cage with respectto sample during the etch, but this process proved to belaborious and imprecise. Instead, IBE offers collimatedbeams of ions several cm in diameter, leading to almostuniform etch rates across the several mm diamond chip.This technique allowed for consistent fabrication of cavi-ties with Q > 104, V < 0.6[λ/(n = 2.4)]3, and resonanceswithin ∼ 10 nm of SiV optical frequencies.

    Once the diamond cavities are fabricated [Fig. 2(a I-III)], SiV centers must be incorporated. To ensure thebest possible atom-photon interaction rate [Sec. V], SiVsshould be positioned at the cavity mode maximum. Ide-ally, this requires implantation accuracy of better than50 nm in all 3 dimensions due to the small mode vol-ume (∼ 0.5[λ/(n = 2.4)]3) of the cavities used. In thepast, implantation of silicon ions (which form SiV cen-ters following a high-temperature anneal) was done us-ing focused ion-beam implantation, but this techniquerequired specialized tools and lacked the accuracy neces-sary for maximally efficient mode coupling [13]. Instead,we adapt the standard masked implantation techniqueand use commercial foundaries for ion implantation.

    For the implantation process, we repeatedly spin andbake MMA EL11 and PMMA C4 (Microchem) to cover

    the nanophotonic cavities completely with polymer re-sist. We then spin-coat a conductive surface layer of Es-pacer (Showa Denko). An E-beam lithography (EBL)tool then aligns with large markers underneath the poly-mer layer, allowing it to expose an area surroundingsmaller, high-resolution alignment markers on the dia-mond. The exposed regions are developed in a 1:3 mix-ture of MIBK:IPA. Espacer is again spin-coated, anda second EBL write can be done, aligned to the high-resolution markers. Based on these alignment markers,holes of less than 65 nm diameter (limited by the reso-lution of PMMA resist) are patterned onto the center ofthe photonic crystal cavity which, after subsequent devel-opment, act as narrow apertures to the diamond surface[Fig. 2(a IV)]. The rest of the diamond surface is stillcovered in sufficiently thick PMMA to prevent ions fromreaching masked portions of the device. Diamonds arethen sent to a commercial foundry (Innovion) where theyare implanted with silicon ions at the appropriate energyand dose [Fig. 2 (b)]. Annealing in a UHV vacuum fur-nace (Kurt-Lesker) at ∼1400 K converts these implantedions into SiV centers [27, 37].

    CPWs are fabricated using a liftoff process similar tothat used to create masked implantation windows. Themost notable difference is an additional oxygen plasmadescum after development to remove PMMA residuefrom the surface. Following development, a 10 nm ti-tanium film serves as an adhesion layer for a 250 nmthick gold CPW [Fig. 2 (a V)]. Liftoff is performed inheated Remover PG (Microchem) [Fig. 2 (a VI)]. Themetal thicknesses used here are chosen to improve adhe-sion of the gold, as well as prevent absorption of cavityphotons by the metallic CPW. We observe that the cav-ity quality factor significantly degrades with gold films >300 nm. Due to ohmic heating, which can degrade the co-herence properties of SiV spins [Sec. VI], the length of theCPW is constrained to address a maximum of roughly 6devices.

    Future improvements in diamond device performancewill be predicated on improvements of the fabricationtechnology. Device quality factors are currently limitedby deviations in device cross section caused by imperfectselectivity of the HSQ hard mask to oxygen etching. Re-placing this mask with a sufficiently smooth metal maskcould result in improved etch selectivity and device per-formance. Isotropic undercut etching could also lead toimproved control over device cross sections and facili-tate more sophisticated device geometries [33, 38] at thecost of reduced control over isotropically etched surfaceroughness. Various techniques exist for the formation ofsmaller implantation apertures [39, 40], but these tech-niques are difficult to use in conjunction with implanta-tion into completed nanophotonic devices. Finally, theuse of superconducting striplines could reduce heating,which would enable the CPW to potentially address alldevices on the diamond chip and allow for faster drivingof SiV spin and nuclear transitions [Sec. VI, IX].

  • 4

    N.C.

    50:50 fiberbeam splitter

    90:10 fiberbeam splitter

    RT

    VC

    N2

    MXC

    4K

    740nm

    CCD

    (a) (b)

    N.C.

    Spec.

    SPCM

    SPCM

    Excitation

    12

    35

    6

    7

    8

    9

    10

    114

    4

    FIG. 3. (a) Experiment schematic. Devices 1 are mountedin the bore of a SC magnet 2 inside of a dilution refrigerator,and imaged with wide-field imaging 3 and piezo steppers4 . Devices are addressed with a tapered optical fiber 5

    positioned using a second set of piezo steppers 6 . Cavitiesare tuned using a nitrogen 7 . (b) Fiber network used toprobe devices. Excitation light is monitored 8 and sent tothe device. Collected light is monitored 9 and filtered 10

    then sent to one or several SPCMs 11 . N.C. indicates noconnection.

    III. EXPERIMENTAL SETUP

    Experiments are performed in a home-built photonic-probe setup inside of a dilution refrigerator (DR, Blue-Fors BF-LD250) [Fig. 3(a)]. The diamond substrate ismounted to a gold-plated copper sample holder via in-dium soldering below the mixing chamber in the bore ofa (6,1,1) T superconducting vector magnet (AmericanMagnetics Inc.) anchored to the 4 K stage. A ther-mal link between the device and the mixing chamberplate is provided by gold-plated copper bars, as well asoxygen-free copper braids (Copper Braid Products), en-suring maximal thermal conductivity between the mixingchamber plate and the sample, which reaches a base tem-perature of roughly 60 mK. We address single nanopho-tonic devices via a tapered optical fiber, which can becoupled in-situ with collection efficiencies exceeding 90%[35]. The tapered fiber is mounted to a 3-axis piezo step-per (ANPx101, ANPz101), and imaged in free-space byan 8f wide-field scanning confocal microscope which fo-cuses onto a cryo-compatible objective (Attocube LT-APO-VISIR). This setup allows for coupling to severalcavities during a single cooldown.

    Once coupled, the cavity resonance is red-shifted vianitrogen gas condensation [22]. A copper tube is weaklythermalized with the 4 K plate of the DR and can beheated above 80 K in order to flow N2 gas onto the de-vices. This gas condenses onto the photonic crystal, mod-ifying its refractive index and red-shifting the cavity res-onance. When the copper tube is not heated, it thermal-

    izes to 4 K, reducing the blackbody load on the sampleand preventing undesired gas from leaking into the vac-uum chamber.

    After red-tuning all devices in this way, each cavitycan be individually blue-tuned by illuminating the de-vice with a ∼100 µW broadband laser via the taperedfiber, locally heating the device and evaporating nitro-gen. This laser-tuning can be performed very slowly toset the cavity resonance with a few GHz. The cavity tun-ing range exceeds 10 nm without significantly degradingthe cavity quality factor, and is remarkably stable insidethe DR, with no observable drift over several months ofmeasurements.

    In previous work [22], SiVs were probed in transmis-sion via the free-space confocal microscope focused ontoa notch opposing the tapered fiber. Mechanical vibra-tions arising from the DR pulse tube (∼1 µm pointingerror at the sample position) result in significant fluctu-ations in power and polarization of incoupled light. Inthis work, we demonstrate a fully integrated solution byutilizing the same tapered fiber to both probe the deviceand collect reflected photons. This approach stabilizesthe excitation path and improves the efficiency of theatom-photon interface, allowing for deterministic interac-tions with single itinerant photons. High-contrast reflec-tion measurements are enabled by the high-cooperativity,critically-coupled atom-cavity system. Resonant light issent via the fiber network [Fig. 3(b)] and reflected off ofthe target device. We pick off a small fraction (∼ 10%)of this signal and use it to monitor the wide-band re-flection spectrum on a spectrometer (Horiba iHR-550) aswell as calibrate the coupling efficiency to the nanocav-ity. The remaining reflection is then routed either di-rectly to a single-photon counting module (SPCM, Ex-celitas SPCM-NIR), or into a time-delay interferometerfor use in spin-photon experiments [Sec. VIII]. Due to thishigh-efficiency fiber-coupled network, we observe overallcollection efficiencies of ∼ 40%, limited by the quantumefficiency of our APDs.

    IV. OPTIMAL STRAIN REGIMES FOR SIVSPIN-PHOTON EXPERIMENTS

    Similar to other solid state emitters [41, 42], the SiVis sensitive to local inhomogeneity in the host crystal. Inthe case of the SiV, which has D3d symmetry, the dom-inant perturbation is crystal strain. In this section, wedescribe the effects of strain on the SiV spin and opticalproperties, and how they can enable efficient microwaveand optical control of SiV centers inside nanostructures.

    A. SiV Hamiltonian in the presence of strain andspin-orbit coupling

    The SiV electronic structure is comprised of spin-orbiteigenstates split by spin-orbit interactions. Optical tran-

  • 5

    Bext

    α

    ↑’〉

    ↓〉 ↑〉

    ↓’〉

    f↓↓’

    f↑↑’

    73

    7n

    m

    LB

    LB’

    f↑↓

    (a)

    (d)

    (b)

    (f )

    0 60Time (min)

    α π/20

    α π/20

    5

    0.2

    UB

    UB’

    ∆gs

    SiV 1 (∆gs~500GHz)SiV 2 (∆gs~140GHz)

    -0.01-1

    0.011

    δf ↑

    ↓ (%

    )

    δf ↑

    ↑’ (

    %)

    δf↑↑’

    (GHz)

    (c)SiV 1SiV 2

    00

    0.1

    1.2

    Bext (T)

    f ↑↑’-f ↓

    ↓’ (

    GH

    z)

    SiV 3

    SiV 3 (∆gs~85GHz)

    00

    20

    0.7Bext (T)

    f ↑↓

    (GH

    z)

    0.4

    -0.60 0.2

    (e)10-300

    0.6

    δf↑↓

    (MHz)

    Pro

    bab

    ility

    Pro

    bab

    ility

    SiV 1SiV 2

    6

    1

    SiV 1SiV 2

    FIG. 4. (a) SiV level diagram. Optical transitionsf↑↑′ , f↓↓′ ∼737 nm are coupled to a nanophotonic cavity withmean detuning ∆. Microwaves at frequency f↑↓ drive rota-tions in the lower branch (LB). (b) Qubit frequency f↑↓ fordifferently strained emitters. Modeled splitting for groundstate g-factors ggs1 = 1.99, ggs2 = 1.89, ggs3 = 1.65 (solidlines) based on independent measurements of ∆gs. (inset) An-gle dependence of f↑↓ at fixed field Bext = 0.19 T. Solid linesare predictions using the same model parameters. (c) Opti-cal splitting f↑↑′ − f↓↓′ . Fits extract excited state g-factorsges1 = 1.97, ges2 = 1.83, ges3 = 1.62 (solid lines). (inset) Angledependence of f↑↑′ − f↓↓′ at fixed field Bext = 0.1 T. (d) His-togram of MW transition frequency for two different emitters.(e) Histogram of Optical transition frequency for two differ-ent emitters. (f) Simultaneous measurement of f↑↓ and f↑↑′reveals correlations between optical and microwave spectraldiffusion for emitter 2.

    sitions connect the ground state manifold (LB, UB) andexcited state manifold (LB′, UB′) [Fig. 4(a)]. In a DR,phonon absorption LB → UB (and LB′ → UB′) is su-pressed, resulting in thermal polarization into LB.

    We consider the ground state SiV Hamiltonian withspin-orbit and strain interactions, in the combined orbitaland spin basis {|ey ↑〉, |ey ↓〉, |ex ↑〉, |ex ↓〉} [23, 25]

    HSiV = HSO +Hstrain (1)

    =

    α− β 0 γ − iλ 00 α− β 0 γ + iλγ + iλ 0 α+ β 00 γ − iλ 0 α+ β

    (2)where α corresponds to axial strain, β and γ correspondto transverse strain, and λ is the strength of spin-orbitinteraction. Diagonalizing this reveals the orbital char-

    acter of the lower branch:

    LB ∝

    |ex ↑〉 −1+√

    1+(γ/β)2+(λ/β)2

    γ/β−iλ/β |ey ↑〉|ex ↓〉 − 1−

    √1+(γ/β)2+(λ/β)2

    γ/β−iλ/β |ey ↓〉(3)

    We investigate these electronic levels in the context ofthe SiV as a spin-photon interface.

    B. Effects of strain on the SiV qubit states

    In the limit of zero crystal strain, the orbital factorssimplify to the canonical form [25]

    LB =

    {|e+ ↓〉|e− ↑〉

    (4)

    In this regime, the spin-qubit has orthogonal electronicorbital and spin components. As result, one would needto simultaneously drive an orbital and spin flip to ma-nipulate the qubit, which is forbidden for direct mi-crowave driving alone. Thus, in the low strain regime,two-photon optical transitions between the qubit statesin a misaligned external field, already demonstrated atmillikelvin temperatures in [21], are likely necessary torealize a SiV spin qubit.

    In the high strain limit (√β2 + γ2 � λ), these orbitals

    become

    LB =

    {(cos(θ/2)|ex〉 − sin(θ/2)|ey〉)⊗ | ↓〉(cos(θ/2)|ex〉 − sin(θ/2)|ey〉)⊗ | ↑〉

    (5)

    where tan(θ) = βγ . In this regime, the ground state or-

    bital components are identical, and the qubit states canbe described by the electronic spin degree of freedomonly. As such, the magnetic dipole transition betweenthe qubit states is now allowed and can be efficientlydriven with microwaves.

    In addition to determining the efficiency of qubit tran-sitions, the spin-orbit nature of the SiV qubit states alsodetermines its susceptibility to external fields. In an ex-ternally applied magnetic field, LB splits due to mag-netic moments associated with both spin– and orbital–angular momenta. This splitting is parameterized by aneffective g-tensor which, for a fixed angle between theexternal field and the SiV symmetry axis, can be sim-plified to an effective g-factor: µ gBext/h = f↑↓. In thelimit of large strain, the orbital component of the twoLB wavefunctions converge, and g trends towards thatof a free electron (g = 2). As a result, the qubit statesbehave akin to a free-electron in the high strain regime,and there is no dependence of g on external field angleor small perturbations in crystal strain.

    While it is difficult to probe β or γ directly, they re-late to the energy difference between UB and LB via

    ∆gs = 2√β2gs + γ

    2gs + λ

    2gs [Fig. 4(a)]. From this, we ex-

    tract√β2 + γ2, given the known value of λgs =46 GHz

  • 6

    [25, 26, 43]. Numerically diagonalizing the SiV Hamil-tonian using the extracted values for β and γ closelymatches the measured ground state splitting, both as afunction of applied field magnitude and angle [Fig. 4(b)].

    C. Effects of strain on the SiV spin-photoninterface

    Strain also plays a crucial role in determining the opti-cal interface to the SiV spin qubit. The treatment shownabove can be repeated for the excited states, with thecaveat that the parameters β, γ, and λ are different in theexcited state manifold as compared to the ground statemanifold [23]. These differences give rise to a differentg-factor in the excited state (ges). If the strain is muchlarger than both λgs =46 GHz and λes =255 GHz, thenggs ≈ ges ≈ 2. In this case, the two spin-cycling transi-tion frequencies f↑↑′ and f↓↓′ are identical, and the onlyspin-selective optical transitions are the dipole-forbiddenspin-flipping transitions f↑↓′ and f↓↑′ .

    Under more moderate strain, the difference δg = |ges−ggs| splits the degenerate optical transitions f↑↑′ and f↓↓′ ,making them spin-selective as well. Due to differences inthe anisotropic g-tensor in the ground and excited states,δg depends on the orientation of the magnetic field aswell, and is minimized in the case of a 〈111〉-aligned field[Fig 4(c), inset]. In such an external field aligned withthe SiV symmetry axis, optical transitions become highlyspin-conserving [20], allowing many photons to scatterwithout altering the SiV spin state. This high cyclicityenables high-fidelity single-shot readout of the spin state[29], even without high collection efficiencies [20]. Thismakes working with the spin-cycling transitions highlydesirable, at the expense of a reduced ability to resolvespin-selective transitions for a given field magnitude. Theneed to resolve individual transitions suggests an optimal

    strain regime where√β2gs + γ

    2gs � λgs, where MW driv-

    ing is efficient, while√β2es + γ

    2es . λes, where one can

    independently address f↑↑′ and f↓↓′ [Fig. 4(c)].

    D. Effects of strain on SiV stability

    Despite the SiV’s symmetry-protected optical transi-tions, spectral diffusion of the SiV has been observed inmany experiments [27, 44] (but still much smaller com-pared to emitters without inversion symmetry, for exam-ple, nitrogen-vacancy centers [45, 46]). While the exactnature of this diffusion has not been studied in depth,it is often attributed to the second-order Stark effect orstrain fluctuations, both of which affect the energies ofSiV orbital wavefunctions. In this paper, we also observesignificant fluctuations of the spin qubit frequency.

    As can be seen in reference [23], for an appropriatelylow static strain value, fluctuating strain can give rise tofluctuations in the g-tensor of the ground state, causing

    spectral diffusion of the qubit frequency f↑↓ [Fig. 4(d)].Since ggs asymptotically approaches 2 as the static strainincreases [23], the qubit susceptibility to this fluctuatingstrain is reduced in the case of highly strained SiV cen-ters, resulting in a more stable qubit.

    While spectral diffusion of the optical transition shouldnot saturate in the same way as diffusion of themicrowave transition, we observe qualitatively differ-ent spectral diffusion properties for different emitters[Fig. 4(e) and Fig. 12]. SiV 1 (∆gs = 500 GHz) dis-plays slow drift of the optical line which is stable to 500 kHz) of the mi-crowave transition for this SiV. On the other hand, SiV2 (∆gs = 140 GHz) drifts over a wider range, and also ex-hibits abrupt jumps between several discrete frequencies[Appendix. C].

    We simultaneously record the optical transition andqubit frequency for SiV 2 and observe correlations be-tween the two frequencies [Fig. 4(f)], indicating that theycould arise from the same environmental perturbation.In Appendix B, we calculate the qubit and optical tran-sition frequencies using the strain Hamiltonian (eq. 2)and find that both correlations and absolute amplitudesof spectral diffusion can simultaneously be explained bystrain fluctuations on the order of 1% (∼ 10−7 strain)[Appendix B].

    In this work we rely on static strain, likely resultingfrom damage induced by ion implantation and nanofab-rication, and select for spectrally stable SiVs with ap-propriate strain profiles. This is characterized by firstmeasuring ∆gs in zero magnetic field at 4 K by excitingthe optical transition LB → LB′ and measuring emis-sion from the LB′ → UB on a spectrometer. We usethis to screen for SiVs with ∆gs >100 GHz to ensureefficient MW driving of the spin qubit. We further ap-ply a static external magnetic field and measure spectralstability properties as well as f↑↑′ − f↓↓′ to guarantee agood spin-photon interface. We measured∼ 10 candidateemitters, and found 4 which satisfy all of the necessarycriteria for spin-photon experiments.

    V. REGIMES OF CAVITY-QED FOR SIVSPIN-PHOTON INTERFACES

    Efficient spin-photon interactions are enabled by incor-porating SiV centers into nanophotonic cavities. In thissection, we describe SiV-cavity measurements in severalregimes of cavity QED, and comment on their viabilityfor spin-photon experiments.

    A. Spectroscopy of cavity-coupled SiVs

    We measure the spectrum of the atom-cavity systemat different atom-cavity detunings in order to character-ize the device and extract key cavity QED parameters

  • 7

    (a)

    ω-ωc (GHz/2π)

    1

    0

    Refle

    ctio

    n

    (d)

    (e)

    1

    Refle

    ctio

    n

    0

    Prob

    abili

    ty

    Photons in 13μs

    F = 0.92

    30

    Laser Detuning (GHz) 4-4

    20-100

    Bare Cavityωa-ωc = 0κωa-ωc = 0.3κωa-ωc = 2.7κ

    1

    Refle

    ctio

    n

    0 1Laser Detuning (GHz)-1

    (b)

    (c)

    Photons in 30μs

    F = 0.97

    Prob

    abili

    ty

    263

    Bext

    α = 0 α = π/2

    Bext

    f↓↓’

    f↓↓’f

    ↑↑’

    f↑↑’

    FIG. 5. (a) SiV-cavity reflection spectrum at several de-tunings. The bare cavity spectrum (black) is modulated bythe presence of the SiV. When the atom cavity detuing issmall (Blue, orange), high-contrast, broad features are theresult of Purcell enhanced SiV transitions. Far from thecavity resonance (green), interaction results in narrow SiV-assisted transmission channels. (b) Spin-dependent reflectionfor large SiV-cavity detuning ∆ ≈ −3κ, Bext = 0.35 T. Inthis regime, SiV spin states can be individually addressed.(c) Probing either transmission dip results in high-fidelitysingle-shot readout in an aligned field (F = 0.97, thresholdon detecting 13 photons). (d) Spin-dependent reflection nearresonance ∆ ≈ 0.5κ, Bext = 0.19 T. Dispersive lineshapes al-low for distinguishable reflection spectra from both SiV spinstates. (e) A probe at the frequency of maximum contrast(fQ) can determine the spin state in a single shot in a mis-aligned field(F = 0.92, threshold on detecting > 1 photon).

    [Fig. 5(a)]. The reflection spectrum of a two-level systemcoupled to a cavity is modeled by solving the frequencyresponse of the standard Jaynes-Cummings Hamiltonianusing input-output formalism for a cavity near criticalcoupling [7]:

    R(ω) =∣∣∣∣1− 2κli(ω − ωc) + κtot + g2/(i(ω − ωa) + γ)

    ∣∣∣∣2 ,(6)

    where κl is the decay rate from the incoupling mirror, κtotis the cavity linewidth, ωc(ωa) is the cavity (atom) res-onance frequency, g is the single-photon Rabi frequency,and γ is the bare atomic linewidth. Interactions betweenthe SiV optical transition and the nanophotonic cavityresult in two main effects. First, the SiV center can mod-ulate the reflection spectrum of the bare cavity, as seenin the colored curves of figure 5(a). Second, the couplingto the cavity can broaden the linewidth of the SiV based

    on the Purcell effect:

    Γ ≈ γ + 4g2/κ 11 + 4(ωc − ωa)2/κ2

    When the cavity is far detuned from the atomic transition|ωc−ωa| ≡ ∆ > κ [Fig. 5(a), green], Purcell enhancementis negligible and the cavity and atomic linewidths κ, γ =2π × {33, 0.1} GHz are estimated. When the cavity ison resonance with the atom (∆ = 0), we fit (6) usingpreviously estimated values of κ and γ to extract g =2π× 5.6 GHz. Together, these measurements allow us todetermine the atom-cavity cooperativity C = 4g2/κγ =38. Importantly, interactions between the SiV and singlephotons becomes deterministic when C > 1.

    As mentioned in section [Sec. IV], we would like tomake use of spectrally resolved spin conserving opticaltransitions (f↑↑′ , f↓↓′) to build a spin-photon interfaceusing the SiV. Here, we make this criteria more explicit:f↑↑′ and f↓↓′ can be resolved when |f↑↑′ − f↓↓′ | & Γ.

    B. Cavity QED in the detuned regime

    In the detuned regime (∆ > κ), Γ ≈ γ, and nar-row atom-like transitions are easily resolved under mostmagnetic field configurations, including when the field isaligned with the SiV symmetry axis [Fig. 5(b)]. In thiscase [sec. IV] [20], optical transitions are highly spin-conserving, and many photons can be collected allow-ing for high-fidelity single-shot readout of the SiV spinstate (F = 0.97) [Fig. 5(c)]. Rapid, high-fidelity, non-destructive single-shot readout can enable projective-readout based initialization: after a single measurementof the SiV spin state, the probability of a measurement-induced spin flip is low, effectively initializing the spininto a known state.

    While this regime is useful for characterizing the sys-tem, the maximum fidelity of spin-photon entanglementbased on reflection amplitude is limited. As seen in fig-ure 5(b), the contrast in the reflection signal between anSiV in |↑〉 (orange) vs. |↓〉 (purple) is only ∼ 80%, imply-ing that in 20% of cases, a photon is reflected from thecavity independent of the spin state of the SiV, resultingin errors. We note that the residual 20% of reflectioncan be compensated by embedding the cavity inside aninterferometer at the expense of additional technical sta-bilization challenges, discussed below.

    C. Cavity QED near resonance

    Tuning the cavity onto the atomic resonance (∆ ≈ 0)dramatically improves the reflection contrast [Fig. 5(a)(blue curve)]. Here, we observe nearly full contrast ofthe reflection spectrum due to the presence of the SiV.Unfortunately, this is associated with a broadened atomiclinewidth (Γ = γ(1 + C) ∼4 GHz). While it is, in prin-ciple, still possible to split the atomic lines by going to

  • 8

    higher magnetic fields, there are several technical con-siderations which make this impractical. Large magneticfields (|Bext| >0.5 T) correspond to large qubit frequen-cies (f↑↓), which can induce spontaneous qubit decay dueto phonon emission (|↑〉 → |↓〉), as well as increased localheating of the device from microwave dissipation, bothof which reduce the SiV spin coherence time rendering itineffective as a quantum memory.

    At intermediate detunings (0 < ∆ < κ), the SiV reso-nance is located on the cavity slope and results in high-contrast, spin-dependent Fano lineshapes which exhibitsharp features smaller than Γ [Fig. 5(a), orange curve].By working at an optimal Bext where the peak of onespin transition is overlapped by the valley of the other,the best features of the resonant and far-detuned regimesare recovered [Fig. 5(e)]. Probing the system at the pointof maximum contrast (fQ ≈ (|f↑↑′ − f↓↓′ |)/2, contrast> 90%) enables single-shot readout of the SiV spin statefor an arbitrary field orientation, even when transitionsare not cycling [Fig. 5(f)].

    This demonstrates an optical regime of cavity QEDwhere we simultaneously achieve high-contrast readoutwhile maintaining spin-dependent transitions. In thisregime, we still expect residual reflections of about 10%,which end up limiting spin-photon entanglement fidelity.This infidelity arises because the cavity is not perfectlycritically coupled (κl 6= κtot/2), and can in principle besolved by engineering devices that are more critically cou-pled. Alternatively, this problem can be addressed forany cavity by interfering the signal with a coherent ref-erence to cancel unwanted reflections. In this case, onewould have to embed the cavity in one arm of a stabi-lized interferometer. This is quite challenging, as it in-volves stabilizing ∼ 10 m long interferometer arms, partof which lie inside the DR (and experience strong vibra-tions from the pulse-tube cryocooler).

    A fundamental issue with critically coupled cavities isthat not all of the incident light is reflected from the de-vice. If the spin is not initialized in the highly-reflectingstate, photons are transmitted and not recaptured intothe fiber network. Switching to overcoupled (single-sided) cavities, where all photons are reflected with aspin-dependent phase, could improve both the fidelityand efficiency of spin-photon entanglement. Once again,however, measurement of this phase would require em-bedding the cavity inside of a stabilized interferometer.As such, the un-compensated reflection amplitude basedscheme employed here is the most technically simple ap-proach to engineering spin-photon interactions.

    VI. MICROWAVE SPIN CONTROL

    While the optical interface described in previous sec-tions enables high-fidelity initialization and readout ofthe SiV spin qubit, direct microwave driving is the moststraightforward path towards coherent single-qubit ro-tations. Typically, microwave manipulation of electron

    (a)

    (e)

    10

    0.2

    1

    〈Sz〉

    ΩR = 8 MHz

    ΩR = 17 MHzΩR = 10 MHz

    τ (μs)

    RT VC4K

    MXC

    1

    3

    5

    6

    24

    7

    (d)

    TM

    ax

    1

    10

    0

    ΩR = 18 MHz

    〈Sz〉

    0 50τ (μs)

    ΩR = 50 MHzMax(TSiV)

    (b)

    Gold

    τth TS

    iV (A

    rb. U

    nit

    s)

    0.35

    010 t (ms)

    (c)

    τ = 10μs

    τ = 30μs

    TMax(10μs)

    FIG. 6. (a) Experimental schematic for microwave control.The amplitude and phase of a CW microwave source 1 aremodulated via a microwave switch and IQ mixer controlledexternally by an AWG 2 . A CW radio frequency source 3 iscontrolled using a digital delay generator 4 . Both signals areamplified by 30dB amplifiers 5 before entering the DR. 0dBcryo-attenuators 6 thermalize coax cables at each DR stage,ultimately mounted to a PCB 7 on the sample stage anddelivered to the devices. (b) Schematic depicting microwave-induced heating of devices. (c) Modeled temperature at theSiV from a dynamical decoupling sequence. At long τ , devicecools down between each decoupling pulse, resulting in lowtemperatures. At short τ , devices are insufficiently cooled,resulting in a higer max temperature (Tmax). (d) Effects ofmicrowave heating on SiV coherence time. (Top panel) Athigh Rabi frequencies, SiV coherence is temporarily reducedfor small τ . (Bottom panel) The local temperature (Tmax) atthe SiV calculated by taking the maximum value of the plotsin figure (c). (e) Hahn-echo for even lower Rabi frequencies,showing coherence times that scale with microwave power

    spins requires application of significant microwave power.This presents a challenge, as SiV spins must be keptat local temperatures below 500 mK in order to avoidheating-related dephasing. In this section, we implementcoherent microwave control of SiV centers inside nanos-tructures at temperatures below 500 mK.

    A. Generating microwave single-qubit gates

    The SiV spin is coherently controlled using amplitudeand phase controlled microwave pulses generated by aHittite signal generator (HMC-T2220). A target pulsesequence is loaded onto an arbitrary waveform generator(Tektronix AWG 7122B), which uses a digital channel tocontrol a fast, high-extinction MW-switch (Custom Mi-crowave Components, CMCS0947A-C2), and the analogchannels adjust the amplitude and phase via an IQ-mixer(Marki, MMIQ-0416LSM). The resulting pulse train issubsequently amplified (Minicircuits, ZVE-3W-183+) toroughly 3 W of power, and sent via a coaxial cable intothe dilution refrigerator. At each cryogenic flange, a 0 dB

  • 9

    attenuator is used to thermalize the inner and outer con-ductors of the coaxial line while minimizing microwavedissipation. The signal is then launched into a coplanarwaveguide on a custom-built circuit board (Rogers4003C,Bay Area Circuits) so it can be wire-bonded directly tothe diamond chip [Sec. II, Fig. 6(c)]. The qubit frequency(f↓↑) is measured by its optically detected magnetic res-onance spectrum (ODMR) identically to the method de-scribed in [20]. We observe ODMR from 2 GHz to 20 GHz(corresponding to fields from 0.1 T to 0.7 T), implyingthat microwave control of SiV centers in this configura-tion is possible at a wide variety of external field magni-tudes. This allows the freedom of tuning the field to op-timize other constraints, such as for resolving spin tran-sitions [Sec. V] and identifying ancillary nuclear spins[Sec. IX].

    Once the qubit frequency has been determined for agiven field, single-qubit gates are tuned up by measur-ing Rabi oscillations. The frequency of these oscillationsscales with the applied microwave power ΩR ∼

    √P and

    determines the single-qubit gate times. We can performπ-pulses (Rπφ) in under 12 ns, corresponding to a Rabi

    frequency exceeding 80 MHz [29]. This coherent controlis used to implement pulse-error correcting dynamicaldecoupling sequences, either CPMG-N sequences of the

    form Rπ/2x −

    (τ −Rπy − τ

    )N −Rπ/2x = x−(Y )N −x[47] orXY8-N sequences of the form x− (XYXY Y XYX)N −x[48]. Sweeping the inter-pulse delay τ measures the co-herence time T2 of the SiV.

    B. Effects of microwave heating on coherence

    As mentioned in sections III and IV, thermally in-duced T1 relaxation can dramatically reduce SiV coher-ence times. To explain this phenomenon, we model thenanobeam as a 1D beam weakly coupled at two anchorpoints to a uniform thermal bath [Fig. 6(b)]. Initially,the beam is at the steady-state base temperature of theDR. A MW pulse instantaneously heats the bath, and thebeam rethermalizes on a timescale τth set by the thermalconduction of diamond and the beam geometry. Oncethe pulse ends, this heat is extracted from the beam ona similar timescale. By solving the time-dependent 1-Dheat equation, we find that the change in temperatureat the SiV caused by a single pulse (starting at time t0)scales as TSiV ∝ (e−(t−t0)/τth−e−9(t−t0)/τth). We take thesum over N such pulses to model the effects of heatingfrom a dynamical-decoupling sequence of size N .

    At early times (τ < τth), the SiV does not see the ef-fects of heating by the MW line, and coherence is high.Similarly, at long times (τ � τth) a small amount ofheat is able to enter the nanostructure and slightly raisethe local temperature, but this heat is dissipated beforethe next pulse arrives [Fig. 6(c), blue curve]. At inter-mediate timescales however, a situation can arise wherethe nanobeam has not fully dissipated the heat from oneMW pulse before the second one arrives [Fig. 6(c), orange

    curve]. We plot the maximum temperature as seen by theSiV as a function of pulse spacing [Fig. 6(d), lower panel],and observe a spike in local temperature for a specificinter-pulse spacing τ , which depends on τth. Dynamical-decoupling sequences using high Rabi frequency pulsesreveal a collapse in coherence at a similar time [Fig. 6(d),upper panel]. This collapse disappears at lower Rabi fre-quencies, suggesting that it is associated with heating-related dephasing. We fit this collapse to a model wherethe coherence time T2 depends on temperature [28], andextract the rate of heating τth = 70 µs.

    Typically, faster π-pulses improve measured spin co-herence by minimizing finite-pulse effects and detuningerrors. Unfortunately, as seen above, faster pulses requirehigher MW powers which cause heating-related decoher-ence in our system. We measure Hahn-echo at lower MWpowers [fig. 6(e)], and find MW heating limits T2 even atΩR ∼10 MHz. For applications where long coherence isimportant, such as electron-nuclear gates [Sec. IX], weoperate at an optimal Rabi frequency ΩR = 2π×10 MHzwhere nuclear gates are as fast as possible while main-taining coherence for the entire gate duration. For appli-cations such as spin-photon entangling gates where fastgates are necessary [Sec. VIII], we operate at higher Rabifrequencies ΩR = 2π×80 MHz at the cost of reduced co-herence times.

    Heating related effects could be mitigated by usingsuperconducting microwave waveguides. This approachwould also enable the fabrication of a single, long su-perconducting waveguide that could simultaneously ad-dress all devices on a single chip. However, it is stillan open question whether or not superconducting waveg-uides with appropriate critical temperature, current, andfield properties can be fabricated around diamond nanos-tructures.

    VII. INVESTIGATING THE NOISE BATH OFSIVS IN NANOSTRUCTURES

    At low temperatures, the coherence time of SiV cen-ters drastically depends on the surrounding spin bath,which can differ from emitter to emitter. As an example,we note that the T2 of two different SiV centers in dif-ferent nanostructures scales differently with the numberof applied decoupling pulses [Fig. 7(a)]. Surprisingly, thecoherence time of SiV 2 does not scale with the num-ber of applied pulses, while the coherence time of SiV 1does scale as T2(N) ∝ N2/3. Notably, both scalings aredifferent as compared to what was previously measuredin bulk diamond: T2(N) ∝ N1 [20]. In this section, weprobe the spin bath of these two SiVs in nanostructuresto investigate potential explanations for the above obser-vations.

  • 10

    (a)

    Con

    tras

    t

    (c)

    (b)

    (d)

    0.05

    0.5T 2

    (ms)

    Spin-EchoDEER Echo

    12.4

    0

    1

    fDEER (GHz)

    00.01

    1

    1τ (ms)

    〈Sz〉

    N = 64

    N = 16N = 32

    N = 8N = 4N = 2

    12.0

    〈Sz〉

    -1

    0

    0 70τ (µs)

    242 Number of pulses

    SiV:

    bath:

    Init π2π2π

    fDEER

    Read

    SiV:

    bath:

    Init π2π2π Read

    π-τ- -τ-

    SiV 1SiV 2

    FIG. 7. (a) T2 scaling for two different SiVs. SiV 2 exhibitsno scaling with number of pulses (T2,SiV2 =30 µs). (b) DEERESR on SiV 2. Vertical red line is the expected frequency ofa g = 2 spin based on our ability to determine the appliedexternal field (Typically to within 10%). (c) DEER Echo onSiV 2. T2,DEER =10 µs. (d) Dynamical-decoupling on SiV 1.Data points are T2 measurements used in part (a, blue curve),and solid lines are a noise model consisting of two Lorentziannoise baths.

    A. Double electron-electron resonancespectroscopy of SiVs in nanostructures

    In order to investigate the poor coherence of SiV 2,we perform double electron-electron resonance (DEER)spectroscopy [49] to probe the spin bath surrounding thisSiV. We perform a Hahn-echo sequence on the SiV, andsweep the frequency of a second microwave pulse (takingthe RF path in figure 6(a)), contemporaneous with theechoing SiV π-pulse [Fig. 7(b), upper panel]. If this sec-ond pulse is resonant with a spin bath coupled to the SiV,the bath can flip simultaneously with the SiV, leadingto increased sensitivity to noise from the bath [Fig. 7(b),lower panel]. We observe a significant reduction of coher-ence at a frequency consistent with that of a free-electronspin bath (gbath = 2) (resonance expected at 12(1) GHz).

    Next, we repeat a standard Hahn-echo sequence wherea π-pulse resonant with this bath is applied simultane-ously with the SiV echo pulse (DEER echo). The coher-ence time measured in DEER echo is significantly shorterthan for standard spin-echo, indicating that coupling tothis spin bath is a significant source of decoherence forthis SiV. One possible explanation for the particularly se-vere bath surrounding this SiV is a thin layer of alumina(Al2O3) deposited via atomic layer deposition on this de-vice in order to tune cavities closer to the SiV transitionfrequency. The amorphous oxide layer–or its interfacewith the diamond crystal–can be host to a large numberof charge traps, all located within ∼50 nm of this SiV.Unfortunately, we could not measure this device with-out alumina layer due to our inability to gas-tune thenanophotonic cavity close enough to the SiV resonance[Sec. III].

    These observations are further corroborated by DEERmeasurements in SiV 1, where the alumina layer was notused (only N2 was used to tune this cavity). In thisdevice, we observe longer coherence times which scaleT2(N) ∝ N2/3, as well as no significant signatures fromgbath = 2 spins using DEER spectroscopy. We fit thisscaling to a model consisting of two weakly-coupled spinbaths [Fig. 7(d), Appendix. D], and extract bath param-eters b1 =5 kHz, τ1 =1 µs, b2 =180 kHz, τ2 =1 ms, whereb corresponds to the strength of the noise bath, and τcorresponds to the correlation time of the noise [50, 51].

    While the source of this noise is an area of futurestudy, we find that the b2 term (likely due to bulk impuri-ties) is the dominant contribution towards decoherence inthe system [Appendix. D]. Removing this term from themodel results in coherence times up to a factor of 1000times larger than measured values. Higher-temperature[27] or in situ [52] annealing could potentially mitigatethis source of decoherence by eliminating paramagneticdefects such as vacancy clusters. Additionally, by accom-panying Si implantation with electron irradiation [53],SiV centers could be created more efficiently, and with re-duced lattice damage. Finally, working with isotopicallypurified diamond samples with very few 13C, a spin-1/2isotope of carbon, could also result in a reduced spin bath[20], [Appendix. D].

    VIII. SPIN-PHOTON ENTANGLEMENT

    The previous sections characterize the SiV as an effi-cient spin-photon interface and a quantum memory with

    (a) (b)

    (d)(c)

    0

    0.5

    ↑〉|e ↓〉|l ↑〉|l ↓〉|e

    |ψ = |-

    |+ →〉→〉|+ ←〉←〉0

    0.5|ψ = |+

    |ψ = |+

    |+ →〉→〉|+ ←〉←〉0

    0.5

    π2 πSpin:

    Photon:Z

    p,X

    p

    Zs,X

    s

    δt

    Init

    FIG. 8. (a) Experimental sequence for generating and ver-ifying spin-photon entanglement. A time-bin encoded qubitis reflected by the cavity, and both the SiV and the photonicqubits are measured in the Z and X bases. (b) Spin-photoncorrelations measured in the Z-Z basis. Light (dark) bars arebefore (after) correcting for known readout error associatedwith single-shot readout of the SiV. (c) Spin-photon corre-lations measured in the X-X basis. Bell-state preparationfidelity of F ≥ 0.89(3) and a concurrence C ≥ 0.72(7). (d)Preparation of second spin-photon Bell state. Changing thephase of the incoming photonic qubit prepares a Bell-statewith inverted statistics in the X basis.

  • 11

    long-lived coherence. Here, we combine these two prop-erties to demonstrate entanglement between a spin qubitand a photonic qubit. The mechanism for generating en-tanglement between photons and the SiV can be seen infigure 5(b,d): Depending on the spin state of the SiV,photons at the probe frequency are either reflected fromthe cavity and detected, or are transmitted and lost.

    A. Generating time-bin qubits

    We begin by explaining our choice of time-bin encod-ing for photonic qubits. One straightforward possibilityis to use the Fock state of the photon. However, it is ex-tremely challenging to perform rotations on a Fock state,and photon loss results in an error in the computationalbasis. Another, perhaps more obvious possibility is to usethe polarization degree of freedom. While the SiV spin-photon interface is not polarization selective (both spinstates couple to photons of the same polarization), onecould consider polarization based spin-photon entanglingschemes already demonstated in nanophotonic systems[54, 55]. However, this requires embedding the nanos-tructure inside of a stabilized interferometer, which hasa number of challenges [Sec. V]. In addition, it requirescareful fabrication of overcoupled, single-sided cavities(unlike the critically coupled diamond nanocavities usedhere [Sec. II]). As such, we believe time-bin encoding isa natural choice given the critically-coupled SiV-cavityinterface described here [Sec. V].

    These qubits are generated by passing a weak co-herent laser though a cascaded AOM, amplitude-EOM,and phase-EOM. The time-bins are shaped by an AWG-generated pulse on the amplitude-EOM, and are chosento be much narrower than the delay δt between time bins.We can choose to prepare arbitrary initial photonic statesby using the phase-EOM to imprint an optional phaseshift to the second bin of the photonic qubit. Since weuse a laser with Poissonian photon number statistics, weset the average photon number 〈nph〉 = 0.008� 1 usingthe AOM to avoid events where two photons are incidenton the cavity.

    Using this encoding, measurements in a rotated ba-sis (X-basis) become straightforward. We send thetime-bin qubit into an actively stabilized, unbalanced,fiber-based, Mach-Zender interferometer, where one armpasses through a delay line of time δt. With 25% prob-ability, |e〉 enters the long arm of the interferometer and|l〉 enters the short arm, and the two time bins interfereat the output. Depending on the relative phase betweenthe two bins, this will be detected on only one of thetwo arms of the interferometer output [Fig. 3(b)], corre-sponding to a measurement in the X basis of |±〉.

    B. Spin-photon Bell states

    We prepare and verify the generation of maximallyentangled Bell states between the SiV and a photonicqubit using the experimental sequence depicted in fig-ure 8(a). First, the SiV is initialized into a superpo-

    sition state |→〉 = 1/√

    2(|↑〉 + |↓〉). Then photons atfrequency fQ [Sec. V] are sent to the cavity, correspond-

    ing to an incoming photon state |+〉 = 1/√

    2(|e〉 + |l〉),conditioned on the eventual detection of only one pho-ton during the experiment run. Before any interac-tions, this state can be written as an equal superposition:Ψ0 = | →〉⊗|+〉 = 1/2(|e ↑〉+|e ↓〉+|l ↑〉+|l ↓〉). The firsttime bin is only reflected from the cavity if the the SiV isin state |↑〉, effectively carving out |e ↓〉 in reflection [56].A π-pulse on the SiV transforms the resulting state toΨ1 = 1/

    √3(|e ↓〉+ |l ↓〉+ |l ↑〉). Finally, reflection of the

    late time-bin off of the cavity carves out the state |l ↓〉,leaving a final entangled state Ψ2 = 1/

    √2(|e ↓〉 + |l ↑〉).

    To characterize the resulting state, we perform tomogra-phy on both qubits in the Z and X bases [Fig. 8(a)].

    In order to enable high-bandwidth operation and re-duce the requirements for laser and interferometric sta-bilization in generating and measuring time-bin qubits,it is generally beneficial to set δt as small as possible.The minimum δt is determined by two factors: First,each pulse must be broad enough in the time-domain(narrow enough in the frequency domain) so that it doesnot distort upon reflection off of the device. From figure5(d), the reflection specturm is roughly constant over a ∼100 MHz range, implying that ∼ nanosecond pulses aresufficient. The second consideration is that a microwaveπ-pulse must be placed between the two pulses. In thisexperiment, we drive fast (12 ns) π-pulses. As such, weset δt = 30 ns and use 5 ns optical pulses to satisfy thesecriteria.

    C. Spin-photon entanglement measurements

    For Z-basis measurements, photons reflected from thecavity are sent directly to a SPCM and the time-of-arrivalof the time-bin qubit is recorded. Afterwards, the SiVis read out in the Z-basis [Sec. V]. Single-shot readoutis calibrated via a separate measurement where the twospin-states are prepared via optical pumping and readout, and the fidelity of correctly determining the | ↑〉(|↓〉) state is F↑ = 0.85 (F↓ = 0.84), limited by the large0 component of the geometric distribution which governsphoton statistics for spin-flip systems [Sec. V]. In otherwords, since we work in a misaligned field in this exper-iment, the probability of a spin flip is high, making itsomewhat likely to measure 0 photons regardless of ini-tial spin state. Even before accounting for this known er-ror [Appendix. E], we observe clear correlations betweenthe photonic and spin qubits [Fig. 8(b), light-shading].Error bars for these correlation histograms (and the fol-

  • 12

    lowing fidelity calculations) are estimated by statisticalbootstrapping, where the scattered photon histograms(post-selected on the detection of |e〉 or |l〉) are randomlysampled in many trials, and the variance of that ensembleis extracted.

    Measurements in the X-basis are performed similarly.The photon is measured through an interferometer as de-scribed above, where now the detector path information

    is recorded for the overlapping time-bin. After a Rπ/2y

    pulse on the SiV, the scattered photon histograms againreveal significant correlations between the ‘+’ and ‘-’ de-tectors and the SiV spin state [fig. 8(c)]. By adding aπ-phase between the early and late time bins, we canprepare an orthogonal Bell state. Measured correlationsof this state are flipped in the X-basis [Fig. 8(d)].

    Measurements of this Bell state in the Z- and X-bases are used to estimate a lower bound on the fidelity:F = 〈Ψ+|ρ|Ψ+〉 ≥ 0.70(3) (F ≥ 0.89(3) after correctingfor readout errors) [Appendix. E]. The resulting entan-gled state is quantified by its concrrence C ≥ 0.42(6)(C ≥ 0.79(7) after correcting for readout errors) [Ap-pendix. E]. This high-fidelity entangled state between aphotonic qubit and a quantum memory is a fundamen-tal resource for quantum communication[2] and quantumcomputing schemes [5], and can be used, for example, todemonstrate heralded storage of a photonic qubit intomemory [29].

    IX. CONTROL OF SIV-13C REGISTER

    While demonstrations of a quantum node with a sin-gle qubit is useful for some protocol, nodes with severalinteracting qubits enable a wider range of applications,including quantum repeaters [57]. In this section, weintroduce additional qubits based on 13C naturally oc-curring in diamond [sec. VII].

    A. Coupling between the SiV and several 13C

    For all of the emitters investigated in section VI, weobserve collapses in the echo signal corresponding to en-tanglement with nearby nuclear spins [Fig. 9(a)]. As thediamond used in this work has 1% 13C [Sec. VII], wetypically observe several such nuclei, with all of their res-onances overlapping due to their second-order sensitiv-ity to hyperfine coupling parameters [29]. Consequently,during a spin echo sequence the SiV entangles with manynuclei, quickly losing coherence and resulting in a col-lapse to 〈Sz〉 = 0 [Fig. 9(a), left side]. If single 13Ccan be addressed however, this entanglement results incoherent population transfer and echo collapses whichcan, in some cases, completely flip the SiV spin state(〈Sz〉 = ±1). This entanglement forms the basis forquantum gates [Fig. 9(a), right side]. These gates canbe tuned by changing the alignment of Bext with respectto the hyperfine coupling tensor, or by using different

    (a)

    SiV:

    13C:Init=

    Rπ/2xInit

    Rφn↑,n↓

    Rπ/2y Init

    SiV:

    13C:Read=

    Rφn↑,n↓ Rφn↑,n↓

    Rπ/2y Rπ/2x

    (c)(b)

    -1

    〈Sz〉

    1

    1 80 7τ (μs)

    ~R3π/4±x

    ~Rπ/2±x

    ~R0±x

    (d)

    -1

    0.5

    2.83

    X

    Y

    2.87τ (μs)

    〈Ix〉

    - 10μs -SiV:

    13C:Init Read Init

    SiV:

    13C:Read-τ-Init Read

    Rπ/2x Rπ/2x

    InitSiV:

    13C:Init Read

    Rπ/2x Rπ/2xR

    π/2±x

    InitSiV:

    13C:Init Read

    Rπ/2xRπ/2±x

    (f ) (g)

    ↓e↓n〉 ↓e↑n〉 ↑e↓n〉 ↑e↑n〉

    Z basis

    0

    0.5

    ↓e↓n〉 ↓e↑n〉 ↑e↓n〉 ↑e↑n〉

    X basis

    0

    0.5

    -0.6

    0.6

    〈Ix〉

    0 5τ (ms)

    (e)

    〈Ix〉

    50 τ (μs) 54

    Rπ/2x

    Initialize ↑〉Initialize ↓〉

    ↑〉SiV inSiV in ↓〉

    Rφn↑,n↓

    FIG. 9. (a) XY8-2 spin echo sequence reveals coupling tonuclear spins. (Left panel) Collapses 〈Sx〉 = 0 at short timesindicate coupling to many nuclei. (Right panel) Collapses〈Sx〉 6= 0 at long times indicate conditional gates on a sin-gle nuclear spin. (b) Trajectory of 13C on the Bloch sphereduring a maximally entangling gate. Orange (purple) linescorrespond to the SiV initially prepared in state | ↑〉 (| ↑〉);transitions from solid to dashed lines represent flips of theSiV electronic spin during the gate. (c) Maximally entan-

    gling gates of the form Rφ~n↑, ~n↓ are used to initialize and read-out the two-qubit register. (d) Tuning up an initializationgate. Inter-pulse spacing τ for Init and Read gates are sweptto maximize polarization. Solid line is the modeled pulse se-quence using the hyperfine parameters extracted from (a). (e)Nuclear Ramsey measurement. Driving the 13C using com-posite gates on the SiV reveals T ∗2 =2.2 ms. (Inset) Orangepoints are coherent oscillations of the Ramsey signal due tohyperfine coupling to the SiV. (f) Electron-nuclear correla-tions measured in the ZZ-basis. Light (dark) bars are before(after) correcting for known errors associated with reading outthe SiV and 13C. (g) Electron-nuclear correlations measuredin the XX-basis. We estimate a Bell state preparation fidelityof F ≥ 0.59(4) and a concurrence C ≥ 0.22(9).

  • 13

    timings. Unfortunately, as a result of the complicatednuclear bath for this device, a majority of field orienta-tions and amplitudes only show collapses to 〈Sz〉 = 0.The highest fidelity nuclear gates demonstrated here arebased on echo resonances with the largest contrast which,crucially, were not commensurate with an aligned field.Thus, in this device, single 13C could only be isolated atthe cost of lower SSR fidelity [Sec. V].

    B. Initializing the nuclear spin

    Once a single nuclear spin is identified, resonances inspin-echo form the building block for quantum gates. Forexample, a complete flip of the SiV is the result of thenuclear spin rotating by π conditionally around the axes±X (Rπ±x,SiV−C), depending on the state of the SiV. Wecan vary the rotation angle of this pulse by choosing dif-ferent spacings τ between pulses [Fig. 9(a)], or by usingdifferent numbers of π-pulses. We find a maximally en-

    tangling gate (Rπ/2±x,SiV−C) by applying N = 8 π-pulsesseparated by 2τ = 2 × 2.859µs. This can be visualizedon the Bloch sphere in figure 9(b), where the state of theSiV (orange or purple) induces different rotations of the13C.

    A similarly constructed entangling gate (Rφ~n↑, ~n↓ , dis-cussed in Appendix F) is used to coherently map popu-lation from the SiV onto the nuclear spin or map popula-tion from the nuclear spin onto the SiV [Fig. 9(c)]. Thefidelity of these gates is estimated by polarizing the SiV,mapping the population onto the 13C, and waiting forT � T ∗2 (allowing coherence to decay) before mappingthe population back and reading out [Fig. 9(d)]. We findthat we can recover 80% of the population in this way,giving us an estimated initialization and readout fidelityof F = 0.9.

    Based on the contrast of resonances in spin-echo (also0.9), this is likely limited by entanglement with othernearby 13C for this emitter, as well as slightly sub-optimal choices for τ and N . Coupling to other 13Cresults in population leaking out of our two-qubit reg-ister, and can be improved by increasing sensitivity tosingle 13C, or by looking for a different emitter with adifferent 13C distribution. The misaligned external fieldfurther results in slight misalignment of the nuclear rota-tion axis and angle of rotation, and can be improved byemploying adapted control sequences to correct for theseerrors [58, 59].

    C. Microwave control of nuclear spins

    As demonstrated above, control of the 13C via compos-ite pulse sequences on the SiV is also possible. A max-imally entangling gate has already been demonstratedand used to initialize the 13C, so in order to build auniversal set of gates, all we require are unconditionalsingle-qubit rotations. This is done following reference

    0.2

    1

    0 3ΩRF (kHz)

    SiV

    cohe

    renc

    e

    (a)

    -0.3

    0.3

    〈Ix〉

    0 2.5τ (ms)

    (b)-τ-Init

    SiV:

    13C:ReadInit Read

    RF

    FIG. 10. (a) RF Rabi oscillations. Applying an RF tonedirectly drives nuclear rotations of a coupled 13C. (b) SiVcoherence in the presence of an RF drive. As the strengthof the RF drive is increased, local heating from the CPWreduces the SiV T2.

    [60], where unconditional nuclear rotations occur in spin-echo sequences when the inter-pulse spacing τ is halfwaybetween two collapses. For the following gates, we usean unconditional π/2-pulse composed of 8 π-pulses sep-arated by τ = 0.731 µs.

    We use this gate to probe the coherence time T ∗2 of the13C. After mapping population onto the nuclear spin, theSiV is re-initialized, and then used to perform uncondi-tional π/2-rotations on the 13C [Fig. 9(d)]. Oscillationsin the signal demonstrate Larmor precession of the nu-cleus at a frequency determined by a combination of theexternal field as well as 13C-specific hyperfine interac-tions [29], which are seen as the orange data points infigure 9(d). The green envelope is calculated by fittingthe oscillations and extracting their amplitude. The de-cay of this envelope T ∗2 =2.2 ms shows that the

    13C hasan exceptional quantum memory, even in the absence ofany dynamical decoupling.

    We characterize the fidelity of our conditional and un-conditional nuclear gates by generating and reading outBell states between the SiV and 13C [Appendix. E]. First,we initialize the 2-qubit register into one of the 4 eigen-states: {| ↑e↑N 〉, | ↑e↓N 〉, | ↓e↑N 〉, | ↓e↓N 〉}, then performa π/2-pulse on the electron to prepare a superpositionstate. Afterward, a CNOT gate, comprised of an uncon-ditional π/2 pulse followed by a maximally entanglinggate, prepares one of the Bell states |Ψ±〉, |Φ±〉 depend-ing on the initial state [Fig. 9 (e,f)]. Following the anal-ysis outlined in appendix E, we report an error correctedfidelity of F ≥ 0.59(4) and C ≥ 0.22(9), primarily limitedby our inability to initialize the 13C [29].

    D. Radio-frequency driving of nuclear spins

    The previous section demonstrated a CNOT gate be-tween SiV and 13C using composite MW pulses. Thisapproach has several drawbacks. First, the gate fidelityis limited by our ability to finely tune the rotation an-gle of the maximally entangled gate which can not bedone in a continuous fashion [see Fig.9(a)]. Second, thisgate requires a specific number of MW pulses and delaysbetween them, making the gate duration (∼50 µs in this

  • 14

    work) comparable to the SiV coherence time. Finally,this scheme relies on a second order splitting of individ-ual 13C resonances to resolve individual ones; residualcoupling to additional 13C limits the fidelity for a pulsesequence of given total length.

    Direct RF control [24] would be a simple way to make afast and high-fidelity CNOT gate since it would require asingle RF π-pulse on a nuclear spin transition [61]. Fur-thermore, since the nuclear spin transition frequenciesdepend on the hyperfine coupling to leading order, thesepulses could have higher 13C selectivity and potentiallyshorter gate duration.

    We use the RF port inside the DR [Sec. VI] to applyRF pulses resonant with nuclear spin transitions. Figure9(a) shows RF Rabi oscillations of the nuclear spin. Sincethe 13C gyromagnetic ratio is about 3 orders of magni-tude smaller compared to the SiV spin, RF driving ismuch less efficient than MW one and requires much morepower. To investigate local heating of the SiV [Sec. VI]we measured the SiV spin coherence contrast in spin-echo sequence right after applying off-resonant RF pulseof 100 µs at different power (calibrated via RF rabi oscil-lations) [Figure 9(b)]. Unfortunately, Even modest Rabifrequencies (ΩRF ∼1 kHz) result in 20% loss in SiV co-herence. Replacing the gold CWG used in this work bysuperconducting ones may solve heating issue and makeRF driving practically useful.

    X. CONCLUSION

    The SiV center in diamond has rapidly become a lead-ing candidate to serve as the building block of a futurequantum network. In this work, we describe the underly-ing technical procedures and optimal parameter regimesnecessary for utilizing the SiV-nanocavity system as aquantum network node. In particular, we discuss the ef-fect of static and dynamic strain on the properties of theSiV spin qubit and its optical interface, with direct appli-cation to quantum networking experiments. We demon-strate techniques for coherently controlling and interfac-ing SiV spin qubits inside of nanophotonic structures atmillikelvin temperatures to optical photons. Finally, weidentify and coherently control auxiliary nuclear spins,forming a nanophotonic two-qubit register.

    The work presented here and in the complementaryletter [29] illustrates the path towards the realization ofa first-generation quantum repeater based on SiV centersinside diamond nanodevices. We note that a key ingre-dient enabling future, large-scale experiments involvingseveral solid-state SiV-nanocavity nodes will be the in-corporation of strain tuning onto each device [62]. Pre-cise tuning of both the static and dynamic strain canovercome the limitations of inhomogeneous broadeningand spectral diffusion, and enable scalable fabrication ofquantum repeater nodes [Sec. IV].

    We would like to thank M. Markham, A. Bennett andD. Twitchen from Element Six Inc. for providing the di-

    amond substrates used in this work, as well as F. Jelezkoand R. Evans for insightful discussions. This work wassupported by DURIP grant No. N00014-15-1-28461234through ARO, NSF, CUA, AFOSR MURI, and ARL.M.K.B. and D.S.L were supported by DoD NDSEG, B.M.and E.N.K. were supported by NSF GRFP, and R.R. wassupported by the Alexander von Humboldt Foundation.Devices were fabricated at Harvard CNS, NSF award no.1541959.

    Appendix A: Nanophotonic cavity design

    We simulate and optimize our nanophotonic structuresto maximize atom-photon interactions while maintaininghigh waveguide coupling, which ensures good collectionefficiency for the devices. In particular, this requires op-timizing the device quality-factor to mode volume ratio,the relative rates of scattering into waveguide modes, andthe size and shape of the optical mode. Each of thesequantities are considered in a three-step simulation pro-cess (FDTD, Lumerical). We first perform a coarse pa-rameter sweep over all possible unit cells which define

    (a)

    (c)

    (d)

    (b)

    FIG. 11. (a) Unit cell of a photonic crystal cavity (boundedby black lines). Hx and Hy define the size and aspect ratioof the hole, a determines the lattice constant, and w setsthe waveguide width. (b) Electric field intensity profile ofthe TE mode inside the cavity, indicating strong confinementof the optical mode inside the waveguide. (c) Schematic ofphotonic crystal design. Blue shaded region is the bandgapgenerating structure, red shaded region represents the cavitystructure. (d) Plot of a, Hx, and Hy for the cavity shownin (c), showing cubic taper which defines the cavity region.All sizes are shown in fractions of anominal, the unperturbedlattice constant.

  • 15

    the photonic crystal gemometry and identify families ofbandgap-generating structures. These structures are thestarting point for a gradient ascent optimization proce-dure, which results in generating high quality-factor, lowmode volume resonators. Finally, the generated designsare modified to ensure efficient resonator-waveguide cou-pling.

    Optimization begins by exploring the full param-eter space of TE-like bandgap generating structureswithin our waveguide geometry. For hole-based cavities[Fig. 11(a)], this sweep covers a 5-dimensional parameterspace: The lattice constant of the unit cell (a), the holesize and aspect ratio (Hx and Hy), the device etch angle(θ) and the waveguide width (w). Due to the size of thisparameter space, we start by performing a low-resolutionsweep over all parameters, with each potential designsimulated by a single unit cell with the following bound-ary conditions: 4 perfectly matched layer (PML) bound-ary conditions in the transverse directions and 2 Blochboundary conditions in the waveguide directions. Theband structure of candidate geometries are determinedby sweeping the effective k-vector of the Bloch boundarycondition and identifying allowed modes. Using this tech-nique, families of similar structures with large bandgapsnear the SiV transition frequency are chosen for furthersimulation. Each candidate photonic crystal is also in-spected for the position of its optical mode maximum,ensuring that it has first-order modes concentrated in thecenter of the diamond, where SiVs will be incorporated[Fig. 11(b)].

    The second step is to simulate the full photonic crystalcavity design, focused in the regions of parameter spaceidentified in step one. This is done by selecting a fixedθ, as well as a total number of unit cells that define thestructure, then modifying the bandgap of the photoniccrystal with a defect region to form a cavity mode. Wedefine this defect using a cubic tapering of one (or sev-eral) possible parameters:

    A(x) = 1− dmax|2x3 − 3x2 + 1| (A1)

    where A is the relative scale of the target parameter(s)at a distance x from the cavity center, and dmax is thedefect depth parameter. Photonic Crystal cavities withmulti-parameter defects are difficult to reliably fabricate,therefore, devices used in this work have cavity defectgeometries defined only by variations in the lattice con-stant. The cavity generated by this defect is scored bysimulating the optical spectrum and mode profile andcomputing the scoring function F :

    F = min(Q,Qcutoff)/(Qcutoff × Vmode) (A2)

    Where Q is the cavity quality-factor, Qcutoff = 5× 105 isan estimated maximum realizable Q based on fabricationconstraints, and Vmode is the cavity mode volume. Basedon this criteria, we employ a gradient ascent process overall cavity design parameters (except θ and the total num-ber of unit cells) until F is maximized, or a maximum

    number of iterations has occurred. Due to the complexityand size of the parameter space, a single iteration of thisgradient ascent is unlikely to find the optimal structure.Instead, several candidates from each family of designsfound in step one are explored, with the best moving onto the final step of the simulation process. These surviv-ing candidates are again checked to ensure confinement ofthe optical mode in the center of the cavity structure andto ensure that the structures fall within the tolerances ofthe fabrication process.

    The final step in the simulation process is to modify theoptimized designs to maximize resonator-waveguide cou-pling. This is done by removing unit cells from the inputport of the device, which decreases the overall quality-factor of the devices in exchange for better waveguidedamping of the optical field. Devices are once again sim-ulated and analyzed for the fraction of light leaving theresonator through the waveguide compared to the frac-tion scattering into free-space. The number of unit cellson the input port is then optimized for this ratio, withsimulations indicating that more than 95% of light is col-lected into the waveguide. In practice, fabrication de-fects increase the free-space scattering rate, placing res-onators close to the critically-coupled regime. Finally,the waveguide coupling fraction is increased by append-ing a quadratic taper to both ends of the devices suchthat the optical mode is transferred adiabatically fromthe photonic crystal region into the diamond waveguide.This process produces the final cavity structure used forfabrication [Fig. 11(c)].

    Appendix B: Strain-induced frequency fluctuations

    In this Appendix we calculate changes the SiV spin-qubit frequency and optical transition frequency arisingfrom strain fluctuations. We start with the Hamiltonianfor SiV in an external magnetic field Bz aligned alongtrhe SiV symmetry axis [23, 25]:

    H = −λ

    0 0 i 00 0 0 −i−i 0 0 00 i 0 0

    ︸ ︷︷ ︸

    spin-orbit

    +

    α− β 0 γ 00 α− β 0 γγ 0 β 00 γ 0 β

    ︸ ︷︷ ︸

    strain

    +

    qγLBz

    0 0 i 00 0 0 i−i 0 0 00 −i 0 0

    ︸ ︷︷ ︸

    orbital Zeeman

    +γSBz

    2

    1 0 0 00 −1 0 00 0 1 00 0 0 −1

    ︸ ︷︷ ︸

    spin Zeeman

    , (B1)

    where λ is a spin-orbit coupling constant, γL = µB andγS = 2µB are Landé g-factors of the orbital and spindegrees of freedom (µB the Bohr magneton), q = 0.1 isa Ham reduction factor of the orbital momentum [25,63], and α, β, γ are local strain parameters which canbe different for the ground and excited sates [Sec. IV].

  • 16

    As measuring the exact strain parameters is challenging[Sec. IV] we assume only one non-zero component in thistensor (�zx) in order to simplify our calculations. In thiscase, strain parameters are:

    β = fg(e)�zx, (B2)

    α = γ = 0, (B3)

    where fg(e) = 1.7×106 (3.4×106) GHz/strain [23] for theground (excited) state and the GS splitting is:

    ∆GS = 2√λ2g + β

    2, (B4)

    where λg ≈ 25 GHz is the SO-constant for the groundstate. Next, we solve this Hamiltonian and investigatehow the qubit frequency changes as a function of relativestrain fluctuations (ξ):

    ∆fMW =2 (fg�zx)

    2λgBzqγL(

    (fg�zx)2

    + λ2g

    )3/2 ξ. (B5)The corresponding change in the optical frequency is:

    ∆foptical =

    (fg�zx)2√(fg�zx)

    2+ λ2g

    − (fe�zx)2√

    (fe�zx)2

    + λ2e

    ξ,(B6)

    where λe ≈ 125 GHz is the SO-constant for the excitedstate.

    For SiV 2 [Sec. IV] we measured ∆GS = 140 GHz andfind �zx = 3.8 × 10−5. With ξ = 1% strain fluctuations(corresponding to ∼ 10−7 strain), frequencies change by∆fMW ≈ 4 MHz and ∆foptical ≈ −300 MHz. This quan-titatively agrees with the data presented in [Fig. 4(f)].

    Appendix C: Mitigating spectral diffusion

    In order to couple SiV centers to a quantum network,electronic transitions must be stabilized with respect to

    (a) (b) (c)

    0

    4

    Laser Detuning (GHz)

    Tim

    e (m

    in)

    -1 1

    SiV 1

    -1 1

    SiV 2

    -1 1

    SiV 2 1

    0

    FIG. 12. (a) Spectral diffusion of SiV 2. We observe slowspectral wandering as well as spectral jumps. (b) Applyinga short green repumping pulse before every measurement sig-nificanly speeds up the timescale for spectral diffusion. (c)Spectral diffusion of SiV 1 in nanostructures. Line is stableto below 100 MHz over many minutes. Scale bar indicatesnormalized SiV reflection signal.

    a probe laser. We note that such spectral diffusion isa universal challenge for solid-state quantum systems[45, 64, 65]. In the case of the SiV center, spectral dif-fusion can be seen explicitly in figure 12(a), where theoptical transition frequency can either drift slowly (cen-tral region), or undergo large spectral jumps. As thisdiffusion can be larger than the SiV linewidth, any giveninstance of an experiment could have the probe laser com-pletely detuned from the atomic transition, resulting ina failed experiment.

    There are several possible solutions to mitigate thisspectral diffusion. First, exploiting a high-cooperativityinterface, one can Purcell-broaden the optical linewidth[sec. V] to exceed the spectral diffusion [13]. Second,a high collection efficiency can be used to read out theoptical position faster than the spectral diffusion. Thefrequency can then be probabilistically stabilized by ap-plying a short laser pulse at 520 nm which dramaticallyspeeds up the timescale of spectral diffusion, [22, 66][Fig. 12(b)]. Alternatively this signal could be used to ac-tively stabilize the line using strain-tuning [62, 67]. Fromthe observations in figure 4(f), this technique should mit-igate spectral diffusion of both the optical and spin tran-sitions. Strain tuning also offers the capability to con-trol the DC strain value, which has important effects onqubit properties as discussed previously, and enables tun-ing multiple SiV centers to a common network operationfrequency. As such, this tunability will likely be an im-portant part of future quantum networking technologiesbased on SiV centers.

    The severity of spectral diffusion is different for differ-ent emitters however, and this control is not always nec-essary, especially for proof-of-principle experiments witha small number of emitters. For SiV 1, the main SiV usedin the following sections, and the SiV used in ref [29], wefind almost no spectral diffusion, with optical transitionsstable over many minutes [Fig. 12(c)]. This is an idealconfiguration, as experiments can be performed withoutany need to verify the optical line position.

    Appendix D: Model for SiV decoherence

    The scaling of T2(N) ∝ N2/3 is identical to that foundfor nitrogen-vacancy centers, where it is assumed thatT2 is limited by a fluctuating electron spin bath [50, 51].Motivated by DEER measurements with SiV 2, we fol-low the analysis of ref. [51] to estimate the noise bathobserved by SiV 1.

    The measured coherence decay is modeled by:

    〈Sz〉 = Exp(−∫

    dω S(ω)FN (t, ω)), (D1)

    where S(ω) is the noise power-spectrum of the bath, andFN (t, ω) = 2 sin(ωt/2)(1−sec(ωt/2N))2/ω2 is filter func-tion for a dynamical-decoupling sequence with an evennumber of pulses [51]. We fit sucessive T2 echo curves

  • 17

    to the functional form A + Be−(t/T2)β

    , with A,B beingfree parameters associated with photon count rates, andβ = 3 providing the best fit to the data. This valueof β implies a decoherence bath with a Lorentzian noisepower-spectrum, S(ω, b, τ) = b2τ/π×1/(1+ω2τ2), whereb is a parameter corresponding to the strength of thenoise bath, and τ is a parameter corresponding to thecorrelation time of the noise [50, 51].

    Empirically, no one set of noise parameters faithfullyreproduces the data for all measured echo sequences.Adding a second source of dephasing S̃ = S(ω, b1, τ1) +S(ω, b2, τ2), gives reasonable agreement with the datausing parameters b1 =5 kHz, τ1 =1 µs, b2 =180 kHz,τ2 =1 ms [Fig. 7(d)]. The two drastically different setof noise parameters for each of the sources can help illu-minate the source of noise in our devices.

    As explained in the previous section, one likely can-didate for this decoherence is a bath of free electronsarising from improper surface termination or local dam-age caused during nanofabrication, which are known tohave correlation times in the ∼µs range. The SiV studiedin this analysis is approximately equidistant from threesurfaces: the two nearest holes which define the nanopho-tonic cavity, and the top surface of the nanobeam [sec: II],all of which are approximately 50 nm away. We estimatea density of σsurf = 0.067 spins/nm

    2 using:

    b1 = γSiV〈Bsurf〉 =g2µ2Bµ0

    ~1

    4πΣd2i

    √π

    4σsurf(D2)

    where b1 is the measured strength of the noise bath, g isthe electron gyromagnetic ratio, and di are the distancesto the nearest surfaces. This observation is consistentwith surface spin densities measured using NVs [51].

    The longer correlation time for the second noise termsuggests a different bath, possibly arising from free elec-tron spins inside the bulk diamond. Vacancy clusters,which can persist under annealing even at 1200 C, areknown to posses g = 2 electron spins, and are one pos-sible candidate for this noise bath [68]. Integrating overd in eq. D2, we estimate the density of spins requiredto achieve the measured b2. We estimate ρbulk ∼ 0.53spins per nm3, which corresponds to a doping of 3ppm.Interestingly, this is nearly identical to the local concen-tration of silicon incorporated during implantation (mostof which is not successfully converted into negativelycharged SiV), and could imply implantation-related dam-age as a possible source of these impurities.

    Another possible explanation for this slower bath couldbe coupling to nuclear spins in the environment. The di-amond used in this experiment has a natural abundanceof 13C, a spin-1/2 isotope, in concentrations of approxi-mately 1.1%. Replacing µB → µN in the term for 〈B〉gives an estimated nuclear spin density of ρbulk,N = 0.6%,only a factor of two different than the expected nuclearspin density.

    Appendix E: Concurrence and Fidelity calculations

    1. Spin-photon concurrence and fidelitycalculations

    From correlations in the Z- and X-bases, we estimatea lower bound for the entanglement in our system. Fol-lowing reference [69], we note that the density matrix ofour system conditioned on the detection of one photoncan be described as:

    ρZZ = 1/2

    pe↑ 0 0 00 pe↓ ce↓,l↑ 0

    0 c†e↓,l↑ pl↑ 00 0 0 pl↓

    (E1)where pij are the probabilities of measuring a photon instate i, and the spin in state j. ce↓,l↑ represents entan-glement between pe↑ and pl↓. We set all other coherenceterms to zero, as they represent negligibly small errors inour system (for example, ce↑,e↓ > 0 would imply that theSiV was not initialized properly at the start of the mea-surement). We quantify the degree of entanglement inthe system by its concurrence C, which is 0 for seperablestates, and 1 for a maximally entangled state [70]:

    C = Max(0, λ1/20 −N∑i=1

    λ1/2i ), (E2)

    where λi are the eigenvalues of the matrix ρZZ ·(σy · ρZZ · σ†y

    ), and σy is the standard Pauli matrix act-

    ing on each qubit basis separately (σy = σy,ph ⊗ σy,el).While this can be solved exactly, the resulting equation iscomplicated. Taking only the first-order terms, this canbe simplified to put a lower bound on the concurrence:

    C ≥ 2(|ce↓,l↑| −√pe↑pl↓) (E3)

    We measure p direcly in the Z basis, and estimate |ce↓,l↑|by performing measurements in the X basis. A π/2-rotation on both the photon and spin qubits rotates:

    |e〉 → 1/√

    2(|e〉+ |l〉), |l〉 → 1/√

    2(|e〉 − |l〉)|