Skip to content
BY 4.0 license Open Access Published by De Gruyter March 13, 2024

Strongly coupled spins of silicon-vacancy centers inside a nanodiamond with sub-megahertz linewidth

  • Marco Klotz , Richard Waltrich ORCID logo , Niklas Lettner , Viatcheslav N. Agafonov and Alexander Kubanek EMAIL logo
From the journal Nanophotonics

Abstract

The search for long-lived quantum memories, which can be efficiently interfaced with flying qubits, is longstanding. One possible solution is to use the electron spin of a color center in diamond to mediate interaction between a long-lived nuclear spin and a photon. Realizing this in a nanodiamond furthermore facilitates the integration into photonic devices and enables the realization of hybrid quantum systems with access to quantum memories. Here, we investigated the spin environment of negatively charged silicon-vacancy centers in a nanodiamond and demonstrate strong coupling of its electron spin, while the electron spin’s decoherence rate remained below 1 MHz. We furthermore demonstrate multi-spin coupling with the potential to establish registers of quantum memories in nanodiamonds.

1 Introduction

In the future, quantum-based networks can provide secure communication or distributed quantum computing [1], [2], [3], [4]. One of the remaining challenges is finding a scalable network node, which can process, distribute, and store quantum information, efficiently. Qubits based on solid-state quantum emitters offer advantages in terms of scalability. First small networks, for example, based on negatively charged nitrogen-vacancy centers in diamond (NV) have been realized in a pioneering work [5]. However, the NV is prone to perturbations from external fields and the rate of coherent photons is low [6], [7]. In contrast, group-IV defects like the negatively charged silicon-vacancy center (SiV) are insensitive to external electric fields and show intrinsically identical emitter [8], [9], [10]. Recent results demonstrated coherent control of the SiV spin with coherence times in the ms range when operating at mK temperatures [11]. Increasing the operation temperature is desirable to reduce technical overhead. One potential solution is SiV in nanodiamonds (NDs) with modified electron–phonon interactions [12], which can be integrated into hybrid quantum systems such as photonic crystal cavities [13], [14] or plasmonic waveguides [15]. In this letter, we show the observation of spins in a ND strongly coupled the electron spin of a SiV center. The coupling strength of one of the spins is in good agreement to theoretical modeling of a nearest neighbor 13C nuclear spin [16]. We further show that for a SiV in a ND the main mechanism for decoherence of its spin qubit, which is phonon-mediated dephasing, can already be mitigated at temperatures of around 4 K. The resulting suppressed decoherence rate, access to a local memory, and fast initialization rates lay the foundations for coherent control of an integratable hybrid quantum network node based on SiV s in NDs.

2 Results

The SiV is a point defect in the diamond lattice, where a silicon atom (Si) with an excess electron is situated between two adjacent carbon vacancies (V) as schematically depicted in Figure 1(a). An atomic force microscope scan of the ND containing the investigated SiV revealed an agglomeration of NDs as shown in Figure 1(b). We estimate the size range of individual NDs to be between 50 nm and 250 nm. The agglomeration is small enough to be integrated into a cavity system [17]. For individual NDs within the agglomeration, a modified phonon density of states (PDOS) at low strain is possible [12]. The electronic level scheme of the SiV consists of four spin-degenerate orbital states, two of which form the ground-state (GS) and excited-state (ES), respectively. As a consequence, four optically active transitions arise, which we label as A, B, C, and D. For the remainder of the text, we only use transition C, for which the spin levels are depicted in Figure 1(c). The spin degeneracy of the GS and ES levels can be lifted by applying a magnetic field, giving access to an electron spin qubit, e.g., the one labeled by |↓⟩ and |↑⟩ [18], [19]. When using such a spin-qubit at liquid helium temperature, its coherence time is mainly limited through phonon-induced dephasing. The latter can be mitigated by either cooling the system to mK temperatures [11], changing the PDOS [12], or increasing the GS splitting, which suppresses phonon absorption [20], [21]. The use of NDs is an appealing choice as a host for the SiV, since they can combine two of the mentioned effects to increase spin-coherence times at temperatures around 4 K. The reduced size can modify the PDOS and commonly present strain in NDs results in an increased GS splitting. We, therefore, investigate spectrally shifted SiV where high strain can be expected [20] and studied them for their optical and spin-coherence properties.

Figure 1: 
SiV center in a ND and partial electronic structure of the SiV. (a) Illustration of a SiV− with a coupled 13C nuclear spin inside a ND. The implied shear represents the presence of strain in the host crystal. (b) AFM scan of the nanodiamond agglomeration containing the SiV−s under consideration. (c) Partial level scheme for transition C of the SiV−. The four levels originating from the Zeeman effect are labeled as |↓⟩, |↑⟩, |↓′⟩, and |↑′⟩. The corresponding hyperfine levels are indicated with an additional ⇓ or ⇑. Λ systems connecting the correspondingly involved electron and nuclear spins are differently colored. Single- and two-photon detunings are labeled with Δ and δ, respectively. Other transitions arising from A, B, and D are not shown for simplicity.
Figure 1:

SiV center in a ND and partial electronic structure of the SiV. (a) Illustration of a SiV with a coupled 13C nuclear spin inside a ND. The implied shear represents the presence of strain in the host crystal. (b) AFM scan of the nanodiamond agglomeration containing the SiVs under consideration. (c) Partial level scheme for transition C of the SiV. The four levels originating from the Zeeman effect are labeled as |↓⟩, |↑⟩, |↓′⟩, and |↑′⟩. The corresponding hyperfine levels are indicated with an additional ⇓ or ⇑. Λ systems connecting the correspondingly involved electron and nuclear spins are differently colored. Single- and two-photon detunings are labeled with Δ and δ, respectively. Other transitions arising from A, B, and D are not shown for simplicity.

The NDs were coated onto a sapphire substrate with good thermal conductivity. The sample was then cooled to liquid helium temperatures in a continuous flow-cryostat and investigated using a home-built confocal microscope. Four permanent magnets in a Hallbach-configuration designed for an in-plane field strength of around 400 mT were used to lift the spin degeneracy. Individual transitions of the SiV were addressed by photo-luminescence excitation spectroscopy (PLE). After finding a suitable SiV, the optical linewidths of the two spin-preserving transitions, labeled as C2 and C3, were investigated by PLE with varying power (P), as shown in Figure 2(a). The frequency splitting between C2 and C3 390 ± 4 MHz was determined by a double-Lorentzian fit. Figure 2(b) shows the fitted power-dependent linewidth Γ of C2 and C3. We extrapolated the linewidth to zero power using Γ ( P ) = Γ 0 s + 1 , where Γ0 is the linewidth at zero power and s = P/P sat with the saturation power P sat. The fit resulted in Γ 0 C 2 = 125 ± 9 MHz and Γ 0 C 3 = 101 ± 18 MHz , respectively. The saturation powers turned out as P sat C 2 = 5.8 ± 1.2 nW and P sat C 3 = 3.4 ± 1.6 nW , respectively. Firstly, Γ0 is close to Fourier limits for commonly found lifetimes inside NDs, suggesting excellent optical quality. Secondly, the narrow linewidths compared to 390 MHz allow for a strong drive without significant cross-talk between C2 and C3.

Figure 2: 
Optical properties of a SiV-center. (a) The data points show power-dependent PLE scans of the spin-preserving transitions C2 and C3 with the frequency relative to 




ν


0




C


2






${\nu }_{0}^{{C}_{2}}$



 of transition C2. The solid lines are double-Lorentzian fits to the data and reveal a splitting of 390 ± 4 MHz for the lowest excitation power P. (b) Fitted linewidth as a function of P with squares and triangles representing C2 and C3, respectively. The dotted and solid lines are the respective square-root fit. (c) Normalized fluorescence during spin initialization for increasing excitation powers with the solid lines being an exponential fit. For the highest power (lower panel), the initialization time is 387 ± 10 ns with a fidelity of 0.806 ± 0.004. (d) The extracted initialization rates Γinit as a function of P.
Figure 2:

Optical properties of a SiV-center. (a) The data points show power-dependent PLE scans of the spin-preserving transitions C2 and C3 with the frequency relative to ν 0 C 2 of transition C2. The solid lines are double-Lorentzian fits to the data and reveal a splitting of 390 ± 4 MHz for the lowest excitation power P. (b) Fitted linewidth as a function of P with squares and triangles representing C2 and C3, respectively. The dotted and solid lines are the respective square-root fit. (c) Normalized fluorescence during spin initialization for increasing excitation powers with the solid lines being an exponential fit. For the highest power (lower panel), the initialization time is 387 ± 10 ns with a fidelity of 0.806 ± 0.004. (d) The extracted initialization rates Γinit as a function of P.

We proceeded by measuring the initialization rate of the spin for each excitation power. To this end, we pumped the population out of a thermal maximally mixed equilibrium state with a resonant pulse of varying power. The resulting fluorescence, shown in Figure 2(c), was then fitted with an exponential decay. From the fit, we evaluated the initialization fidelity by comparing the maximum fluorescence at the beginning of the pulse with the fluorescence at the end of the pulse, corresponding to the steady state population under resonant drive. In addition, the corresponding initialization rates Γinit are shown in Figure 2(d). For the highest power, we measured an initialization time of 387 ± 10 ns and a fidelity of 0.806 ± 0.004. From the power dependence of Γinit, we determined the spin-branching ratio η = 22.7 ± 1.2 by fitting Γ init = 1 η s ( 1 + s ) Γ 0 C 2 2 , where Γ 0 C 2 and P sat C 2 are taken from the previously fitted PLE measurements [22].

Extending the measurement procedure of Figure 2(c) to multiple consecutive resonant pulses with increasing temporal spacing probes the exponentially recovering spin population T 1. Fitting an exponential to the peak intensities in each fluorescence pulse resulted in T 1 = 2.0 ± 0.7μs, which ultimately limits spin coherence. In addition, T 1 gives insight into the orientation of the SiV symmetry axis with respect to the magnetic field, where a misalignment leads to spin mixing and thus shorter relaxation times [23]. From the relatively short T 1 together with the spin-branching η, we estimate only a moderate alignment. A tunable magnet system or nanomanipulation [24] may improve the alignment and thus increase the maximum achievable coherence time and initialization fidelity in future experiments [23].

To probe the spin-coherence, we used coherent population trapping (CPT). Here, the laser resonantly (Δ = 0, see Figure 1(c)) drove a spin-flipping transition. Simultaneously, an electro-optical modulator (EOM) generated sidebands from which one was swept over the corresponding spin-preserving transition. If the Raman condition (δ = 0) is fulfilled, the system is pumped into a dark state quenching the fluorescence signal. For the studied SiV, two dips with a frequency splitting of A = 38.47 ± 0.12 MHz were present, as shown in Figure 3(a). The splitting A is composed of two terms, the parallel coupling term A and the perpendicular coupling term A . The observed splitting is in close agreement with a theoretically predicted strongly hyperfine-coupled next-nearest neighbor 13C nuclear spin with a coupling strength of A = 37 MHz [16].

Figure 3: 
Coupling and coherence properties of a SiV-center. (a) CPT measurement of a SiV−-center showing two dips, indicating a nearby 13C nuclear spin. The solid line is a triple-Lorentzian fit to the data points. The resulting Zeeman splitting is around 11 GHz and the splitting between the two dips is A = 38.47 ± 0.12 MHz. (b) Measurement of the CPT width of the dip highlighted by the gray area in (a) for varying power. (c) Power-dependent CPT width 




Γ


2


*




${{\Gamma }}_{2}^{\ast }$



. A linear fit yields a rate of 




Γ


2


*


=
320

±

120


kHz


${{\Gamma }}_{2}^{\ast }=320\,\pm \,120\enspace \,\text{kHz}$



 at zero power.
Figure 3:

Coupling and coherence properties of a SiV-center. (a) CPT measurement of a SiV-center showing two dips, indicating a nearby 13C nuclear spin. The solid line is a triple-Lorentzian fit to the data points. The resulting Zeeman splitting is around 11 GHz and the splitting between the two dips is A = 38.47 ± 0.12 MHz. (b) Measurement of the CPT width of the dip highlighted by the gray area in (a) for varying power. (c) Power-dependent CPT width Γ 2 * . A linear fit yields a rate of Γ 2 * = 320 ± 120 kHz at zero power.

Furthermore, since the dip’s linewidth gives insight on the spin’s dephasing rate, we performed a power-dependent measurement, which is shown in Figure 3(b). Fitting the data with a Lorentzian and extracting the linewidth Γ 2 * allowed us to linearly extrapolate the dephasing rate to zero laser power to suppress power-induced broadening. As a result, we obtained a zero-power linewidth of Γ 2 * = 320 ± 120 kHz . This is approximately a ten-fold reduced decoherence rate compared to previously reported measurements in bulk diamond at similar temperatures with Γ 2 * 4.5 MHz [25], [26], [27], and a minor improvement to measurements in a strain-engineered nano-beam, which reported Γ 2 * = 640 kHz [20], [21]. Because the minimal size of individual NDs is 50 nm, we do not expect a significantly changed PDOS in a high strain environment. Instead, we attribute the improved coherence time mainly to the suppressed phonon absorption caused by the increased GS splitting. Combining the latter improvements of spin coherence in high strain environments with the present spin transition frequency of roughly 11 GHz and the spin-cycling transitions’ splitting of 390 MHz suggests a comparably highly strained SiV-center with a GS splitting on the order of 500 GHz [20].

We further investigated the spin environment in the same agglomeration of NDs by performing CPT measurements on different SiV-centers. The measurement revealed coupling of multiple spins, as shown in Figure 4. For example, the SiV in Figure 4(a) displayed two dips split by A = 41.5 ± 0.7 MHz. Measurements with reduced power resolved two more dips, split by A′ = 5.47 ± 0.17 MHz as shown in the inset. They exhibited linewidths of 3.2 ± 0.4 MHz and 4.3 ± 0.5 MHz, for the left and right dip, respectively. The distinct linewidths suggest coupling to two other surrounding SiV-centers’ electron spins with coherence properties commonly found with SiV in low to moderate strain bulk diamond [25], [26], [27]. Assuming a dipolar electron–electron coupling would correspond to a distance on the order of 10 nm, reasonable for the size of the ND and density of SiV-centers under study. Another SiV, displayed in Figure 4(b), exhibited two individual dips at low optical power with linewidths of 2.0 ± 0.3 MHz and 2.6 ± 0.5 MHz, for the left and right, respectively. In contrast to the previous SiVs’ multi-dip structure with distinct linewidths, the fact that the present SiV has closely matching and relatively narrow linewidths is indicative of a coupled nearby 13C nuclear spin with a coupling strength of 5.20 ± 0.14 MHz [21]. In this case, the distance between the two coupled spins is on the order of 1 Å [21].

3 Conclusions

The presented results open up new possibilities to utilize the electron and nuclear spin environment of SiV-centers in NDs as an elementary unit of diamond-based qubits with integrability into photonic platforms, like photonic crystal or open microcavities [14], [17], [28], [29]. Depending on the size of the ND, evanescent coupling to other systems such as plasmonic waveguides is also possible, where high Purcell factors can be expected [15]. In addition, phononic [30] or optomechanical [31] platforms could be used to further reduce phonon-induced dephasing via a modified PDOS while maintaining or enhancing the photon collection efficiency. The thereby formed hybrid quantum system enables efficient mapping of the quantum state of the electron spin to flying photonic qubits as well as coupling to local memory units consisting of nearby nuclear spins. Our work also suggests that the access to small spin registers is feasible for SiV-centers in NDs.

Figure 4: 
CPT measurements of two more SiV− with low and high optical power, indicated in yellow and blue, respectively. (a) Shows two dips, split by A = 41.5 ± 0.7 MHz. Low-power measurements reveal a coupled system with an additional splitting of A′ = 5.47 ± 0.17 MHz. (b) Measurements show a double dip, split by 5.20 ± 0.14 MHz, indicating a weakly coupled C13 nuclear spin.
Figure 4:

CPT measurements of two more SiV with low and high optical power, indicated in yellow and blue, respectively. (a) Shows two dips, split by A = 41.5 ± 0.7 MHz. Low-power measurements reveal a coupled system with an additional splitting of A′ = 5.47 ± 0.17 MHz. (b) Measurements show a double dip, split by 5.20 ± 0.14 MHz, indicating a weakly coupled C13 nuclear spin.

Increased strain, e.g., caused by hydrostatic pressure might further improve coherence times [32]. Relying on intrinsic, nonuniform natural strain challenges the scalability of spectrally matched emitters. However, either postselection of NDs, strain tuning [20], or electric field-induced fine-tuning of the emission wavelength [33] relaxes spectral constraints. Our results are also applicable to other group IV defects in diamond with intrinsically higher GS splitting, such as the germanium vacancy (GeV) and tin vacancy (SnV). For example, highly strained GeV centers in NDs with a GS splitting of 870 GHz have been reported in Ref. [34]. The dephasing rates bring coherent control of different spin qubits in NDs, either by means of direct microwave drive [35] or all-optical control using a Raman-type lambda scheme [22], into reach. Additionally, temperatures around 4 K within a flow-cryostat become sufficient. For the system parameters in our measurements, an all-optical Rabi driving strength in the order of several MHz can be expected, comparable to driving rates achieved with a microwave [35]. Additionally, the high sensitivity enables to detect close-by electron or nuclear spins. Furthermore, with coherent electron spin control, direct [36] or indirect [37] control of a nuclear spin becomes possible. Mapping information of the electron spin to a strongly coupled long-lived nuclear spin establishes SiV-centers in NDs as a viable candidate for an interchangeable hybrid quantum memory.


Corresponding author: Alexander Kubanek, Institute for Quantum Optics, Ulm University, 89081 Ulm, Germany, E-mail:

Marco Klotz and Richard Waltrich contributed equally to this work.


Award Identifier / Grant number: 16KISQ006

Award Identifier / Grant number: 16KISQO43K

Acknowledgments

The authors thank V.A. Davydov for synthesis and processing of the nanodiamond material.

  1. Research funding: The project was funded by the Baden-Württemberg Stiftung in Project Internationale Spitzenforschung. A.K. acknowledges support of the BMBF/VDI in Project HybridQToken (16KISQO43K) and QR.X (16KISQ006) and Spinning (13N16215). The authors acknowledge funding by the European Union and the DFG within the Quanteraproject SensExtreme.

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: Authors state no conflicts of interest.

  4. Informed consent: Informed consent was obtained from all individuals included in this study.

  5. Ethical approval: The conducted research is not related to either human or animals use.

  6. Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] H. J. Kimble, “The quantum internet,” Nature, vol. 453, no. 7198, p. 1023, 2008. https://doi.org/10.1038/nature07127.Search in Google Scholar PubMed

[2] S. Wehner, D. Elkouss, and R. Hanson, “Quantum internet: a vision for the road ahead,” Science, vol. 362, no. 6412, p. 6412, 2018. https://doi.org/10.1126/science.aam9288.Search in Google Scholar PubMed

[3] H.-J. Briegel, W. Dür, J. I. Cirac, and P. Zoller, “Quantum repeaters: the role of imperfect local operations in quantum communication,” Phys. Rev. Lett., vol. 81, no. 26, p. 5932, 1998. https://doi.org/10.1103/physrevlett.81.5932.Search in Google Scholar

[4] L. Childress, J. M. Taylor, A. S. Sørensen, and M. D. Lukin, “Fault-tolerant quantum communication based on solid-state photon emitters,” Phys. Rev. Lett., vol. 96, no. 7, p. 070504, 2006. https://doi.org/10.1103/physrevlett.96.070504.Search in Google Scholar

[5] M. Pompili, et al.., “Realization of a multinode quantum network of remote solid-state qubits,” Science, vol. 372, no. 6539, p. 259, 2021. https://doi.org/10.1126/science.abg1919.Search in Google Scholar PubMed

[6] A. Faraon, C. Santori, Z. Huang, V. M. Acosta, and R. G. Beausoleil, “Coupling of nitrogen-vacancy centers to photonic crystal cavities in monocrystalline diamond,” Phys. Rev. Lett., vol. 109, no. 3, p. 033604, 2012. https://doi.org/10.1103/physrevlett.109.033604.Search in Google Scholar

[7] Y. Chu, et al.., “Coherent optical transitions in implanted nitrogen vacancy centers,” Nano Lett., vol. 14, no. 4, p. 1982, 2014. https://doi.org/10.1021/nl404836p.Search in Google Scholar PubMed

[8] L. J. Rogers, et al.., “Multiple intrinsically identical single-photon emitters in the solid state,” Nat. Commun., vol. 5, no. 1, p. 4739, 2014. https://doi.org/10.1038/ncomms5739.Search in Google Scholar PubMed

[9] A. Sipahigil, et al.., “Indistinguishable photons from separated silicon-vacancy centers in diamond,” Phys. Rev. Lett., vol. 113, no. 11, p. 113602, 2014. https://doi.org/10.1103/physrevlett.113.113602.Search in Google Scholar

[10] R. Waltrich, M. Klotz, V. N. Agafonov, and A. Kubanek, “Two-photon interference from silicon-vacancy centers in remote nanodiamonds,” Nanophotonics, vol. 12, no. 15, p. 3663, 2023. https://doi.org/10.1515/nanoph-2023-0379.Search in Google Scholar

[11] D. D. Sukachev, et al.., “Silicon-vacancy spin qubit in diamond: a quantum memory exceeding 10 ms with single-shot state readout,” Phys. Rev. Lett., vol. 119, no. 22, p. 223602, 2017. https://doi.org/10.1103/physrevlett.119.223602.Search in Google Scholar

[12] M. Klotz, et al.., “Prolonged orbital relaxation by locally modified phonon density of states for the siv− center in nanodiamonds,” Phys. Rev. Lett., vol. 128, no. 15, p. 153602, 2022. https://doi.org/10.1103/physrevlett.128.153602.Search in Google Scholar

[13] K. G. Fehler, et al.., “Purcell enhanced emission from individual siv center in nanodiamonds coupled to a si3n4 based photonic crystal cavity,” Nanophotonics, vol. 9, no. 11, p. 3655, 2020. https://doi.org/10.1515/nanoph-2020-0257.Search in Google Scholar

[14] K. G. Fehler, et al.., “Hybrid quantum photonics based on artificial atoms placed inside one hole of a photonic crystal cavity,” ACS Photonics, vol. 8, no. 9, p. 2635, 2021. https://doi.org/10.1021/acsphotonics.1c00530.Search in Google Scholar

[15] H. Siampour, S. Kumar, V. A. Davydov, L. F. Kulikova, V. N. Agafonov, and S. I. Bozhevolnyi, “On-chip excitation of single germanium vacancies in nanodiamonds embedded in plasmonic waveguides,” Light: Sci. Appl., vol. 7, no. 1, p. 61, 2018. https://doi.org/10.1038/s41377-018-0062-5.Search in Google Scholar PubMed PubMed Central

[16] A. P. Nizovtsev, et al.., “Spatial and hyperfine characteristics of SiV– and SiV0 color centers in diamond DFT simulation,” Semiconductors, vol. 54, no. 12, p. 1685, 2020. https://doi.org/10.1134/s1063782620120271.Search in Google Scholar

[17] G. Bayer, et al.., “Optical driving, spin initialization and readout of single siv- centers in a fabry-perot resonator,” Commun. Phys., vol. 6, no. 1, p. 300, 2023. https://doi.org/10.1038/s42005-023-01422-7.Search in Google Scholar

[18] C. Hepp, et al.., “Electronic structure of the silicon vacancy color center in diamond,” Phys. Rev. Lett., vol. 112, no. 3, p. 036405, 2014. https://doi.org/10.1103/physrevlett.112.036405.Search in Google Scholar

[19] L. J. Rogers, et al.., “Electronic structure of the negatively charged silicon-vacancy center in diamond,” Phys. Rev. B, vol. 89, no. 23, p. 235101, 2014. https://doi.org/10.1103/physrevb.89.235101.Search in Google Scholar

[20] S. Meesala, et al.., “Strain engineering of the silicon-vacancy center in diamond,” Phys. Rev. B, vol. 97, no. 20, p. 205444, 2018. https://doi.org/10.1103/physrevb.97.205444.Search in Google Scholar

[21] Y.-I. Sohn, et al.., “Controlling the coherence of a diamond spin qubit through its strain environment,” Nat. Commun., vol. 9, no. 1, p. 2012, 2018. https://doi.org/10.1038/s41467-018-04340-3.Search in Google Scholar PubMed PubMed Central

[22] R. Debroux, et al.., “Quantum control of the tin-vacancy spin qubit in diamond,” Phys. Rev. X, vol. 11, no. 4, p. 041041, 2021. https://doi.org/10.1103/physrevx.11.041041.Search in Google Scholar

[23] E. I. Rosenthal, et al.., “Microwave spin control of a tin-vacancy qubit in diamond,” Phys. Rev. X, vol. 13, no. 3, p. 031022, 2023. https://doi.org/10.1103/physrevx.13.031022.Search in Google Scholar

[24] N. Lettner, et al.., “Controlling all degrees of freedom of the optical coupling in hybrid quantum photonics,” ACS Photonics, vol. 11, no. 2, 2023. https://doi.org/10.1021/acsphotonics.3c01559.Search in Google Scholar

[25] L. J. Rogers, et al.., “All-optical initialization, readout, and coherent preparation of single silicon-vacancy spins in diamond,” Phys. Rev. Lett., vol. 113, no. 26, p. 263602, 2014. https://doi.org/10.1103/physrevlett.113.263602.Search in Google Scholar PubMed

[26] J. N. Becker, et al.., “All-optical control of the silicon-vacancy spin in diamond at millikelvin temperatures,” Phys. Rev. Lett., vol. 120, no. 5, p. 053603, 2018. https://doi.org/10.1103/physrevlett.120.053603.Search in Google Scholar PubMed

[27] B. Pingault, et al.., “All-optical formation of coherent dark states of silicon-vacancy spins in diamond,” Phys. Rev. Lett., vol. 113, no. 26, p. 263601, 2014. https://doi.org/10.1103/physrevlett.113.263601.Search in Google Scholar

[28] L. Antoniuk, et al.., “All-optical spin initialization via a cavity broadened optical transition in on-chip hybrid quantum photonics,” 2023, arXiv:2308.15544 [quant-ph].Search in Google Scholar

[29] C. T. Nguyen, et al.., “An integrated nanophotonic quantum register based on silicon-vacancy spins in diamond,” Phys. Rev. B, vol. 100, no. 31, p. 165428, 2019. https://doi.org/10.1103/physrevb.100.165428.Search in Google Scholar

[30] K. Kuruma, et al.., “Engineering phonon-qubit interactions using phononic crystals,” arXiv preprint arXiv:2310.06236, 2023.Search in Google Scholar

[31] M. J. Burek, et al.., “Diamond optomechanical crystals,” Optica, vol. 3, no. 12, p. 1404, 2016. https://doi.org/10.1364/optica.3.001404.Search in Google Scholar

[32] B. Vindolet, et al.., “Optical properties of siv and gev color centers in nanodiamonds under hydrostatic pressures up to 180 gpa,” Phys. Rev. B, vol. 106, no. 21, p. 214109, 2022. https://doi.org/10.1103/physrevb.106.214109.Search in Google Scholar

[33] S. Aghaeimeibodi, D. Riedel, A. E. Rugar, C. Dory, and J. Vučković, “Electrical tuning of tin-vacancy centers in diamond,” Phys. Rev. Appl., vol. 15, no. 6, p. 064010, 2021. https://doi.org/10.1103/physrevapplied.15.064010.Search in Google Scholar

[34] H. Siampour, et al.., “Ultrabright single-photon emission from germanium-vacancy zero-phonon lines: deterministic emitter-waveguide interfacing at plasmonic hot spots,” Nanophotonics, vol. 9, p. 953, 2020, https://doi.org/10.1515/nanoph-2020-0036.Search in Google Scholar

[35] B. Pingault, et al.., “Coherent control of the silicon-vacancy spin in diamond,” Nat. Commun., vol. 8, no. 1, p. 15579, 2017. https://doi.org/10.1038/ncomms15579.Search in Google Scholar PubMed PubMed Central

[36] P.-J. Stas, et al.., “Robust multi-qubit quantum network node with integrated error detection,” Science, vol. 378, no. 6619, p. 557, 2022. https://doi.org/10.1126/science.add9771.Search in Google Scholar PubMed

[37] H. H. Vallabhapurapu, C. Adambukulam, A. Saraiva, and A. Laucht, “Indirect control of the 29siv− nuclear spin in diamond,” Phys. Rev. B, vol. 105, no. 20, p. 205435, 2022. https://doi.org/10.1103/physrevb.105.205435.Search in Google Scholar

Received: 2023-12-18
Accepted: 2024-02-28
Published Online: 2024-03-13

© 2024 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 1.6.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2023-0927/html
Scroll to top button