Abstract
Solid-state quantum emitters with manipulable spin-qubits are promising platforms for quantum communication applications. Although such light-matter interfaces could be realized in many systems only a few allow for light emission in the telecom bands necessary for long-distance quantum networks. Here, we propose and implement an optically active solid-state spin-qubit based on a hole confined in a single InAs/GaAs quantum dot grown on an InGaAs metamorphic buffer layer emitting photons in the C-band. We lift the hole spin-degeneracy using an external magnetic field and demonstrate hole injection, initialization, read-out and complete coherent control using picosecond optical pulses. These results showcase a solid-state spin-qubit platform compatible with preexisting optical fiber networks.
Subject terms: Quantum dots, Quantum optics, Quantum information, Single photons and quantum effects
Long-distance quantum communication relies on interfacing qubits with telecom-band photons. Here the authors implement a solid-state spin-qubit based on a hole confined in a semiconductor quantum dot emitting photons in the telecom C-band.
Introduction
Long-distance quantum network technologies require reliable interfaces between stationary qubits and photons with frequencies in the telecom bands1,2. In this regard, spin degrees of freedom in solid-state quantum emitters are of particular interest due to the potential of realizing quantum entanglement between a confined spin and a propagating photon3–6. Spin-based photonics have been to date realized using a wide range of material systems. Each platform varies in the terms of electronic structure, emission wavelength, optical performance, spin-coherence, and integration capabilities with photonic devices7,8. Semiconductor quantum dots (QDs) for instance, show excellent optical properties9,10, achieve spin-coherence times on the level of tens of microseconds11,12, and allow for efficient generation of spin–photon4–6,13 and spin–spin14,15 entanglement. Another well-established spin-system, vacancy centers in diamond, while suffering intrinsically low emission efficiency into the zero-phonon-line, show very long coherence times on the level of single seconds16, and allowed for demonstration of spin–spin entanglement on the record distance of 1.3 km17. While diamond is still a rather challenging material for micro-processing, platforms based on defects in SiC show similar optical-performance with milliseconds spin-coherence times18,19, and stand out by maturity of SiC and processing methods. Another promising class of spin-active emitters are rare-earth-ion doped crystals embedded in cavities20–22. Such systems might in principle achieve over seconds spin-coherence times23, but require the usage of high-finesse cavities to enhance the intrinsically slow emission rates.
Among those many types of solid-state platforms, only a few systems demonstrated capability of directly interfacing spins with telecom-wavelength photons. Recently, spin-control has been demonstrated at 1220 nm wavelength in nitrogen-vacancy centers in SiC24, at 1330 nm using ensemble of radiation T-damage centers in 28Si25, and at 1550 nm in Er3+ in Y2SiO521,22. In particular, the latter system operates in the lowest-loss telecom C-band centered at 1550 nm, promising a clear advantage for long-distance quantum communication applications2. Emitters with transitions incompatible with fiber networks, such as defects in diamond, could potentially take advantage of quantum frequency conversion into telecom wavelengths4,26. While such an approach offers flexibility in terms of frequency matching of multiple emitters27, it is intrinsically limited by the efficiency of the frequency conversion process and requires a considerable technological overhead.
Alternative systems that could potentially allow interfacing spins with telecom-wavelength photons are InAs QDs relying on InP substrate28–30 or strain-relaxed layers in GaAs31–33. Spectroscopic studies of such systems identified already charged exciton complexes with intrinsic optical transitions in the C-band offering access to isolated electron and hole states29,30,33,34.
In this work, we introduce a spin-qubit system based on a single hole confined in a InAs self-assembled QD based on the strain-relaxed GaAs emitting radiation directly in the telecom C-band. We demonstrate the full set of spin-control operations consisting of spin injection, initialization, read-out and coherent rotation. We perform all protocols at 76 MHz repetition rate, where the spin-rotation is achieved on the picosecond time-scale using detuned optical pulses. This demonstration provides a clear pathway toward direct spin–photon entanglement in the telecom C-band for applications in long-distance quantum communication. Thanks to the QDs’ intrinsically short radiative lifetimes of around 1 ns, our system provides a direct alternative to Er3+ ions or platforms relying on frequency conversion.
Results
Spin system under investigation
The physical system used in our experiment is presented in Fig. 1a. It consists of InAs QDs grown on an InGaAs metamorphic buffer (MMB) layer, reducing the lattice mismatch between the QDs and the underneath material35. This results in an emission wavelength in the telecom C-band. In addition, the QDs and the MMB are placed on a near lattice-matched AlAs/GaAs distributed Bragg reflector (DBR) that significantly enhances the photon extraction efficiency (see Methods). For optical experiments, we cool-down the sample to 1.6 K temperature and focus on a single QD by using a confocal microscope. Under nonresonant optical excitation, sharp emission lines are identified in photoluminescence spectra stemming from radiative recombination of various excitonic complexes. In particular, positively charged trions formed by a single electron and two holes are identified as dominant transitions in agreement with previous theoretical and experimental studies36. Consequently, in this work, we focus on an isolated emitter with an optical transition between a positively charged exciton and a ground hole state (see Supplementary Note 2). The benefit of using the hole state over the electron is hole’s smaller hyperfine interaction strength with nuclear spins due to the hole’s p-like symmetry37–43 and thus potentially better hole coherence.
Under a magnetic field oriented in Voigt geometry (perpendicular to the growth direction/optical axis), two Λ-systems are formed from an initial two-level system, where each optically active trion state / is coupled to each of two ground hole spin states and . The optical selection rules are indicated in Fig. 1b. In this work, we define a qubit as a quantum-dot hole spin-state , which can be schematically represented on the Bloch sphere, where θ and ϕ are spherical co-ordinates as shown in Fig. 1c. The north and south pole represent the pure and states. Access to any arbitrary Bloch vector on the sphere requires the ability of spin rotation along two perpendicular axes to control both θ and ϕ angle.
In Fig. 1d polarization dependence of the preselected trion photoluminescence at 3 T magnetic field in Voigt configuration is shown. The emission spectrum shows a four-fold splitting, with polarized transitions consistent with selection rules shown in Fig. 1b. By fitting the emission peaks at vertical (V) and horizontal (H) polarization (as defined in Fig. 1) we recover the trion and hole splitting of 7 ± 1 and 115 ± 1 GHz, respectively. Systematic investigation of the peaks in the function of the magnetic field reveal the ground spin-state and trion spin-state g-factors of 2.7 and 0.16, respectively (see Supplementary Note 3).
Optical charge injection
Within the current study, we utilize a nominally undoped QD sample, thus most of the QDs are uncharged. Interestingly, trion transitions are visible under nonresonant pumping, suggesting that photo-excited electrons and holes relax with different rates predominantly forming positive trion states. This, in turn, could be utilized to load the QD with a single charge, since directly after trion recombination the QD is left in the ground hole state. Moreover, since trions can recombine with equal probability into both spin states, such a charge injection scheme randomizes the initial spin-configuration serving additionally as an active-spin-reset.
To experimentally demonstrate the optical charge injection, we perform a two-color pulse experiment as schematically described in Fig. 2a. First, we excite a QD with a 2 ps width nonresonant pulse at 850 nm to form a trion (1–2). Upon radiative recombination (3), the QD is left with single hole in random or spin-configuration (4). After 4.5 ns we send another H-polarized pulse with 100 ps width and central wavelength of 1558 nm, which is resonant with the - transition (5). Taken, that a charge was indeed loaded into the QD, the second pulse should excite the system into the trion state (6), which will then recombine radiatively (7). By spectrally filtering the diagonal V-polarized transition - and performing a time-resolved measurement, we can trace the QD emission dynamics following both pulses, as shown in Fig. 2c. Emission decay after the first pulse exhibits a mono-exponential decay with a time constant of 1.4 ns, as typically observed for such type of QDs35. Emission after the second pulse exhibits a similar mono-exponential decay, however, with a slight amplitude oscillation with a period of 147 ± 1 ps as indicated in the Fig. 2c inset. This oscillations period corresponds to a 6.80 ± 0.05 GHz frequency which is in good agreement with trion state splitting of 7 ± 1 GHz and is a fingerprint of coherent excitation. Interestingly, when the sample is excited with the resonant pulse only, emission from the QD is not observed (black curve). This allows recording the reference background time-trace related to the scattered laser. By integration of the counts corresponding to the first or second decay on the graph (blued area), we can plot the time-integrated intensity of the QD emission after each pulse. In Fig. 2d the emission intensity after each excitation pulse is plotted as a function of the first (charge injection) pulse power. Clear intensity increase and saturation at around 1 μW are observed for both decays. In Fig. 2e we plot the emission after the second (resonant) pulse as a function of the square root of the second pulse power. In this case, clear oscillatory behavior is observed related to coherent Rabi rotation of trion state population. We can observe rotation up to 5π-pulse area, which is clear evidence of coherent trion control.
Spin initialization
To initialize the spin into the desired / state we use optical pumping44,45. For that a continuous wave (cw) laser at H polarization is tuned into resonance with the transition -/- (brown arrows in Fig. 3a, b inset). The fluorescence is monitored at the corresponding V-polarized transitions -/-. A significant increase of the emission intensity is observed when the resonance with the given transition is established as shown in Fig. 3a, b. After a charge is injected into the QD by the nonresonant pulse, the cw pumping laser re-excites the state into , until the event of V-polarized emission, which will initialize the system into . Analogously, by optical pumping of the transition -, the system can be initialized in the state. Since the probability of V and H-polarized fluorescence is equal, the initialization process will require few emission cycles to reach high yield. Based on our simulations, we expect that within 4.5 ns (6 ns) initialization time we can reach 96% (99%) spin-preparation precision (see Supplementary Note 4).
To read the spin state in the QD we send a subsequent H-polarized pulse resonant with -, which probes the state population (pulse sequence shown in Fig. 3c). In the case of the initialization, around two-fold enhancement of the emission intensity can be observed, while in the case of initialization clear suppression is visible as shown in Fig. 3d. In the case of pumping a re-excitation process is visible directly in the time-trace as a prolongation of the emission decay after the nonresonant pulse. In Fig. 3e integrated counts corresponding to the read-out pulse are plotted as a function of the spin-pumping laser power in case of the and initialization, respectively. For the pumping, the emission intensity reaches double the counts observable in case of no initialization, which is related to the increase of the state preparation probability from 50% to almost 100%. In analogy, in the case of pumping, significant suppression of the signal is observed. For spin-pumping powers above 10 μW, a spin-preparation precision of 94.5 ± 0.5% is achieved for 4.5 ns optical pumping time. This could be further improved by increasing the delay between charge injection and spin-read-out pulses.
Coherent control of the hole spin
For coherent rotation of the hole spin-state to any desired point on the Bloch sphere we employ detuned circularly polarized laser pulses46. If the laser detuning Δ is much larger than the splitting between the hole ground states δh, the net effect on the double Λ system is such that the population of the spin state should oscillate with an effective Rabi frequency Ωeff, without physically exciting the system into the trion state. By keeping the pulse length constant and by controlling its power, it is possible to coherently rotate the spin state between and . The rotation angle θ will be in such case proportional to the square root of the detuned laser power and inversely proportional to the detuning46.
For spin-control experiments we apply a magnetic field of 1.5 T, decreasing the hole spin splitting to 57 GHz and increasing the Larmor precession period to around 17.5 ps. First, we send a nonresonant 850 nm charge injection pulse and initialize the system into the state by optical pumping. Then, after around 6 ns we send a circularly polarized ~400 GHz red detuned spin-rotation pulse with a length of 4 ps. To read if the spin was rotated from to state, a cw spin-pumping laser excites the system into the trion state. Following radiative recombination of a V or H-polarized photon signalizes the spin-rotation. The respective pulse sequence used is shown in Fig. 4b. By monitoring the emission evolution of the V-polarized - transition we can detect the spin-rotation events as shown in the time-trace in Fig. 4c. We note here that pumping powers required for spin rotation are quite significant such that the scattered laser suppression is challenging. To take into account the non-filtered laser contribution we subtract the background by recording reference time-trace with the initial charge injection-pulse switched off. In Fig. 4d we plot the emission signal after the spin-rotation pulse as a function of the square root of this pulse power (blue data points). Clear oscillatory behavior can be observed, related to Rabi rotations up to 3π angle. The evolution of the Bloch vector trajectory by angle θ during this process is schematically shown in the Fig. 4d inset. To estimate the off-resonant incoherent trion excitation contribution due to the rotation pulse, we record the emission time-trace with the cw spin-pumping laser switched off (red curve Fig. 4c). For the rotation powers above 100 μW, the incoherent part of the emission becomes dominant over the coherent part. The oscillation amplitude damping is most likely related to the combination of a few effects, such as finite rotation pulse length in respect to Larmor period (4 vs. 17.5 ps), presence of cw spin-pumping/read-out laser during the spin rotation and slight tilt of the rotation vector due to the QD asymmetry and non-perfect circular polarization of the rotation pulse.
Ramsey interference and full coherent control
To access all the qubit states on the Bloch sphere, spin rotation by a second axis is required. For that, we utilize the inherent Larmor precession of the spin state around the applied magnetic field46. This can be probed by the means of Ramsey interferometry. For that population is probed after two π/2 rotation pulses separated by the variable time delay τ. The pulse sequence used and the corresponding evolution of the spin-state trajectory on the Bloch sphere are shown in Fig. 5a. After a first π/2 rotation pulse, the hole spin is rotated from the north pole to the equator, where it is allowed to freely precess around the external magnetic field with a Larmor period defined by - splitting. By delaying a second π/2 pulse by an amount τ, Ramsey fringes are observed. In Fig. 5b Ramsey interference for a pair π/2 pulses is shown, demonstrating integrated emission counts as a function of the time delay τ between the pulses. The data are fitted with a damped sine function exhibiting a Larmor period of 17.44 ± 0.03 ps (57.3 ± 0.2 GHz hole splitting). Additionally, in Fig. 5c the amplitude of the Ramsey fringes, as a function of the delay time between the pulses is shown. The data are fitted by an exponential decay, revealing the inhomogeneous dephasing time of 240 ± 30 ps. While typically on the level of nanoseconds is observed in QDs11,12,40, we relate the value recorded in our case, to the presence of the cw pumping between the rotation pulses46, thus, could be potentially increased by temporal modulation of the pumping laser. The fringe contrast of the first Ramsey period reaches the level of 85% (after correcting for incoherent contribution).
Finally, to access any arbitrary state on the Bloch sphere, we demonstrate simultaneous control of θ and ϕ rotation angles by adjusting the power of the pulses and delay time τ between them, using a pulse sequence as shown in Fig. 5d. In Fig. 5e we plot the QD integrated emission counts as a function of rotation pulse area—θ and delay time between the pulses τ. For θ increasing from 0 to π/2 the Ramsey fringe amplitude rises and achieves a maximum value at π/2. When θ is further increased from π/2 to π, the fringe amplitude decreases and a phase shift is observed. Ramsey interference fringes recorded for π/2 and π pulse areas are in clear anti-phase as shown in Fig. 5f. At θ = π the oscillations are supposed to vanish completely, however, some small signal is remaining. We relate this to the non-negligible spin Larmor precession with respect to the effective Rabi rotation frequency (nonzero pulse length, the QD asymmetry). Access to θ rotation in the range of 0 to π and ϕ of 0 to 2π (one Larmor period) allows to explore the entire surface of the Bloch sphere and constitutes full coherent control of the spin-qubit. Taking into account the rotation pulse length and Larmor precession period, a single gate operation within our system could be performed within a 26 ps time window.
Discussion
In conclusion, we have demonstrated optical spin injection, initialization, read-out and full coherent control of a spin-qubit system based on a hole confined in a single InAs/GaAs QD grown on a MMB layer. The measurements performed highlight this system as particularly promising for long-distance quantum network applications thanks to the possibility of interfacing spins with photons directly at C-band telecom frequencies. The QD-spin coherent control was achieved using spin rotations by two perpendicular axes using a detuned pair of picosecond pulses. Furthermore, Ramsey interferometry allowed us to probe the spin dephasing time, which in our case was limited to 240 ps due to the presence of optical pumping during the spin-rotation sequence. Regular InAs/GaAs QD systems operating at shorter wavelengths and driven by modulated initialization/read-out sequences have shown the times of over few ns11,12,40.
While the intrinsic coherence limitations of the spin system investigated in this work remain to be revealed in further studies, the presence of strain-relaxed layers suggest that it could outperform the typical well-established InAs/GaAs QDs. It was shown that reducing the lattice strain in semiconductor QDs increases the spin-coherence by limiting the electric quadrupolar effects which play a dominant role in hole hyperfine interaction with nuclear spins47,48. In case of the hole qubits, less strain translates also into smaller light-hole admixture, which was shown to limit the spin-hole lifetime49,50 and coherence time48.
Generating the spin–photon and spin–spin entanglement with the current device architecture should be realistically achieved in the near future. Thanks to mature GaAs fabrication technology, our QD-spin system could be further combined with cavities and other photonic structures. This in turn might allow increasing the photon collection efficiency, speed up the spin-initialization process and achieve high optical transition cyclicity51 for single-shot read-out. Further performance improvement could be achieved by utilizing an n–i–p diode structure, which was shown to allow trapping the charge state via Coulomb blockade, tuning the emission frequency using gate voltage and reducing the charge noise50. These advances combined with emission in the telecom wavelengths will be of immediate relevance for long-distance quantum networks technologies based on optical fibers or satellites.
Methods
Sample description
The QD sample used in this study was grown by metalorganic vapor phase epitaxy (MOVPE) on (100) GaAs substrate. First, the 200 nm buffer layer of GaAs was deposited, followed by 20 alternating DBR pairs of AlAs/GaAs with thickness 134.4 and 114.6 nm, respectively. Next, a 1080 nm thick InGaAs metamorphic buffer (MMB) layer was grown by linearly increasing the In flux over the time from 0 to 17 μmol/min. The indium concentration increase with the thickness in the MMB allows increasing the relaxation of the lattice constant. Finally, InAs material was deposited to form the QD layer, following capping with 220 nm of In0.25Ga0.75As. The bottom DBR and the top InGaAs–air interface nominally form a 3-λ cavity optimized at 1550 nm and with 100 nm stop-band width. The QD layer is placed in the cavity mode anti-node for increased photon extraction efficiency from the top. The schematic sample layer structure is shown in Fig. 1a. The QD areal density varies on the sample position and is estimated to be 107 cm−2. More details regarding the sample fabrication process can be found in ref. 35.
Experimental setup
The sample is cooled down to a temperature of 1.6 K inside a closed-cycle liquid helium cryostat (Attodry 2100). The sample is placed in a dipstick cage-system, in which helium gas acts as an exchange gas. Superconducting magnet surrounding the sample chamber is used to apply a magnetic field up to 8 T. The optical part of the setup consists of a home-built confocal microscope mounted on top of the cryostat, and an apochromatic lens (NA = 0.68) inside the sample chamber. The sample is mounted on piezoelectric stages (Attocube) allowing to position the sample with respect to the lens. Nonresonant excitation of the QDs is performed using a mode-locked Ti:Sa pulsed laser (Coherent Mira) at 850 nm and 76 MHz repetition rate (2 ps pulse width). For resonant excitation, two lasers are used: (i) continuous wave tunable diode laser at 1550–1570 nm (iBlue ModBox) and (ii) tunable Optical Parametric Oscillator (Coherent OPO-X) pumped by the Ti:Sa laser at 850 nm. For the Ramsey interference experiments, a laser beam was passed through the free-space Mach-Zender interferometer where the time delay between the pulses is controlled via a retro-reflector mounted on the motorized translation stage. The fluorescence signal is collected through the same lens. A 1200 nm long-pass filter in the confocal microscope filters out the 850 nm laser excitation. Furthermore, polarization optics is setup in cross-polarization configuration for resonant laser suppression. Additional suppression of the laser is achieved by spatial filtering using single-mode fiber. Emission is then sent to a 75 cm focal length monochromator where it could be resolved spectrally and imaged on the array detector (Princeton Instruments NIRvana) or filtered spectrally and coupled to superconducting single-photon counting detector (Single Quantum EOS) with a time resolution better than 20 ps. The signal from the single-photon counting detectors is sent to a time-tagger module (Pico Harp 300) triggered by the laser. The total timing resolution of the setup (instrumental response function fwhm) is around 30 ps. The experimental setup scheme can be found in the Supplementary Note 1.
Data analysis
The non-filtered laser contribution to the time-resolved histograms (visible under high pumping powers) is subtracted from the data by recording the reference histograms under identical conditions but with the charge injection laser (850 nm) switched off. Data shown in Fig. 5b, c, e, f are corrected for the incoherent emission contribution, decreasing the Ramsey fringes contrast from 85% to 72% under π/2 driving. In Fig. 3d there are two very small intensity peaks visible at 2.5 and 7.7 ns time. Those artefacts are related nonideal 76 MHz read-out laser pulse sequence, consisting of the main pulse and two extra satellite pulses at ±3 ns with around two orders of magnitude lower intensity.
Supplementary information
Acknowledgements
We thank Yu-Ming He and Chao-Yang Lu for fruitful discussions at the beginning stages of this study. This work was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)-INST 93/932-1 FUGG. We acknowledge financial support by the German Ministry of Education and Research (BMBF) within the project “Q.Link.X” (FKZ: 16KIS0871 and 16KIS0862). We are furthermore grateful for the support by the State of Bavaria.
Author contributions
C.N., M.J., S.L.P., and P.M. designed and provided the wafer. Ł.D. established an experimental setup, carried out the optical experiments, analyzed, and interpreted the experimental data. T.H. and S.K. helped with the project administration and acquiring the funding. S.H. and P.M. guided the work. Ł.D. wrote the paper with input from all authors.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.
Funding
Open Access funding enabled and organized by Projekt DEAL.
Data availability
The data that support the findings of this study are available from the corresponding author on reasonable request.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-022-28328-2.
References
- 1.Kimble HJ. The quantum internet. Nature. 2008;453:1023–1030. doi: 10.1038/nature07127. [DOI] [PubMed] [Google Scholar]
- 2.Chen Y-A, et al. An integrated space-to-ground quantum communication network over 4,600 kilometres. Nature. 2021;589:214. doi: 10.1038/s41586-020-03093-8. [DOI] [PubMed] [Google Scholar]
- 3.Togan E, et al. Quantum entanglement between an optical photon and a solid-state spin qubit. Nature. 2010;466:730–734. doi: 10.1038/nature09256. [DOI] [PubMed] [Google Scholar]
- 4.De Greve K, et al. Quantum-dot spin-photon entanglement via frequency downconversion to telecom wavelength. Nature. 2012;491:421. doi: 10.1038/nature11577. [DOI] [PubMed] [Google Scholar]
- 5.Gao WB, Fallahi P, Togan E, Miguel-Sanchez J, Imamoglu A. Observation of entanglement between a quantum dot spin and a single photon. Nature. 2012;491:426–430. doi: 10.1038/nature11573. [DOI] [PubMed] [Google Scholar]
- 6.Schaibley JR, et al. Demonstration of quantum entanglement between a single electron spin confined to an InAs quantum dot and a photon. Phys. Rev. Lett. 2013;110:1. doi: 10.1103/PhysRevLett.110.167401. [DOI] [PubMed] [Google Scholar]
- 7.Atatüre M, Englund D, Vamivakas N, Lee S-Y, Wrachtrup J. Material platforms for spin-based photonic quantum technologies. Nat. Rev. Mater. 2018;3:38–51. [Google Scholar]
- 8.Awschalom DD, Hanson R, Wrachtrup J, Zhou BB. Quantum technologies with optically interfaced solid-state spins. Nat. Photon. 2018;12:516–527. [Google Scholar]
- 9.Senellart P, Solomon G, White A. High-performance semiconductor quantum-dot single-photon sources. Nat. Nanotechnol. 2017;12:1026–1039. doi: 10.1038/nnano.2017.218. [DOI] [PubMed] [Google Scholar]
- 10.Michler, P. Quantum dots for quantum information technologies, Nano-optics and nanophotonics (Springer International Publishing, 2017).
- 11.Bechtold A, et al. Three-stage decoherence dynamics of an electron spin qubit in an optically active quantum dot. Nat. Phys. 2015;11:1005–1008. [Google Scholar]
- 12.Stockill R, et al. Quantum dot spin coherence governed by a strained nuclear environment. Nat. Commun. 2016;7:12745. doi: 10.1038/ncomms12745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.He Y, et al. Quantum state transfer from a single photon to a distant quantum-dot electron spin. Phys. Rev. Lett. 2017;119:060501. doi: 10.1103/PhysRevLett.119.060501. [DOI] [PubMed] [Google Scholar]
- 14.Delteil A, et al. Generation of heralded entanglement between distant hole spins. Nat. Phys. 2016;12:218–223. [Google Scholar]
- 15.Stockill R, et al. Phase-tuned entangled state generation between distant spin qubits. Phys. Rev. Lett. 2017;119:010503. doi: 10.1103/PhysRevLett.119.010503. [DOI] [PubMed] [Google Scholar]
- 16.Abobeih MH, et al. One-second coherence for a single electron spin coupled to a multi-qubit nuclear-spin environment. Nat. Commun. 2018;9:2552. doi: 10.1038/s41467-018-04916-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Hensen B, et al. Loophole-free Bell inequality violation using electron spins separated by 1.3 kilometres. Nature. 2015;526:682. doi: 10.1038/nature15759. [DOI] [PubMed] [Google Scholar]
- 18.Koehl WF, Buckley BB, Heremans FJ, Calusine G, Awschalom DD. Room temperature coherent control of defect spin qubits in silicon carbide. Nature. 2011;479:84–87. doi: 10.1038/nature10562. [DOI] [PubMed] [Google Scholar]
- 19.Christle DJ, et al. Isolated electron spins in silicon carbide with millisecond coherence times. Nature Materials. 2015;14:160–163. doi: 10.1038/nmat4144. [DOI] [PubMed] [Google Scholar]
- 20.Kindem JM, et al. Control and single-shot readout of an ion embedded in a nanophotonic cavity. Nature. 2020;580:201–204. doi: 10.1038/s41586-020-2160-9. [DOI] [PubMed] [Google Scholar]
- 21.Raha M, et al. Optical quantum nondemolition measurement of a single rare earth ion qubit. Nat. Commun. 2020;11:1605. doi: 10.1038/s41467-020-15138-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Chen S, Raha M, Phenicie CM, Ourari S, Thompson JD. Parallel single-shot measurement and coherent control of solid-state spins below the diffraction limit. Science. 2020;370:592. doi: 10.1126/science.abc7821. [DOI] [PubMed] [Google Scholar]
- 23.Rančić M, Hedges MP, Ahlefeldt RL, Sellars MJ. Coherence time of over a second in a telecom-compatible quantum memory storage material. Nat. Phys. 2018;14:50–54. [Google Scholar]
- 24.Wang JF, et al. Coherent control of nitrogen-vacancy center spins in silicon carbide at room temperature. Phys. Rev. Lett. 2020;124:223601. doi: 10.1103/PhysRevLett.124.223601. [DOI] [PubMed] [Google Scholar]
- 25.Bergeron L, et al. Silicon-integrated telecommunications photon-spin interface. PRX Quantum. 2020;1:020301. [Google Scholar]
- 26.Bock M, et al. High-fidelity entanglement between a trapped ion and a telecom photon via quantum frequency conversion. Nat. Commun. 2018;9:1998. doi: 10.1038/s41467-018-04341-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Weber JH, et al. Two-photon interference in the telecom C-band after frequency conversion of photons from remote quantum emitters. Nat. Nanotechnol. 2019;14:23–26. doi: 10.1038/s41565-018-0279-8. [DOI] [PubMed] [Google Scholar]
- 28.Mrowiński P, et al. Magnetic field control of the neutral and charged exciton fine structure in single quantum dashes emitting at 1.55 μm. Appl. Phys. Lett. 2015;106:053114. [Google Scholar]
- 29.Ha N, et al. Single photon emission from droplet epitaxial quantum dots in the standard telecom window around a wavelength of 1.55 μm. Appl. Phys. Express. 2020;13:025002. [Google Scholar]
- 30.Anderson M, et al. Coherence in single photon emission from droplet epitaxy and Stranski-Krastanov quantum dots in the telecom C-band. Appl. Phys. Lett. 2021;118:014003. [Google Scholar]
- 31.Kettler J, et al. Neutral and charged biexciton-exciton cascade in near-telecom-wavelength quantum dots. Phys. Rev. B. 2016;94:045303. [Google Scholar]
- 32.Mrowiński P, et al. Excitonic complexes in MOCVD-grown InGaAs/GaAs quantum dots emitting at telecom wavelengths. Phys. Rev. B. 2019;100:115310. [Google Scholar]
- 33.Nawrath C, et al. Coherence and indistinguishability of highly pure single photons from non-resonantly and resonantly excited telecom C-band quantum dots. Appl. Phys. Lett. 2019;115:023103. [Google Scholar]
- 34.Rudno-Rudziński, W., Burakowski, M., Reithmaier, J. P., Musiał, A. & Benyoucef, M. Magneto-optical characterization of trions in symmetric InP-Based quantum dots for quantum communication applications. Materials14, 942 10.3390/ma14040942 (2021). [DOI] [PMC free article] [PubMed]
- 35.Paul M, et al. Single-photon emission at 1.55 μm from MOVPE-grown InAs quantum dots on InGaAs/GaAs metamorphic buffers. Appl. Phys. Lett. 2017;111:033102. [Google Scholar]
- 36.Carmesin C, et al. Structural and optical properties of InAs/(In)GaAs/GaAs quantum dots with single-photon emission in the telecom C-band up to 77 K. Phys. Rev. B. 2018;98:125407. [Google Scholar]
- 37.Brunner D, et al. A coherent single-hole spin in a semiconductor. Science. 2009;325:70. doi: 10.1126/science.1173684. [DOI] [PubMed] [Google Scholar]
- 38.Fallahi P, Yılmaz ST, Imamoğlu A. Measurement of a heavy-hole hyperfine interaction in InGaAs quantum dots using resonance fluorescence. Phys. Rev. Lett. 2010;105:257402. doi: 10.1103/PhysRevLett.105.257402. [DOI] [PubMed] [Google Scholar]
- 39.Chekhovich EA, Krysa AB, Skolnick MS, Tartakovskii AI. Direct measurement of the hole-nuclear spin interaction in single InP/GaInP quantum dots using photoluminescence spectroscopy. Phys. Rev. Lett. 2011;106:027402. doi: 10.1103/PhysRevLett.106.027402. [DOI] [PubMed] [Google Scholar]
- 40.De Greve K, et al. Ultrafast coherent control and suppressed nuclear feedback of a single quantum dot hole qubit. Nat. Phys. 2011;7:872–878. [Google Scholar]
- 41.Greilich A, Carter SG, Kim D, Bracker AS, Gammon D. Optical control of one and two hole spins in interacting quantum dots. Nat. Photon. 2011;5:702–708. [Google Scholar]
- 42.Godden TM, et al. Coherent optical control of the spin of a single hole in an InAs/GaAs quantum dot. Phys. Rev. Lett. 2012;108:017402. doi: 10.1103/PhysRevLett.108.017402. [DOI] [PubMed] [Google Scholar]
- 43.Prechtel JH, et al. Decoupling a hole spin qubit from the nuclear spins. Nat. Mater. 2016;15:981–986. doi: 10.1038/nmat4704. [DOI] [PubMed] [Google Scholar]
- 44.Atatüre M, et al. Quantum-dot spin-state preparation with near-unity fidelity. Science. 2006;312:551. doi: 10.1126/science.1126074. [DOI] [PubMed] [Google Scholar]
- 45.Gerardot BD, et al. Optical pumping of a single hole spin in a quantum dot. Nature. 2008;451:441–444. doi: 10.1038/nature06472. [DOI] [PubMed] [Google Scholar]
- 46.Press D, Ladd TD, Zhang B, Yamamoto Y. Complete quantum control of a single quantum dot spin using ultrafast optical pulses. Nature. 2008;456:218–221. doi: 10.1038/nature07530. [DOI] [PubMed] [Google Scholar]
- 47.Bulutay C. Quadrupolar spectra of nuclear spins in strained InGaAs quantum dots. Phys. Rev. B. 2012;85:115313. [Google Scholar]
- 48.Urbaszek B, et al. Nuclear spin physics in quantum dots: an optical investigation. Rev. Mod. Phys. 2013;85:79. [Google Scholar]
- 49.Krzykowski M, Gawarecki K, Machnikowski P. Hole spin-flip transitions in a self-assembled quantum dot. Phys. Rev. B. 2020;102:205301. [Google Scholar]
- 50.Zhai L, et al. Low-noise GaAs quantum dots for quantum photonics. Nat. Commun. 2020;11:4745. doi: 10.1038/s41467-020-18625-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Appel MH, et al. Coherent spin-photon interface with waveguide induced cycling transitions. Phys. Rev. Lett. 2021;126:013602. doi: 10.1103/PhysRevLett.126.013602. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data that support the findings of this study are available from the corresponding author on reasonable request.