Skip to content
BY 4.0 license Open Access Published online by De Gruyter May 9, 2024

Reconfigurable quantum photonic circuits based on quantum dots

  • Adam McCaw , Jacob Ewaniuk ORCID logo , Bhavin J. Shastri ORCID logo and Nir Rotenberg ORCID logo EMAIL logo
From the journal Nanophotonics

Abstract

Quantum photonic integrated circuits, composed of linear-optical elements, offer an efficient way for encoding and processing quantum information on-chip. At their core, these circuits rely on reconfigurable phase shifters, typically constructed from classical components such as thermo- or electro-optical materials, while quantum solid-state emitters such as quantum dots are limited to acting as single-photon sources. Here, we demonstrate the potential of quantum dots as reconfigurable phase shifters. We use numerical models based on established literature parameters to show that circuits utilizing these emitters enable high-fidelity operation and are scalable. Despite the inherent imperfections associated with quantum dots, such as imperfect coupling, dephasing, or spectral diffusion, we show that circuits based on these emitters may be optimized such that these do not significantly impact the unitary infidelity. Specifically, they do not increase the infidelity by more than 0.001 in circuits with up to 10 modes, compared to those affected only by standard nanophotonic losses and routing errors. For example, we achieve fidelities of 0.9998 in quantum-dot-based circuits enacting controlled-phase and – not gates without any redundancies. These findings demonstrate the feasibility of quantum emitter-driven quantum information processing and pave the way for cryogenically-compatible, fast, and low-loss reconfigurable quantum photonic circuits.

1 Introduction

Reconfigurable quantum photonic integrated circuits (qPICs) are versatile tools capable of simulating molecular dynamics [1], executing quantum logic [2], and generating multidimensional entanglement [3]–[7]. They utilize quantum properties such as entanglement and indistinguishability for information processing, which is unachievable through classical means. This capability is crucial for developing emerging quantum communication and computation technologies.

To date, qPICs have predominantly operated at room temperature, harnessing the advancements of foundry photonics to create increasingly complex devices [8], [9]. As depicted in Figure 1, the core of these circuits is a mesh of Mach–Zehnder interferometers (MZIs) [10], where each MZI is comprised of two directional couplers and two phase shifters (see Figure 1(b)), typically thermo-optical in nature [11], [12]. That is, this mesh is fully classical, based on low-loss and highly accurate (albeit slow and hot) phase shifters. More recently, electro-optic [13], [14] and opto-mechanical [15], [16] phase shifters have become popular alternatives, however, they are limited to large footprints ( > 100 μ m lengths) and MHz modulation speeds, respectively. Although any mesh, regardless of the type of phase shifter, is susceptible to unbalanced nanophotonic imperfections including losses and imperfect routing, high-fidelity operation is achievable with redundancy and optimization [17]–[22].

Figure 1: 
Scalable reconfigurable MZI meshes. (a) An exemplary 4-mode MZI mesh colored to represent nanophotonic losses in each interferometer. (b) Normal distribution for nanophotonic loss with a mean of 4.5 % and standard deviation that is 5 % of the mean. (c) A zoom-in on the basic component of the mesh (dashed region in a), a MZI with two phase shifters ϕ and 2θ and two imperfect beam splitters along with the imperfect transfer matrices for each component and for the overall nanophotonic loss.
Figure 1:

Scalable reconfigurable MZI meshes. (a) An exemplary 4-mode MZI mesh colored to represent nanophotonic losses in each interferometer. (b) Normal distribution for nanophotonic loss with a mean of 4.5 % and standard deviation that is 5 % of the mean. (c) A zoom-in on the basic component of the mesh (dashed region in a), a MZI with two phase shifters ϕ and 2θ and two imperfect beam splitters along with the imperfect transfer matrices for each component and for the overall nanophotonic loss.

The quantum nature of qPICs stems from the individual photons that propagate through the circuits. Single photons are often generated by optical nonlinearities such as spontaneous parametric down-conversion, which is compatible with room-temperature chips but inherently probabilistic [23]–[25]. In contrast, single-photon sources based on solid-state quantum emitters, such as the quantum dots (QDs) we consider, can operate on-demand but require cryogenic temperatures [26]–[29]. Similarly, high-efficiency integrated single-photon detectors also necessitate a cryogenic environment [30]–[32]. As a result, both state-of-the-art sources and detectors cannot be heterogeneously integrated with qPICs and instead currently rely on lossy interconnects [33].

Here, we propose that QDs can be used not only as single-photon sources, but also as reconfigurable phase shifters for creating fast, cryogenically-compatible meshes. As photons scatter from QDs, and indeed all solid-state emitters, the imparted phase shift depends on the detuning between the photon and emitter transition frequency. For the case of QDs, the detuning can be modulated electrically [29], [34], optically [35], [36], or by using strain [37]. Our model of QD-based qPICs builds from this operating principle, yet additionally incorporates standard nanophotonic imperfections, such as losses and routing errors, along with QD-specific non-idealities, like imperfect interactions and both fast and slow noise processes. We use this model to evaluate the fidelity of both the resultant unitary operations and the desired output states. Our findings reveal that these QD-based meshes can be optimized to achieve remarkable scalability, with a unitary infidelity less than 0.001 for circuits up to 10 × 10 in dimension, using state-of-the-art QD parameters from the literature. We further consider QD-based controlled-phase and – not gates as examples, where we find that state-of-the-art circuits process logical states with fidelities of 0.9998. In sum, our results offer a roadmap to cryogenically-compatible, reconfigurable qPICs based on solid-state quantum emitters.

2 Quantum-emitter phase shifters

The scattering of photons from a quantum emitter embedded in a nanophotonic waveguide is a complex process [38] that may modulate the photons’ amplitude or phase [39], [40] or, when more than a single photon is present, induce complex correlations [35], [41]. The exact response depends on the properties of the emitter and the efficiency with which it couples to the various available modes, as sketched in Figure 2(a), yet in the most general case, an input single-photon state with phase φ0, 1 in = φ 0 , will scatter into the mixed state,

(1) 1 out = α co φ 0 + Δ φ α inc φ 0 ,

where, in the presence of losses the sum of the probabilities to scatter coherently and incoherently, |αco|2 and |αinc|2, do not sum to unity.

Figure 2: 
Waveguide-coupled chiral quantum dots (QDs). (a) A QD chirally-coupled to a waveguide with its forward direction to the right, βR ≫ βL, imparts a phase shift Δφ on coherently-scattered photons but not to those with which it interacts incoherently (due to fast dephasing). If βR < 1, then some photons are either reflected or scattered out of the waveguide. (b) Transmission spectrum of the interaction shown in (a). In the ideal case (βR = 1 and no noise, solid orange curve) all photons are transmitted, while if βR = 0.9 and Γdp = 0.1Γ then the presence of the QD will be imprinted on the total transmission spectrum (solid red curve) and the coherent transmission spectrum (solid purple curve). The associated phase shift applied to the coherently-scattered photons in both the ideal (orange) and non-ideal (purple) scenarios is shown in the dashed curves (right axis), which are nearly identical such that they almost completely overlap. (c) On-resonance phase shift as a function of directionality and dephasing rate (inset: similar map for a photon-emitter detuning of Δp = 0.3Γ). The large region over which Δφ = π demonstrates the robustness of QD-based phase shifters.
Figure 2:

Waveguide-coupled chiral quantum dots (QDs). (a) A QD chirally-coupled to a waveguide with its forward direction to the right, βRβL, imparts a phase shift Δφ on coherently-scattered photons but not to those with which it interacts incoherently (due to fast dephasing). If βR < 1, then some photons are either reflected or scattered out of the waveguide. (b) Transmission spectrum of the interaction shown in (a). In the ideal case (βR = 1 and no noise, solid orange curve) all photons are transmitted, while if βR = 0.9 and Γdp = 0.1Γ then the presence of the QD will be imprinted on the total transmission spectrum (solid red curve) and the coherent transmission spectrum (solid purple curve). The associated phase shift applied to the coherently-scattered photons in both the ideal (orange) and non-ideal (purple) scenarios is shown in the dashed curves (right axis), which are nearly identical such that they almost completely overlap. (c) On-resonance phase shift as a function of directionality and dephasing rate (inset: similar map for a photon-emitter detuning of Δp = 0.3Γ). The large region over which Δφ = π demonstrates the robustness of QD-based phase shifters.

The description of Eq. (1) only holds if the single-photon pulse is not reshaped during scattering, requiring that the pulse linewidth σp be much shorter than that of the emitter. More formally, we require that σp ≤ Γ/1000, ensuring the phase shifter response is linear. As shown in the Supplementary Information, this condition holds for any two-level system, including QDs.

In this regime, the single-photon transmission coefficient t and total transmission T are the same as that of a weak coherent beam (see Supplementary Information for derivation),

(2) t = 1 Γ β R Γ 2 + i Δ p Γ 2 2 + Δ p 2 ,
(3) T = 1 2 Γ Γ 2 β R ( 1 β R ) Γ 2 2 + Δ p 2 ,

where βR is the coupling efficiency for right-traveling photons (with a total coupling efficiency β = βR + βL), Δp is the detuning between the photon and emitter-transition frequencies, Γ is the decay rate of the emitter, and Γ2 = Γ/2 + Γdp where Γdp is the pure dephasing rate (i.e. fast noise). We note that in the presence of dephasing, T t 2 (c.f. Figure 2(b)), and that the emitter might also suffer from slow noise leading to spectral diffusion with a characteristic linewidth σsd.

Eqs. (2) and (3) allow us to quantify the results of the scattering. The coefficients, αco and αinc, are related to the transmission, as shown in Figure 2(b). In the ideal case, where βR = 1 and Γdp = 0, the transmission is always unity (orange curve) meaning that αco = 1 and αinc = 0. Conversely, in the presence of losses and/or dephasing, the situation is more complex with α co 2 = t 2 and α inc 2 = T t 2 as shown by the purple and red curves.

The phase of the coherently-scattered photons is likewise calculated from Δ φ = arg t , here shown in dashed curves in Figure 2(b) corresponding to the ideal and non-ideal scenarios. As can be seen, the imparted phase shift is nearly identical in both cases, spanning the full 2π and demonstrating the robustness of emitter-based phase shifters. Full, 2D maps of the induced phase shift on resonance, Δp = 0, and at Δp = 0.3Γ, are shown in Figure 2(c) and its inset, respectively, again demonstrating that a 2π phase change is possible even when the directionality is below D < 0.25 or the dephasing rate is above Γdp > 0.3Γ. However, this range also depends on emitter parameters such as β. Corresponding maps of α co 2 are presented in the Supplementary Information. Together, these enable us to pick an emitter detuning for each desired phase shift, and then calculate the scattered state (c.f. Eq. (1)).

3 QD-based qPICs

Having seen that quantum emitters such as QDs can serve as reconfigurable phase shifters, we quantify the performance of qPICs based on this technology. To do so, we first compare how well we can reproduce any unitary (i.e. operation) with our emitter-based qPICs relative to the ideal, summarizing the results in Figure 3. Here, we show the dependence of the mean circuit infidelity I (i.e. error) as a function of (a) β, (b) D, (c) Γdp and (d) σsd, where, in all, I is limited by the nanophotonic errors as noted in the caption (see the Supplementary Information for details on how these were selected).

Figure 3: 
The effects of QD imperfections on circuit unitary infidelity. All plots have 4.5 % nanophotonic loss [8] and 4 % beam splitter error [21]. (a–d) With no other QD imperfections, we examine the effects of (a) coupling efficiency, β, (b) directionality, D, (c) pure dephasing rate, Γdp, and (d) spectral diffusion, σsd on infidelity, showing both the imperfect (dashed) and optimized (solid) results. In all cases, state-of-the-art QD parameters from literature are indicated by vertical dashed lines. (e) The infidelity of QD-based qPICs as a function of mesh size ranging from N = 2 to 10 for circuits suffering only from nanophotonic imperfections (red), as well as those that additionally suffer from state-of-the-art QD imperfections (green) or typical QD imperfections (blue). See Table 1 for QD parameter values. Also shown are infidelities for circuits based on typical QDs but no noise (orange) or with only fast dephasing (purple), demonstrating that circuit errors are largely due to dephasing.
Figure 3:

The effects of QD imperfections on circuit unitary infidelity. All plots have 4.5 % nanophotonic loss [8] and 4 % beam splitter error [21]. (a–d) With no other QD imperfections, we examine the effects of (a) coupling efficiency, β, (b) directionality, D, (c) pure dephasing rate, Γdp, and (d) spectral diffusion, σsd on infidelity, showing both the imperfect (dashed) and optimized (solid) results. In all cases, state-of-the-art QD parameters from literature are indicated by vertical dashed lines. (e) The infidelity of QD-based qPICs as a function of mesh size ranging from N = 2 to 10 for circuits suffering only from nanophotonic imperfections (red), as well as those that additionally suffer from state-of-the-art QD imperfections (green) or typical QD imperfections (blue). See Table 1 for QD parameter values. Also shown are infidelities for circuits based on typical QDs but no noise (orange) or with only fast dephasing (purple), demonstrating that circuit errors are largely due to dephasing.

As an example, consider the β-dependence of I , shown in Figure 3(a). In dashed curves, we show the infidelity of the non-ideal unitary Unon, constructed following the procedure of [10] using the set of phases 2 θ i , ϕ i (c.f. Figure 1(c)) calculated for the ideal circuit, but with imperfections subsequently added (see Supplementary Information for details). This corresponds to the offline training of a photonic circuit. For each coupling efficiency β, we compare 100 non-ideal unitaries Unon to the ideal U to calculate [10]

(4) I = 1 t r U U non N t r U non U non 2 ,

which we then average. In the figure, we show the cases for N = 2, 6, 10 mode circuits, where in all cases, we observe a monotonic increase from a baseline I due to the nanophotonic imperfections when β = 1, to near-unity error when the coupling efficiency is β = 0.5.

Encouragingly, we can optimize the performance of the emitter-based qPICs, following a fast and efficient routine that finds the optimal set 2 θ i , ϕ i at once [42]. This procedure results in an optimized unitary Uopt that, together with Eq. (4), enables us to again calculate I , which we show as solid curves in Figure 3. In all cases, we observe a significant reduction in errors due to the optimization, with this reduction particularly pronounced for larger circuit sizes and in near-optimal conditions. For example, in Figure 3(a), as β → 1 corresponding to state-of-the-art performance (dashed line) [43], we observe almost no increased infidelity as the circuit size increases. A similar dependence is observed as D and Γdp are scanned (Figure 3(b) and (c), respectively), while spectral diffusion only begins affecting the performance when σsd ≳ 0.01Γ (Figure 3(d)).

Overall, we summarize the circuit scaling in Figure 3(e), where we plot the raw and optimized infidelity as a function of circuit size, both using typical and state-of-the-art parameters (see Table 1). For typical values (blue curve), we see that the infidelity quickly approaches unity, yet by adding imperfections sequentially (purple and orange curves), we see that this is almost entirely caused by the residual dephasing. This is consistent with the optimized state-of-the-art qPIC performance (green curve), where I < 0.006 is observed for all circuits simulated (up to N = 10), as several recent experiments based on QDs have measured Γdp ≈ 0 [35], [44]–[47].

Table 1:

State-of-the-art and typical quantum dot parameters. Listed parameters include coupling, β, directionality, D, dephasing, Γdp, spectral diffusion detuning standard deviation, σsd. Though all parameters are taken from the literature, and are justifiable individually, it should be noted that they correspond to different types of waveguides. For more details on these parameters, including the reported uncertainties, see Section S3.1 of the Supplementary Information.

Typical Ref(s) State-of-the-art Ref(s)
β 0.90 [44] 0.9843 [43]
D 0.90 [44] 0.95 [48]
Γdp 0.01Γ [35] [35], [44]–[47]
σ sd 0.06Γ [49] [50], [51]

4 Examples: CZ gate

To demonstrate the possibilities of emitter-based qPICs, we consider the controlled-phase (CZ) gate, which can be used to generate entanglement [52], and, in the Supplementary Information, a similar realization of a controlled-not (CNOT) gate that enables universal quantum computation [53]. A linear-optical unheralded CZ gate can be realized on a 6 × 6 mesh [52], and in Figure 4(a) we plot the unitary infidelity I (Eq. (4)) calculated as a function of Γdp for circuits with only nanophotonic imperfections (red curve, no dephasing), the addition of typical QD imperfections (blue curve), and state-of-the-art QD imperfections (green curve), where for the latter two cases the dephasing varies along the horizontal axis (different from Table 1). For typical parameters, I > 0.25 regardless of the dephasing rate, consistent with Figure 3. Interestingly, the optimized circuit based on typical QD parameters is nearly identical to that based on state-of-the-art when Γdp > 0.01Γ, meaning that in this regime, the effect of dephasing dominates all other emitter parameters. In contrast, the optimized curve for state-of-the-art parameters, including all QD imperfections (green), tends toward that for conventional phase shifters (red) in the limit that dephasing approaches its state-of-the-art value (Γdp ≈ 0). Additionally, while circuits with only nanophotonic imperfections achieve an optimized unitary infidelity of 0.0023, independent of Γdp, it increases only to 0.0037 at Γdp = 10−8Γ, 0.0065 at Γdp = 10−6Γ, and remains less than 0.01 up to Γdp ∼ 10−5Γ.

Figure 4: 
Unheralded CZ gate performance for nanophotonic, state-of-the-art, and typical imperfections (as previously explained in Figure 3 and Table 1). (a) Imperfect (dashed) and optimized (solid) unitary infidelities (


I

$\mathcal{I}$


, c.f. Eq. (4)) as a function of dephasing. All QD parameters other than Γdp are as listed in Table 1. Since circuits with only nanophotonic imperfections do not suffer from dephasing, their performance is flat in Γdp. (b–d) Optimized nanophotonic, state-of-the-art, and typical post-selected 4 × 4 computational basis matrices for the unheralded CZ gate, with corresponding conditional output state fidelities listed below each matrix.
Figure 4:

Unheralded CZ gate performance for nanophotonic, state-of-the-art, and typical imperfections (as previously explained in Figure 3 and Table 1). (a) Imperfect (dashed) and optimized (solid) unitary infidelities ( I , c.f. Eq. (4)) as a function of dephasing. All QD parameters other than Γdp are as listed in Table 1. Since circuits with only nanophotonic imperfections do not suffer from dephasing, their performance is flat in Γdp. (b–d) Optimized nanophotonic, state-of-the-art, and typical post-selected 4 × 4 computational basis matrices for the unheralded CZ gate, with corresponding conditional output state fidelities listed below each matrix.

More significantly, we consider the fidelity with which logical states are processed by the CZ gate ( F ) across Figure 4(b)–(d). Specifically, here we present the post-selected 4 × 4 matrices for the CZ gate in the computational basis for nanophotonic, state-of-the-art, and typical imperfections (as in Table 1). The state fidelity is the chance the CZ gate produces the correct output in the computational basis for any given input, given that the output is in the computational basis. It is shown below the corresponding matrix in each case (state fidelity calculation details are provided in the Supplementary Information). Although the performance of the typical circuit (Figure 4(d)) significantly differs from the nanophotonic-only (Figure 4(b)), as expected, the circuit based on state-of-the-art QD parameters (Figure 4(c)) achieves F = 0.9998 , matching the performance of circuits constructed from traditional phase shifters. A similar figure and result is presented in the Supplementary Information for a CNOT gate. As demonstrated by these results, the state fidelity is typically better than the unitary fidelity, reinforcing the viability of QD-based qPICs.

5 Conclusions

Our research demonstrates that qPICs built with reconfigurable QD-phase shifters can perform comparable to those using classical phase shifters. Notably, we show that with state-of-the-art QD parameters, circuits can be scaled up to 10 modes without significant increases in unitary infidelity. This advancement allows QD-based qPICs to efficiently perform operations such as multi-qubit gates, as demonstrated in our study, and to simulate molecular dynamics [54] at cryogenic temperatures. In this respect, QD-based phase shifters join a select class that includes electro-optical [13], [14] and opto-mechanical [15], [16] shifters, but with a much smaller footprint, low operational energy and fast response times. As with current quantum photonic circuits, the performance of emitter-based systems could be further enhanced by incorporating redundancies, where extra MZIs and phase shifters provide better compensation for imperfections [17], [19]–[21].

We recognize that while QDs are the only quantum emitters currently integrated into photonic circuits, alternative technologies based on single organic molecules [55], [56] and defects in diamond [57], [58], silicon [59], or silicon carbide [60], [61] are rapidly maturing. In fact, even with QDs, not all state-of-the-art parameters have been demonstrated using a single chiral quantum photonic platform (c.f. Table 1). To date, however, high quality chiral quantum interfaces have been demonstrated with nanobeams [48], glide-plane waveguides [44] and topological photonics [62], [63]. Recent calculations suggest that, were the emitter to be placed at exactly the correct location within either photonic resonators [64] or waveguides [65], these could act as near-ideal chiral interfaces. This is particularly promising in light of recent developments demonstrating the ability to pre-select specific QDs and integrate them deterministically within a circuit [66]–[71].

Finally, we note that fabricating and managing circuits with numerous emitters remains a topic of inquiry. Recent experiments with QDs have successfully demonstrated the integration of deterministic QD-based single-photon sources with qPICs [72], [73], and independent control of multiple emitters within a circuit [74]–[76], respectively. These advancements open the door to larger-scale implementations where emitters would function both as sources and processing elements. Such circuits would be entirely cryogenic, and thus compatible with deterministic sources and detectors, with phase shifters whose operational speeds are determined by emitter lifetimes, potentially enabling GHz rate operation with mild enhancement [50].


Corresponding author: Nir Rotenberg, Centre for Nanophotonics, Department of Physics, Engineering Physics & Astronomy, Queen’s University, 64 Bader Lane, K7L 3N6, Kingston, Ontario, Canada, E-mail:

Acknowledgment

The authors thank J. Carolan for generating stimulating discussions on quantum photonic integrated circuits, and gratefully acknowledge support by the Natural Sciences and Engineering Research Council of Canada (NSERC), the Canadian Foundation for Innovation (CFI), the Vector Institute, and Queen’s University.

  1. Research funding: Natural Sciences and Engineering Research Council of Canada (https://doi.org/10.13039/501100000038); Canadian Foundation for Innovation (https://doi.org/10.13039/501100000196); Vector Institute (https://doi.org/10.13039/501100019117).

  2. Author contributions: The project was conceived and planned by N.R. with support from B.J.S. A.M. programmed the model of the emitter-based-network, with help from J.E., and performed all simulations. Analysis and interpretation was led by A.M. and N.R. with support from all authors. All authors contributed to the writing and editing of the manuscript. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

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

  4. Data availability: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

[1] C. Sparrow, et al.., “Simulating the vibrational quantum dynamics of molecules using photonics,” Nature, vol. 557, no. 7707, pp. 660–667, 2018. https://doi.org/10.1038/s41586-018-0152-9.Search in Google Scholar PubMed

[2] J. Carolan, et al.., “Universal linear optics,” Science, vol. 349, no. 6249, pp. 711–716, 2015. https://doi.org/10.1126/science.aab3642.Search in Google Scholar PubMed

[3] J. Wang, et al.., “Multidimensional quantum entanglement with large-scale integrated optics,” Science, vol. 360, no. 6386, pp. 285–291, 2018. https://doi.org/10.1126/science.aar7053.Search in Google Scholar PubMed

[4] J. C. Adcock, C. Vigliar, R. Santagati, J. W. Silverstone, and M. G. Thompson, “Programmable four-photon graph states on a silicon chip,” Nat. Commun., vol. 10, no. 1, p. 3528, 2019. https://doi.org/10.1038/s41467-019-11489-y.Search in Google Scholar PubMed PubMed Central

[5] D. Llewellyn, et al.., “Chip-to-chip quantum teleportation and multi-photon entanglement in silicon,” Nat. Phys., vol. 16, no. 2, pp. 148–153, 2020. https://doi.org/10.1038/s41567-019-0727-x.Search in Google Scholar

[6] C. Vigliar, et al.., “Error-protected qubits in a silicon photonic chip,” Nat. Phys., vol. 17, no. 10, pp. 1137–1143, 2021. https://doi.org/10.1038/s41567-021-01333-w.Search in Google Scholar

[7] J. Bao, et al.., “Very-large-scale integrated quantum graph photonics,” Nat. Photonics, vol. 17, no. 7, pp. 573–581, 2023. https://doi.org/10.1038/s41566-023-01187-z.Search in Google Scholar

[8] W. Bogaerts, et al.., “Programmable photonic circuits,” Nature, vol. 586, no. 7828, pp. 207–216, 2020. https://doi.org/10.1038/s41586-020-2764-0.Search in Google Scholar PubMed

[9] H. Hou, P. Xu, Z. Zhou, and H. Su, “Hardware error correction for MZI-based matrix computation,” Micromachines, vol. 14, no. 5, p. 955, 2023. https://doi.org/10.3390/mi14050955.Search in Google Scholar PubMed PubMed Central

[10] W. R. Clements, P. C. Humphreys, B. J. Metcalf, W. S. Kolthammer, and I. A. Walmsley, “Optimal design for universal multiport interferometers,” Optica, vol. 3, no. 12, pp. 1460–1465, 2016. https://doi.org/10.1364/optica.3.001460.Search in Google Scholar

[11] J. Parra, J. Hurtado, A. Griol, and P. Sanchis, “Ultra-low loss hybrid ITO/Si thermo-optic phase shifter with optimized power consumption,” Opt. Express, vol. 28, no. 7, pp. 9393–9404, 2020. https://doi.org/10.1364/oe.386959.Search in Google Scholar PubMed

[12] M. Jacques, A. Samani, E. El-Fiky, D. Patel, Z. Xing, and D. V. Plant, “Optimization of thermo-optic phase-shifter design and mitigation of thermal crosstalk on the SOI platform,” Opt. Express, vol. 27, no. 8, pp. 10456–10471, 2019. https://doi.org/10.1364/oe.27.010456.Search in Google Scholar PubMed

[13] L. Midolo, et al.., “Electro-optic routing of photons from a single quantum dot in photonic integrated circuits,” Opt. Express, vol. 25, no. 26, pp. 33 514–533 526, 2017. https://doi.org/10.1364/oe.25.033514.Search in Google Scholar

[14] G. Sinatkas, T. Christopoulos, O. Tsilipakos, and E. E. Kriezis, “Electro-optic modulation in integrated photonics,” J. Appl. Phys., vol. 130, no. 1, p. 010901, 2021. https://doi.org/10.1063/5.0048712.Search in Google Scholar

[15] F. Beutel, T. Grottke, M. A. Wolff, C. Schuck, and W. H. P. Pernice, “Cryo-compatible opto-mechanical low-voltage phase-modulator integrated with superconducting single-photon detectors,” Opt. Express, vol. 30, no. 17, pp. 30066–30074, 2022. https://doi.org/10.1364/oe.462163.Search in Google Scholar PubMed

[16] C. Papon, et al.., “Nanomechanical single-photon routing,” Optica, vol. 6, no. 4, pp. 524–530, 2019. https://doi.org/10.1364/optica.6.000524.Search in Google Scholar

[17] D. A. B. Miller, “Perfect optics with imperfect components,” Optica, vol. 2, no. 8, pp. 747–750, 2015. https://doi.org/10.1364/optica.2.000747.Search in Google Scholar

[18] J. Mower, N. C. Harris, G. R. Steinbrecher, Y. Lahini, and D. Englund, “High-fidelity quantum state evolution in imperfect photonic integrated circuits,” Phys. Rev. A, vol. 92, no. 3, p. 032322, 2015. https://doi.org/10.1103/physreva.92.032322.Search in Google Scholar

[19] C. M. Wilkes, et al.., “60 dB high-extinction auto-configured Mach–Zehnder interferometer,” Opt. Lett., vol. 41, no. 22, pp. 5318–5321, 2016. https://doi.org/10.1364/ol.41.005318.Search in Google Scholar PubMed

[20] S. Pai, B. Bartlett, O. Solgaard, and D. A. B. Miller, “Matrix optimization on universal unitary photonic devices,” Phys. Rev. Appl., vol. 11, no. 6, p. 064044, 2019. https://doi.org/10.1103/physrevapplied.11.064044.Search in Google Scholar

[21] S. Bandyopadhyay, R. Hamerly, and D. Englund, “Hardware error correction for programmable photonics,” Optica, vol. 8, no. 10, pp. 1247–1255, 2021. https://doi.org/10.1364/optica.424052.Search in Google Scholar

[22] F. Shokraneh, M. Sanadgol Nezami, and O. Liboiron-Ladouceur, “Theoretical and experimental analysis of a 44 recongurable MZI-based linear optical processor,” J. Lightwave Technol., p. 1, 2021, https://doi.org/10.1109/jlt.2021.3068083.Search in Google Scholar

[23] C. Wagenknecht, et al.., “Experimental demonstration of a heralded entanglement source,” Nat. Photonics, vol. 4, no. 8, pp. 549–552, 2010. https://doi.org/10.1038/nphoton.2010.123.Search in Google Scholar

[24] F. Kaneda, K. Garay-Palmett, A. B. U’Ren, and P. G. Kwiat, “Heralded single-photon source utilizing highly nondegenerate, spectrally factorable spontaneous parametric downconversion,” Opt. Express, vol. 24, no. 10, pp. 10733–10747, 2016. https://doi.org/10.1364/oe.24.010733.Search in Google Scholar

[25] C. Couteau, “Spontaneous parametric down-conversion,” Contemp. Phys., vol. 59, no. 3, pp. 291–304, 2018. https://doi.org/10.1080/00107514.2018.1488463.Search in Google Scholar

[26] N. Somaschi, et al.., “Near-optimal single-photon sources in the solid state,” Nat. Photonics, vol. 10, no. 5, pp. 340–345, 2016. https://doi.org/10.1038/nphoton.2016.23.Search in Google Scholar

[27] P. Senellart, G. Solomon, and A. White, “High-performance semiconductor quantum-dot single-photon sources,” Nat. Nanotechnol., vol. 12, no. 11, pp. 1026–1039, 2017. https://doi.org/10.1038/nnano.2017.218.Search in Google Scholar PubMed

[28] R. S. Daveau, et al.., “Efficient fiber-coupled single-photon source based on quantum dots in a photonic-crystal waveguide,” Optica, vol. 4, no. 2, pp. 178–184, 2017. https://doi.org/10.1364/optica.4.000178.Search in Google Scholar

[29] R. Uppu, et al.., “Scalable integrated single-photon source,” Sci. Adv., vol. 6, no. 50, p. eabc8268, 2020. https://doi.org/10.1126/sciadv.abc8268.Search in Google Scholar PubMed PubMed Central

[30] W. H. Pernice, et al.., “High-speed and high-efficiency travelling wave single-photon detectors embedded in nanophotonic circuits,” Nat. Commun., vol. 3, no. 1, p. 1325, 2012. https://doi.org/10.1038/ncomms2307.Search in Google Scholar PubMed PubMed Central

[31] D. Rosenberg, A. Kerman, R. Molnar, and E. Dauler, “High-speed and high-efficiency superconducting nanowire single photon detector array,” Opt. Express, vol. 21, no. 2, pp. 1440–1447, 2013. https://doi.org/10.1364/oe.21.001440.Search in Google Scholar

[32] I. Esmaeil Zadeh, et al.., “Single-photon detectors combining high efficiency, high detection rates, and ultra-high timing resolution,” APL Photonics, vol. 2, no. 11, pp. 111301-1-111301-7, 2017. https://doi.org/10.1063/1.5000001.Search in Google Scholar

[33] Y. Wang, et al.., “Deterministic photon source interfaced with a programmable silicon-nitride integrated circuit,” Npj Quantum Inf., vol. 9, no. 1, pp. 94-1-94-5, 2023. https://doi.org/10.1038/s41534-023-00761-1.Search in Google Scholar

[34] D. Hallett, et al.., “Electrical control of nonlinear quantum optics in a nano-photonic waveguide,” Optica, vol. 5, no. 5, pp. 644–650, 2018. https://doi.org/10.1364/optica.5.000644.Search in Google Scholar

[35] H. Le Jeannic, et al.., “Dynamical photon–photon interaction mediated by a quantum emitter,” Nat. Phys., vol. 18, no. 10, pp. 1191–1195, 2022. https://doi.org/10.1038/s41567-022-01720-x.Search in Google Scholar

[36] T. Unold, K. Mueller, C. Lienau, T. Elsaesser, and A. D. Wieck, “Optical Stark effect in a quantum dot: ultrafast control of single exciton polarizations,” Phys. Rev. Lett., vol. 92, no. 15, p. 157401, 2004. https://doi.org/10.1103/physrevlett.92.157401.Search in Google Scholar PubMed

[37] J. Q. Grim, et al.., “Scalable in operando strain tuning in nanophotonic waveguides enabling three-quantum-dot superradiance,” Nat. Mater., vol. 18, no. 9, pp. 963–969, 2019. https://doi.org/10.1038/s41563-019-0418-0.Search in Google Scholar PubMed

[38] A. S. Sheremet, M. I. Petrov, I. V. Iorsh, A. V. Poshakinskiy, and A. N. Poddubny, “Waveguide quantum electrodynamics: collective radiance and photon-photon correlations,” Rev. Mod. Phys., vol. 95, no. 1, p. 015002, 2023. https://doi.org/10.1103/revmodphys.95.015002.Search in Google Scholar

[39] H. Le Jeannic, et al.., “Experimental reconstruction of the few-photon nonlinear scattering matrix from a single quantum dot in a nanophotonic waveguide,” Phys. Rev. Lett., vol. 126, no. 2, p. 023603, 2021. https://doi.org/10.1103/physrevlett.126.023603.Search in Google Scholar PubMed

[40] M. J. Staunstrup, et al.., “Direct observation of non-linear optical phase shift induced by a single quantum emitter in a waveguide,” arXiv preprint arXiv:2305.06839, 2023.Search in Google Scholar

[41] J.-T. Shen and S. Fan, “Strongly correlated two-photon transport in a one-dimensional waveguide coupled to a two-level system,” Phys. Rev. Lett., vol. 98, no. 15, p. 153003, 2007. https://doi.org/10.1103/physrevlett.98.153003.Search in Google Scholar PubMed

[42] M. J. Powell, et al.., The BOBYQA Algorithm for Bound Constrained Optimization without Derivatives, Cambridge NA Report NA2009/06, vol. 26, Cambridge, University of Cambridge, 2009.Search in Google Scholar

[43] M. Areari, et al.., “Near-unity coupling efficiency of a quantum emitter to a photonic crystal waveguide,” Phys. Rev. Lett., vol. 113, no. 9, p. 093603, 2014. https://doi.org/10.1103/physrevlett.113.093603.Search in Google Scholar

[44] I. Söllner, et al.., “Deterministic photon–emitter coupling in chiral photonic circuits,” Nat. Nanotechnol., vol. 10, no. 9, pp. 775–778, 2015. https://doi.org/10.1038/nnano.2015.159.Search in Google Scholar PubMed

[45] W. Langbein, P. Borri, U. Woggon, V. Stavarache, D. Reuter, and A. Wieck, “Radiatively limited dephasing in inas quantum dots,” Phys. Rev. B, vol. 70, no. 3, p. 033301, 2004. https://doi.org/10.1103/physrevb.70.033301.Search in Google Scholar

[46] C. Matthiesen, A. N. Vamivakas, and M. Atatüre, “Subnatural linewidth single photons from a quantum dot,” Phys. Rev. Lett., vol. 108, no. 9, p. 093602, 2012. https://doi.org/10.1103/physrevlett.108.093602.Search in Google Scholar

[47] H.-S. Nguyen, G. Sallen, C. Voisin, P. Roussignol, C. Diederichs, and G. Cassabois, “Ultra-coherent single photon source,” Appl. Phys. Lett., vol. 99, no. 26, pp. 261904-1-261904-3, 2011. https://doi.org/10.1063/1.3672034.Search in Google Scholar

[48] R. J. Coles, et al.., “Chirality of nanophotonic waveguide with embedded quantum emitter for unidirectional spin transfer,” Nat. Commun., vol. 7, no. 1, p. 11183, 2016. https://doi.org/10.1038/ncomms11183.Search in Google Scholar PubMed PubMed Central

[49] H. Thyrrestrup, et al.., “Quantum optics with near-lifetime-limited quantum-dot transitions in a nanophotonic waveguide,” Nano Lett., vol. 18, no. 3, pp. 1801–1806, 2018. https://doi.org/10.1021/acs.nanolett.7b05016.Search in Google Scholar PubMed

[50] F. T. Pedersen, et al.., “Near transform-limited quantum dot linewidths in a broadband photonic crystal waveguide,” ACS Photonics, vol. 7, no. 9, pp. 2343–2349, 2020. https://doi.org/10.1021/acsphotonics.0c00758.Search in Google Scholar

[51] A. V. Kuhlmann, et al.., “Charge noise and spin noise in a semiconductor quantum device,” Nat. Phys., vol. 9, no. 9, pp. 570–575, 2013. https://doi.org/10.1038/nphys2688.Search in Google Scholar

[52] N. Kiesel, C. Schmid, U. Weber, R. Ursin, and H. Weinfurter, “Linear optics controlled-phase gate made simple,” Phys. Rev. Lett., vol. 95, no. 21, p. 210505, 2005. https://link.aps.org/doi/10.1103/PhysRevLett.95.210505.10.1103/PhysRevLett.95.210505Search in Google Scholar PubMed

[53] J. L. O’Brien, “Optical quantum computing,” Science, vol. 318, no. 5856, pp. 1567–1570, 2007. https://doi.org/10.1126/science.1142892.Search in Google Scholar PubMed

[54] P. J. Ollitrault, A. Miessen, and I. Tavernelli, “Molecular quantum dynamics: a quantum computing perspective,” Acc. Chem. Res., vol. 54, no. 23, pp. 4229–4238, 2021. https://doi.org/10.1021/acs.accounts.1c00514.Search in Google Scholar PubMed

[55] P. Türschmann, et al.., “Chip-based all-optical control of single molecules coherently coupled to a nanoguide,” Nano Lett., vol. 17, no. 8, pp. 4941–4945, 2017. https://doi.org/10.1021/acs.nanolett.7b02033.Search in Google Scholar PubMed

[56] C. Toninelli, et al.., “Single organic molecules for photonic quantum technologies,” Nat. Mater., vol. 20, no. 12, pp. 1615–1628, 2021. https://doi.org/10.1038/s41563-021-00987-4.Search in Google Scholar PubMed

[57] R. E. Evans, et al.., “Photon-mediated interactions between quantum emitters in a diamond nanocavity,” Science, vol. 362, no. 6415, pp. 662–665, 2018. https://doi.org/10.1126/science.aau4691.Search in Google Scholar PubMed

[58] P. K. Shandilya, et al.., “Diamond integrated quantum nanophotonics: spins, photons and phonons,” J. Lightwave Technol., vol. 40, no. 23, pp. 7538–7571, 2022. https://doi.org/10.1109/jlt.2022.3210466.Search in Google Scholar

[59] M. Veldhorst, et al.., “A two-qubit logic gate in silicon,” Nature, vol. 526, no. 7573, pp. 410–414, 2015. https://doi.org/10.1038/nature15263.Search in Google Scholar PubMed

[60] D. M. Lukin, et al.., “4H-silicon-carbide-on-insulator for integrated quantum and nonlinear photonics,” Nat. Photonics, vol. 14, no. 5, pp. 330–334, 2020. https://doi.org/10.1038/s41566-019-0556-6.Search in Google Scholar

[61] D. M. Lukin, M. A. Guidry, and J. Vučković, “Integrated quantum photonics with silicon carbide: challenges and prospects,” PRX Quantum, vol. 1, no. 2, p. 020102, 2020. https://doi.org/10.1103/prxquantum.1.020102.Search in Google Scholar

[62] M. J. Mehrabad, et al.., “Chiral topological photonics with an embedded quantum emitter,” Optica, vol. 7, no. 12, pp. 1690–1696, 2020. https://doi.org/10.1364/optica.393035.Search in Google Scholar

[63] S. Barik, et al.., “A topological quantum optics interface,” Science, vol. 359, no. 6376, pp. 666–668, 2018. https://doi.org/10.1126/science.aaq0327.Search in Google Scholar PubMed

[64] D. Martin-Cano, H. R. Haakh, and N. Rotenberg, “Chiral emission into nanophotonic resonators,” ACS Photonics, vol. 6, no. 4, pp. 961–966, 2019. https://doi.org/10.1021/acsphotonics.8b01555.Search in Google Scholar

[65] N. V. Hauff, H. Le Jeannic, P. Lodahl, S. Hughes, and N. Rotenberg, “Chiral quantum optics in broken-symmetry and topological photonic crystal waveguides,” Phys. Rev. Res., vol. 4, no. 2, p. 023082, 2022. https://doi.org/10.1103/physrevresearch.4.023082.Search in Google Scholar

[66] T. Pregnolato, et al.., “Deterministic positioning of nanophotonic waveguides around single self-assembled quantum dots,” APL Photonics, vol. 5, no. 8, p. 086101, 2020. https://doi.org/10.1063/1.5117888.Search in Google Scholar

[67] S. M. Thon, et al.., “Strong coupling through optical positioning of a quantum dot in a photonic crystal cavity,” Appl. Phys. Lett., vol. 94, no. 11, p. 111115, 2009. https://doi.org/10.1063/1.3103885.Search in Google Scholar

[68] P. Schnauber, et al.., “Deterministic integration of quantum dots into on-chip multimode interference beamsplitters using in situ electron beam lithography,” Nano Lett., vol. 18, no. 4, pp. 2336–2342, 2018, https://doi.org/10.1021/acs.nanolett.7b05218.Search in Google Scholar PubMed

[69] Y.-M. He, et al.., “Deterministic implementation of a bright, on-demand single-photon source with near-unity indistinguishability via quantum dot imaging,” Optica, vol. 4, no. 7, pp. 802–808, 2017. https://doi.org/10.1364/optica.4.000802.Search in Google Scholar PubMed PubMed Central

[70] S. Liu, et al.., “A deterministic quantum dot micropillar single photon source with >65 % extraction efficiency based on fluorescence imaging method,” Sci. Rep., vol. 7, no. 1, p. 13986, 2017. https://doi.org/10.1038/s41598-017-13433-w.Search in Google Scholar PubMed PubMed Central

[71] J. Liu, et al.., “Cryogenic photoluminescence imaging system for nanoscale positioning of single quantum emitters,” Rev. Sci. Instrum., vol. 88, no. 2, p. 023116, 2017. https://doi.org/10.1063/1.4976578.Search in Google Scholar PubMed PubMed Central

[72] A. Chanana, et al.., “Ultra-low loss quantum photonic circuits integrated with single quantum emitters,” Nat. Commun., vol. 13, no. 1, p. 7693, 2022. https://doi.org/10.1038/s41467-022-35332-z.Search in Google Scholar PubMed PubMed Central

[73] J.-H. Kim, S. Aghaeimeibodi, C. J. K. Richardson, R. P. Leavitt, D. Englund, and E. Waks, “Hybrid integration of solid-state quantum emitters on a silicon photonic chip,” Nano Lett., vol. 17, no. 12, pp. 7394–7400, 2017. https://doi.org/10.1021/acs.nanolett.7b03220.Search in Google Scholar PubMed

[74] C. Papon, et al.., “Independent operation of two waveguide-integrated quantum emitters,” Phys. Rev. Appl., vol. 19, no. 6, p. L061003, 2023. https://doi.org/10.1103/physrevapplied.19.l061003.Search in Google Scholar

[75] X.-L. Chu, et al.., “Independent electrical control of two quantum dots coupled through a photonic-crystal waveguide,” Phys. Rev. Lett., vol. 131, no. 3, p. 033606, 2023. https://doi.org/10.1103/physrevlett.131.033606.Search in Google Scholar

[76] S. Li, et al.., “Scalable deterministic integration of two quantum dots into an on-chip quantum circuit,” ACS Photonics, vol. 10, no. 8, pp. 2846–2853, 2023. https://doi.org/10.1021/acsphotonics.3c00547.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2024-0044).


Received: 2024-01-22
Accepted: 2024-04-23
Published Online: 2024-05-09

© 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 23.5.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2024-0044/html
Scroll to top button