Abstract
Solid-state single-photon emitters (SPEs) are attracting significant attention as fundamental components in quantum computing, communication, and sensing. Low-dimensional materials-based SPEs (LD-SPEs) have drawn particular interest due to their high photon extraction efficiency, ease of integration with photonic circuits, and strong coupling with external fields. The accessible surfaces of LD materials allow for deterministic control over quantum light emission, while enhanced quantum confinement and light–matter interactions improve photon emissive properties. This perspective examines recent progress in LD-SPEs across four key materials: zero-dimensional (0D) semiconductor quantum dots, one-dimensional (1D) nanotubes, two-dimensional (2D) materials, including hexagonal boron nitride (hBN) and transition metal dichalcogenides (TMDCs). We explore their structural and photophysical properties, along with techniques such as spectral tuning and cavity coupling, which enhance SPE performance. Finally, we address future challenges and suggest strategies for optimizing LD-SPEs for practical quantum applications.
1 Introduction
Photonic quantum technologies that harness the quantum properties of light (photons) to process quantum information have drawn increasing interest over the past decade. Single-photon sources [1], [2], [3], [4] – isolated quantum systems designed to emit precisely one photon per excitation cycle – have garnered widespread research interest. These flying qubits play a crucial role in encoding, transmitting, and transducing quantum information, forming the foundation of many approaches to quantum computing [5], metrology [6], sensing [7], and secure communication [8].
Solid-state SPEs stand out by combining the application-specific tailoring of optical properties, such as wavelength, intensity, and polarization, while maintaining the scalability of solid-state systems [9], [10], [11], offering on-demand emission that is not available in other quantum light sources, such as those utilizing nonlinear optical techniques like spontaneous parametric down-conversion (SPDC) [12], [13], [14] and spontaneous four-wave mixing (SFWM) [15]. SPEs have been reported in various material platforms such as quantum dots (QDs) [16], [17], rare earth ions [18], [19], [20], and defect centers [21] (e.g., nitrogen-vacancy (NV) centers in diamond [22], [23], [24]), offering high brightness, purity, and indistinguishability. Additionally, some SPEs exhibit optically addressable spin states, suitable for quantum sensing, transduction, and nuclear magnetic resonance spectroscopy [25].
SPEs in 3D bulk materials such as silicon [26], [27] and diamond suffer from low (<10 %) intrinsic photon extraction efficiency due to internal reflection, although extensive efforts have gone into patterning 3D photonic platforms with improved extraction efficiency [28]. In addition, integrating 3D SPEs into photonic structures [29] is usually not straightforward and requires demanding fabrication process [30], [31], [32]. LD-materials, on the other hand, offer high photon extraction efficiencies and comparably easy integration with photonic interfaces [33], [34], [35]. Their accessible surface and enhanced surface-volume ratio allows for deterministic SPE creation [36], [37] and efficient tuning of emission properties [38], [39]. Additionally, low-dimensional materials could host exotic physical phenomena, such as valley degrees of freedom [40] and exciton–magnon coupling [41], [42], which pave the way for coherent control of photonic qubits via external fields [43], [44].
This review examines the advancement in LD-SPEs developed from four key material systems: 0D semiconductor QDs, 1D nanotubes, 2D hBN, and 2D TMDCs (Figure 1a). We begin by discussing the mechanism behind single-photon emission (Section 2) and the characterization of SPEs (Section 3). This is followed by an in-depth review of the structures, photophysical properties, state-of-the-art performance, applications, and challenges associated with LD-SPEs (Sections 4–7). In Sections 8 and 9, we explore techniques like spectral tuning and cavity coupling, which are widely utilized to enhance SPE performance. Finally, in Section 10, we summarize the key findings in LD-SPEs and provide perspectives on future research directions.

Wavelength centric overview of low-dimensional materials based SPEs. (a) Electromagnetic spectrum showing spectral ranges and applications for SPEs across ultraviolet, visible, near-infrared, and telecommunication wavelengths, from left to right. The top portion highlights applications, while the bottom shows schematic illustrations of SPE materials: quantum dots, nanotubes, hBN, and TMDCs. Colored polygons indicate the spectral ranges covered by each material. (b)–(e) Different mechanisms for generating single photons. (b) Spontaneous decay of excited states, where an excitation laser promotes an electron to the excited state, and single photons are emitted during relaxation. (c) Spontaneous decay of localized excitons, where the laser creates excitons, which recombine to emit single photons. (d) Ambipolar emission in electroluminescent devices, where electron–hole recombination between the source and drain generates photons. (e) Unipolar emission mechanism via impact excitation in electroluminescent devices, where high-energy carriers excite electrons to emit photons during relaxation.
2 Single-photon emitters: realization and excitation
Over the past four decades, SPEs have been successfully demonstrated in a wide variety of materials. These include bulk materials such as diamond [45], [46], silicon carbide (SiC) [47], silicon nitride (SiN) [48], aluminum nitride (AlN) [49], [50], and gallium nitride (GaN) [51], [52]; two-dimensional materials like hBN [53], [54] and tungsten diselenide (WSe2) [55], [56], [57], [58]; one-dimensional materials like single-walled carbon nanotubes (SWCNTs) [59], [60] and nanowires [61]; and zero-dimensional materials such as colloidal quantum dots [62], graphene quantum dots [33], and epitaxial quantum dots [63], [64]. Single-photon emission in these materials typically originates from two mechanisms: (1) color centers and (2) excitons confined within nanostructures or potential wells [34]. In both cases, a well-defined two-level energy structure is required to avoid unwanted emission from multiple optical transitions [65].
Color centers, usually observed in wide-bandgap systems such as diamond [31] and hBN [66], are defects in the crystal lattice where an atom is either missing or replaced by a different atom (i.e., vacancies or impurities). They disrupt the periodic potential within the solid, creating localized electronic states within the material’s bandgap (Figure 1b). These defects, which are usually found in nanocrystals, exfoliated layers, powders, and bulk materials, can be created through thermal annealing [67], femtosecond laser direct writing [68], electron or ion beam irradiation [69], [70], mechanical indentation [71], and chemical/plasma treatment [72], [73]. However, these techniques face physical limitations that hinder the consistent creation of identical emitters and frequently lead to the formation of unintended defects, thereby impeding the production of uniform SPEs.
Confined excitons, usually observed in LD materials such QDs [17], SWCNTs [74], and TMDCs [75], are bound states of excitons that are spatially localized (Figure 1c). These regions, often caused by defects, strain, or size confinement, disrupt the uniform electronic potential, resulting in quantized energy states [76] that can be engineered through mechanical exfoliation [56], chemical vapor deposition (CVD) [77], chemical functionalization [78], and strain engineering [79]. However, these methods may result in spatial and spectral inhomogeneities due to intrinsic solid-state environment noise and fluctuations in fabrication conditions, complicating the generation of reproducible and indistinguishable SPEs. Thus, new techniques that can inherently ensure the production of identical SPEs are highly desired.
Excitation involves pumping electrons from the ground state to an excited state. We categorize the methods for exciting LD-SPEs based on various perspectives, including optical versus electrical excitation, continuous wave (CW) versus pulsed excitation, and resonant versus nonresonant excitation, as detailed below.
2.1 Optical versus electrical excitation
Optical excitation is widely used for SPEs and can be controlled by adjusting light intensity and wavelength. Jungwirth et al. [80] employed optical excitation to survey the temperature-dependent emission of point defects in multilayer hBN. They identified single-photon emission from individual SPEs by leveraging the spectral selectivity and high spatial resolution of optical excitation. Electrical excitation, alternatively, uses an electric current or voltage to stimulate SPEs, enabling single-photon emission through ambipolar emission (Figure 1d), where both electrons and holes recombine, or unipolar emission (Figure 1e), where a single type of carrier recombines with existing states. Clark et al. [81] demonstrated electrically driven SPEs in WSe2 and observed their electroluminescence (EL) intensities at cryogenic temperature, paving the way for on-chip SPEs in TMDCs. Figure 2a and b illustrate the different mechanisms of optical and electrical excitations.
![Figure 2:
Different excitation schemes in 2D SPEs. (a) Schematic of the optical excitation mechanism in WSe2-based SPEs. An excitation laser excites electrons from the ground state to the excited state, forming excitons in the WSe2 layer positioned on a nanopillar. The excitons drift, and the electron–hole recombination results in the emission of single photons via spontaneous emission. (b) A vertical electrical excitation device comprised of WSe2 sandwiched between few-layer boron nitride and metal contacts. The device structure includes graphene as a conductive layer, with electrical excitation applied through metal contacts to generate excitons within the WSe2 layer. (c) Resonant excitation. The excitation laser and emitted single photons are spectrally filtered based on their polarization. (d) Resonant excitation, spectrally filtering is applied to separate the propagation direction of the excitation laser and the emitted photons. (e) Nonresonant excitation. The excitation laser (ω′) and emitted single photons (ω) have different wavelengths. Spectral filtering is applied to isolate the single photons from the excitation laser based on their distinct wavelengths. (f) A kind of quasi-resonant excitation. Two near-resonant excitation laser pulses (ω
+ and ω
−) are used to predictably manipulate the emitter into its excited state, leading to the emission of single photons (ω). Adapted with permission from: (b), ref. [81] Copyright 2016, American Chemical Society.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_002.jpg)
Different excitation schemes in 2D SPEs. (a) Schematic of the optical excitation mechanism in WSe2-based SPEs. An excitation laser excites electrons from the ground state to the excited state, forming excitons in the WSe2 layer positioned on a nanopillar. The excitons drift, and the electron–hole recombination results in the emission of single photons via spontaneous emission. (b) A vertical electrical excitation device comprised of WSe2 sandwiched between few-layer boron nitride and metal contacts. The device structure includes graphene as a conductive layer, with electrical excitation applied through metal contacts to generate excitons within the WSe2 layer. (c) Resonant excitation. The excitation laser and emitted single photons are spectrally filtered based on their polarization. (d) Resonant excitation, spectrally filtering is applied to separate the propagation direction of the excitation laser and the emitted photons. (e) Nonresonant excitation. The excitation laser (ω′) and emitted single photons (ω) have different wavelengths. Spectral filtering is applied to isolate the single photons from the excitation laser based on their distinct wavelengths. (f) A kind of quasi-resonant excitation. Two near-resonant excitation laser pulses (ω + and ω −) are used to predictably manipulate the emitter into its excited state, leading to the emission of single photons (ω). Adapted with permission from: (b), ref. [81] Copyright 2016, American Chemical Society.
Cathodoluminescence (CL) microscopies have separately emerged as a powerful resource for near-field imaging of LD-SPEs [82], [83], [84], [85], [86], [87], [88], [89]. CL microscopy leverages converged electron-beams in scanning (transmission) electron microscopes combined with far-field optical detection of photons generated by the material after electron-beam excitation in order to probe LD-SPEs with true nanoscale resolution well below the optical diffraction limit. While CL microscopies have been used to probe the photon statistics of SPEs, much of the literature exhibits photon bunching instead of photon antibunching under electron-beam excitation because the high-energy electron-beam can easily excite unwanted electronic transitions, resulting in the concurrent emission of photons from many excited states at once [83], [84], [90], [91]. However, appropriate choices of electron-beam current do allow for nanoscale probes of photon antibunching [87], [88], [92]. CL microscopy is a particularly appealing tool for LD-SPEs because of the potential for in situ patterning and modification of SPEs and their environments. For instance, CL microscopy has been used for in situ monitoring of e-beam patterned defects in hBN [86], [89], and e-beam induced etching in water vapor environments has been used to pattern nanoscale diamond cavities with in situ CL feedback [32]. Ultimately, the ability to image and pattern LD-SPEs at these length scales may lead to the development of new integrated quantum photonic systems with optimized cavity interactions designed to achieve ideal SPE properties.
2.2 Continuous wave (CW) versus pulsed excitation
CW excitation, which uses a steady energy source such as a laser or electric current, enables sustained single-photon emission but can reduce purity at higher excitation power. In contrast, pulsed excitation employs short, intense bursts of energy, offering advantages in time-resolved studies. Li et al. [93] utilized both CW and pulsed excitation to investigate the emission properties of hBN SPEs, reporting emission rates of 44 MHz under CW and 10 MHz under 80 MHz pulsed excitation. Notably, the purity of the SPEs under CW excitation reduced significantly as the excitation power increased, whereas purity under pulsed excitation remained high even at saturation power. Pulsed excitation also allows for time-gated correlation measurement that can improve measured single-photon purity [94].
2.3 Resonant versus nonresonant excitation
Resonant excitation utilizes an excitation energy that exactly matches the optical bandgap [4]. This method allows for near-deterministic excitation of the emitter, minimizing excess energy that could cause unwanted emission or phonon-induced spectral broadening. However, resonant excitation requires complicated excitation or detection schemes, such as polarization filtering [95], [96], phonon sideband (PSB) detection [97], and non-normal excitation [98], to separate the emitted photons from the excitation laser (Figure 2c and d). Wang et al. [95] employed polarization filtering to collect single photons from SPEs in InGaAs quantum dot. They reduced the polarization loss to 3.8 % instead of 50 % by coupling to polarization-selective Purcell microcavities. Extra filtering requirements sometimes limit the polarization direction and intensity. Nonresonant excitation, on the other hand, uses energy higher than the bandgap (Figure 2e). While simpler to implement, this method often results in lower-quality single photons and degrades the indistinguishability of the SPEs. Alternatively, “quasi-resonant” excitation emerged to address the above issues [99]. In this approach, two near-resonant pulses are used to predictably excite the SPEs [100], [101] (Figure 2f), allowing for natural decay or stimulated decay with a second pulse [102]. Jayakumar et al. [99] utilized a two-photon excitation scheme on InAs/GaAs QDs embedded into a microcavity. It allows for the deterministic generation of photon pairs, making the scheme suitable for generating time-bin entanglement.
3 Characterization of single-photon emission
SPEs are characterized by BPI values (Brightness, Purity, and Indistinguishability). Brightness ( B ) is quantified by PL intensity, which represents the number of collected photons per second. B is proportional to the excitation rate, quantum yield (QY), and collection efficiency. The QY quantifies the efficiency of photon emission in response to excitation, which is calculated from the number of emitted single photons at saturation normalized to the laser repetition rate. The measured value of B can vary depending on the measurement location: at the first collection element (B 1), coupled inside a single mode fiber or optical path (B 2), and at the detector (B 3) [11]. Figure 3a shows a simple illustration of a fiber-coupled measurement scheme. B is wavelength dependent (Figure 3b) and is a function of excitation power (inset Figure 3b). The inset of Figure 3b shows the PL intensity of a defect in hBN nanoflakes before and after coupled to a metallo-dielectric antenna [93], which saturates at high power due to the finite availability of excited states and competition with nonradiative recombination processes, such as Auger recombination.
![Figure 3:
Criteria of SPEs, brightness, purity, and indistinguishability. (a) Experiment set up used to measure brightness
B
of an SPE under pulsed excitation conditions. The emitted photons are collected through a single-mode fiber and detected by a detector (D), with the brightness measured at different points (B
1, B
2, B
3) depending on the collection efficiency. The logic count system is used to record photon detection events. (b) Brightness measurement results showing the integrated PL intensity of an SPE at two excitation powers, 600 μW (blue) and 1,200 μW (red), with emission peaks around 515 nm. The inset displays the pump-power-dependent PL intensity of a hBN SPE before and after coupled to a metallo-dielectric antenna. (c) HBT experiment setup used to measure purity of an SPE under pulsed excitation conditions. Emitted photons from the source are split by a beam splitter and directed to two detectors, D
1 and D
2, with the photon coincidences recorded by the logic count system. (d) Single-photon purity measured under continuous wave excitation. The data (points) and the fit (blue line) yield
g
2
0
=
${g}^{\left(2\right)}\left(0\right)=$
0.33 ± 0.02. The inset shows the single-photon purity measured under pulsed excitation, demonstrating clear photon antibunching behavior with well-separated peaks in the coincidence counts. (e) HOM experiment setup used to measure indistinguishability of an SPE under pulsed excitation conditions. Two consecutive photons, separated by a time delay τ, are sent through a beam splitter, and their interference is measured at detectors D
1 and D
2. (f) Indistinguishability measurement results using the HOM effect, showing the
g
HOM
(
2
)
0
${g}_{\text{HOM}}^{(2)}\left(0\right)$
as a function of time delay for two interferometer settings: 0 ps (dark green) and 5 ps (orange). The reduction in photon correlations at zero-time delay demonstrates two-photon interference. Reproduced with permission from: (a), (c), (e), ref. [11], John Wiley and Sons; (f), ref. [103], Springer Nature.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_003.jpg)
Criteria of SPEs, brightness, purity, and indistinguishability. (a) Experiment set up used to measure brightness
B
of an SPE under pulsed excitation conditions. The emitted photons are collected through a single-mode fiber and detected by a detector (D), with the brightness measured at different points (B
1, B
2, B
3) depending on the collection efficiency. The logic count system is used to record photon detection events. (b) Brightness measurement results showing the integrated PL intensity of an SPE at two excitation powers, 600 μW (blue) and 1,200 μW (red), with emission peaks around 515 nm. The inset displays the pump-power-dependent PL intensity of a hBN SPE before and after coupled to a metallo-dielectric antenna. (c) HBT experiment setup used to measure purity of an SPE under pulsed excitation conditions. Emitted photons from the source are split by a beam splitter and directed to two detectors, D
1 and D
2, with the photon coincidences recorded by the logic count system. (d) Single-photon purity measured under continuous wave excitation. The data (points) and the fit (blue line) yield
Single-photon purity (
P
) is quantified by the second order correlation function
where
Figure 3d shows
Indistinguishability (
I
) is quantified by V
HOM, the visibility in the Hong–Ou–Mandel (HOM) experiment [108], [109]. Figure 3e illustrates a HOM experiment, where single photons are split into two paths by the first beam splitter with a delay time τ introduced between the paths. At the second beam splitter, quantum interference occurs, and if the photons are indistinguishable, they will exit together, leading to reduced coincidence counts for τ = 0. By polarization filtering or adjusting τ, we can obtain V
HOM [109], given by
The BPI values influence the design rules for SPEs. Table 1 demonstrates the performance of SPEs in quantum applications, where source efficiency refers to the probability of collecting a photon in each excitation pulse (proportional to B ). In QKD protocols, such as BB84 [128], both P and B contribute to improving the secure key rate. In other applications, such as Greenberger–Horne–Zeilinger (GHZ) state generation [125], [126], [127], all BPI values are crucial for enhancing fidelities and overall efficiencies. Other factors also affect the design rules of SPEs. For example, the emission wavelength determines the transmission properties in various media [17] (Figure 1a). SPEs with telecommunication band emission minimize the transmission losses in optical fibers, enable long-distance QKD [113], and are promising for quantum internet construction. Emission in the ultraviolet range is ideal for free-space transmission. SPEs with electrical excitation capability are ideal for integration into Complementary Metal-Oxide-Semiconductor (CMOS) circuits [34], [129]. SPEs that can emit photons at room temperature or higher [130] reduce the energy cost associated with cooling systems. SPEs that emit in a Gaussian mode facilitate seamless waveguide coupling [131].
Performance of LD-SPEs in various quantum applications.
Purity g (2)(0) | Indistinguishability | Source efficiency | |
---|---|---|---|
Quantum key distribution [113], [114], [115] | ∼0.5 % | – | ∼5 % |
Boson sampling [116], [117], [118], [119] | ∼2 % | ∼95 % | ∼55 % |
Quantum teleportation [120], [121] | ∼18 % | ∼65 % | ∼15 % |
Generate cluster state [122], [123], [124] | ∼5 % | ∼95 % | ∼18 % |
Generate GHZ state [125], [126], [127] | ∼2 % | ∼95 % | ∼30 % |
In Sections 4–7, we review the progress of LD-SPEs categorized by their materials and emission wavelengths (Figure 1a). We discuss (1) semiconductor QD SPEs with emission wavelengths spanning approximately from 280 nm to 1,550 nm, (2) nanotube SPEs from 570 nm to 2,000 nm, (3) hBN SPEs from 300 nm to 850 nm, and (4) TMDCs from 600 nm to 1,550 nm. We focus in particular on SPEs capable of emitting single photons at telecommunication wavelengths, based on materials such as InAs/InP QDs, SWCNTs, and MoTe2.
4 Semiconductor quantum dots
LD-SPEs have been demonstrated in CQDs [132], graphene QDs (GQDs) [33], [133], and epitaxially grown QDs (EQDs) [17]. Single-photon emission in QDs originates from excitons formed within discrete energy levels due to quantum confinement.
CQDs are semiconductor nanocrystals with core sizes typically ranging from 2 to 10 nm, synthesized in a colloidal solution [132]. CQD SPEs are attractive due to their flexibility in synthesis [134], ease of integration, and ability to operate at room temperature [135]. CQD SPEs offer tunable emission wavelengths, which can be controlled by adjusting their size, morphology, and structure [132], [136]. Krishnamurthy et al. [137] demonstrated PbS/CdS SPEs with tunable emission wavelengths covering the telecom S band (1,460–1,530 nm) and O band (1,260–1,360 nm) at room temperature. The tunability was achieved by adjusting the core size and shell thickness of PbS/CdS CQDs. Chandrasekaran et al. [135] demonstrated near-blinking free, high purity (
As a specific type of CQD, halide perovskites QDs (PQDs) exhibit room-temperature single-photon emission, near-unity QY, and high photostability. PQDs are synthesized through methods [138] such as hot injection [139], [140], ligand-assisted precipitation [141], [142], and ultrasonic synthesis [143], [144]. Liu et al. [145] demonstrated 100 % QY in CsPbI3 PQDs at room temperature, employing a synthetic protocol involving the introduction of an organometallic compound, trioctylphosphine-PbI2, as the reactive precursor. Tang et al. [146] demonstrated CsPbBr3/CdS core/shell PQDs with nonblinking PL and a high QY of 90 %, attributed to the reduction of electronic traps within the stable core/shell structure. Utzat et al. [112] showed that CsPbBr3 SPEs exhibit fast emission lifetimes of 210–280 ps (Figure 4a) with a large T
2/2T
1 ratio (∼0.2) and stable emission over several minutes at cryogenic temperatures at 520 nm (2.38 eV). Zhu et al. [149] reported CsPbI3 SPEs with 98 %
P
(
![Figure 4:
Single-photon emitters based on quantum dots. (a) The PL decay of a single PQD. The emission exhibits an initial fast decay (∼210–280 ps), followed by a slower mono-exponential decay. (b) Single-photon purity of GQD SPEs under nonresonant excitation at room temperature, yielding g
(2)(0) equals to 0.05 ± 0.05. (c) Summary plot showing emission wavelengths and operational temperatures of various EQD SPEs. (d) PL spectra of EQD SPEs measured at different temperatures: 3.9 K (left), 150 K (middle), and 300 K (right). At 3.9 K, the inset shows a power dependence plot with a fitted slope. At 150 K, the spectrum displays an acoustic phonon sideband. At 300 K, the PL peak exhibits a shift of 87 meV. (e) Indistinguishability M of InGaAs SPEs as a function of excitation power, measured at 4.2 K. The indistinguishability reaches a maximum value of 0.9956. Error bars are based on Poissonian statistics from detected events. (f) The purity of an InAs/InP QD in an optical horn at 8 K under quasi-resonant excitation with a
g
2
0
${g}^{\left(2\right)}\left(0\right)$
value of 4.4 × 10−4 and a background correction value of 2.2 × 10−4. Adapted with permission from: (a), ref. [112], The American Association for the Advancement of Science; (b), ref. [33], Springer Nature; (c), ref. [17], AIP Publishing; (d), ref. [16], Copyright 2014, American Chemical Society; (e), ref. [147], Springer Nature; (f), ref. [148], AIP Publishing.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_004.jpg)
Single-photon emitters based on quantum dots. (a) The PL decay of a single PQD. The emission exhibits an initial fast decay (∼210–280 ps), followed by a slower mono-exponential decay. (b) Single-photon purity of GQD SPEs under nonresonant excitation at room temperature, yielding g
(2)(0) equals to 0.05 ± 0.05. (c) Summary plot showing emission wavelengths and operational temperatures of various EQD SPEs. (d) PL spectra of EQD SPEs measured at different temperatures: 3.9 K (left), 150 K (middle), and 300 K (right). At 3.9 K, the inset shows a power dependence plot with a fitted slope. At 150 K, the spectrum displays an acoustic phonon sideband. At 300 K, the PL peak exhibits a shift of 87 meV. (e) Indistinguishability M of InGaAs SPEs as a function of excitation power, measured at 4.2 K. The indistinguishability reaches a maximum value of 0.9956. Error bars are based on Poissonian statistics from detected events. (f) The purity of an InAs/InP QD in an optical horn at 8 K under quasi-resonant excitation with a
CQD SPEs face several bottlenecks waiting to be addressed: strong Auger recombination reduces the PL intensity; intense multiexciton emission restricts HOM experiments to cryogenic temperatures; blinking leads to poor photostability at ambient conditions; and there are no reports to date of telecommunications band operation.
Graphene QDs (GQDs) are atomically thin fragments of graphene, typically consisting of 1 or 2 layers with lateral sizes below 10 nm [153]. Compared to graphene, GQDs exhibit desirable properties for SPEs, such as a bandgap opened as a result of quantum confinement and tunable physical properties enabled by geometry engineering and chemical functionalization. GQD SPEs are synthesized by bottom-up methods, such as molecular fusion, allowing control over size, morphology, doping, functionalization, and synthesis techniques. Zhao et al. [33] demonstrated GQD SPEs at room temperature with high
B
,
P
(
EQDs are synthesized through layer-by-layer techniques such as Molecular Beam Epitaxy (MBE) and Metal–Organic Chemical Vapor Deposition (MOCVD) [155]. SPEs with tailored nanostructures can be formed by tuning growth conditions, strain, and utilizing prepatterned substrates along with postgrowth techniques like etching and lithography [156], [157]. EQD SPEs cover a broad emission wavelength range, from the ultraviolet (below 280 nm) to the telecom band (around 1,550 nm) (Figure 4c). Emission in the UV range is achieved using III-nitrides, particularly GaN/AlGaN EQDs. Holmes et al. [16] demonstrated that GaN/AlGaN SPEs maintain high
P
(
InGaN, InGaN/GaN, InP, and various II–VI EQDs [159], [160], such as CdSe/ZnSe, have been used to realize SPEs with emission in the visible region. Fedorych et al. [160] demonstrated CdSe/ZnSSe/MgS EQD SPEs with
III-arsenide materials, such as InAs/GaAs EQDs, have been used to achieve single-photon emission in the near-infrared (NIR) region with near-unity
I
. He et al. [165] demonstrated InAs/GaAs SPEs operating at a wavelength of 940 nm with
The telecommunication band is primarily covered by InAs/InP and InAs/GaAs EQDs due to their narrow bandgaps (0.2 eV–1.2 eV). The emission wavelengths of InAs/InP SPEs can be tuned to 1,550 nm (telecom C-band) [167]. Takemoto et al. [113] demonstrated InAs/InP SPEs with high purities (
EQD SPEs demonstrated room-temperature emission in the UV or visible range, with tunable emission wavelengths controlled by adjusting the material’s stoichiometry during growth [17]. Thanks to the maturity of EQD growth techniques, these SPEs exhibit promising scalability for quantum applications such as boson sampling [116], [117], [118], linear cluster state generation [122], [123], [124], QKD [113], [114], quantum logic gate operation [171], and quantum teleportation [120], [121]. While promising, EQD SPEs face several challenges: HOM measurements are only favorable at cryogenic temperatures due to strong dephasing rates at room temperature; it remains challenging to achieve usable photon indistinguishability for photons emitted from distinct EQD SPEs; and the difficulty in achieving electrical excitation limits the potential for integration into on-chip devices.
5 Nanotubes
1D nanotubes SPEs are primarily realized using two materials: boron nitride nanotubes (BNNTs) and SWCNTs. BNNTs, structurally similar to rolled boron nitride sheets, are known for their wide bandgap and high thermal and chemical stability [172]. BNNT SPEs exhibit room-temperature single-photon emission in the 570–610 nm range [173], [174], [175]. Chejanovsky et al. [173] reported SPEs in BNNT with
SWCNTs consist of covalently bonded carbon atoms arranged in an ordered tubular structure, with their diameter and roll-up angle defined by the chiral index ( n , m ) where n and m specify the wrapping direction of the graphene lattice. The chiral index also defines the emission wavelength for intrinsic SWCNTs. SWCNTs stand out as candidates for SPEs due to their structure-specific NIR PL [176] (Figure 5a). Single-photon emission from SWCNTs originates from excitons confined in potential wells created through noncovalent or covalent functionalization.
![Figure 5:
Single-photon emitters based on SWCNTs. (a) PL spectrum of a single SWCNT at room temperature (black line) and at 10 K (blue line) [176]. Inset: a polymer wrapped nanotube leading to reduced spectral diffusion and blinking [177] (left); PL polarization diagram (right). (b) Purity of electrically excited SWCNT SPEs at 1.6 K, yielding g
(2)(0) equals to 0.49 at 0.08 μW excitation power. (c) PL spectrum and purity for SPEs in (10, 3) SWCNTs functionalized with OCH3-Dz. The PL spectrum shows an emission peak at 1.55 μm corresponding to the E
11
∗ transition. The
g
2
0
${g}^{\left(2\right)}\left(0\right)$
equals to 0.01 measured at 220 K demonstrates high purity of the single-photon emission. (d) Schematic of nanotube defect (NTD) SPEs operating at room temperature and coupled to a tunable fiber cavity. The fiber cavity setup allows precise control of the cavity length (L
C) and enhances emission properties of the NTD SPEs. (e) The HOM second-order correlation function of an NTD SPE. HOM autocorrelation function of an NTD, measured in a copolarized interferometer configuration with interferometer delays of 0 ps (dark green) and 5 ps (orange). The zero-interferometer delay corresponds to a delay equal to the separation of one excitation pulse. The visibility is then measured to be v = 0.65 ± 0.24 at room temperature. Adapted with permission from: (a), ref. [176], Springer Nature; (b), ref. [178], Springer Nature; (c), ref. [74], Springer Nature; (d), (e), ref. [103], Springer Nature.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_005.jpg)
Single-photon emitters based on SWCNTs. (a) PL spectrum of a single SWCNT at room temperature (black line) and at 10 K (blue line) [176]. Inset: a polymer wrapped nanotube leading to reduced spectral diffusion and blinking [177] (left); PL polarization diagram (right). (b) Purity of electrically excited SWCNT SPEs at 1.6 K, yielding g
(2)(0) equals to 0.49 at 0.08 μW excitation power. (c) PL spectrum and purity for SPEs in (10, 3) SWCNTs functionalized with OCH3-Dz. The PL spectrum shows an emission peak at 1.55 μm corresponding to the E
11
∗ transition. The
Noncovalent functionalization creates localized potential wells through unintentional molecular adsorption or local inhomogeneities at the interface with the surrounding matrix or substrate [179], allowing control of exciton diffusion and inducing strong photon antibunching. This approach leverages the sensitivity of SWCNTs to their dielectric environment while preserving their excellent optical characteristics [180]. Högele et al. [181] demonstrated strong photon antibunching in CoMoCat SWCNTs encapsulated in sodium dodecylbenzenesulfonate, achieving
Covalent functionalization has enabled a variety of approaches for creating localized excitons through methods such as oxygen doping, diazonium salts, DNA, or photoexcited aromatics-based functionalization. Ma et al. [59] demonstrated solitary oxygen dopant SWCNT SPEs with
Aryl sp3 defects created through diazonium-based reactions have emerged as promising candidates for SPEs, offering stable, shot-noise limited emission. These defects are synthetically tunable, allowing for enhanced trapping potentials and red-shifted emission, particularly in large-diameter tubes emitting at telecom wavelengths. He et al. [74] demonstrated stable SPEs with
6 Few-layer hexagonal boron nitride
Two-dimensional hBN is a wide-bandgap insulator (E g ∼ 5.97 eV) with a graphene-like honeycomb lattice of alternating boron and nitrogen atoms. The wide bandgap makes hBN SPEs robust against thermal fluctuations, enabling stable single-photon emission at room-temperature. Few-layer hBN SPEs show promising characteristics such as high BPI values and spin-photon interfaces. hBN SPEs have been successfully integrated with photonic circuits in initial demonstrations of scalable quantum photonic technologies [38], [39]. Single-photon emission in hBN arises from trapped excitons at defect sites, including nitrogen vacancies (VN), boron vacancies (VB), antisite carbon vacancies (VNCB), and antisite nitrogen vacancies (VNNB) [184] (Figure 6a), which can be introduced by annealing [67], [187], electron beam [185], [188], [189] or ion beam [69], [189] irradiation, nanopillars [190], [191], plasma processing [72], [73], or femtosecond pulses [68], [192]. Theoretical studies [193], [194], [195], [196] have proposed new type of defects, such as C2CN and C2CB carbon clusters [197], [198], as candidates for SPEs in hBN. hBN SPEs exhibit zero phonon lines (ZPLs) across the NIR-visible range (∼560–780 nm) [67], [80] and the UV range (∼300 nm) [87], [199].
![Figure 6:
Structure, defects, and performance of hBN SPEs. (a) Schematic of a hBN lattice structure highlighting various types of defects it can host. The lattice is composed of boron (B, yellow) and nitrogen (N, blue) atoms. Common defects include nitrogen vacancies (VN), boron vacancies (VB), oxygen substituting nitrogen or boron (ON, OB), carbon substituting nitrogen or boron (CN, CB), and complex vacancies with multiple atoms missing or substituted (e.g., VB3H, VB2O). (b) Schematic demonstrating the separation of excitation and emission light using a dichroic mirror (DM) during the optical characterization of hBN. The excitation (Exc.) light is directed onto the sample via an objective (Obj.), while the emitted (Emi.) light is reflected by the DM and collected for analysis. (c) Low-temperature spectra of the eight hBN SPEs, labeled 1 through 8. The ZPL for all emitters is reproducible, centered around 436 ± 0.7 nm. (d) Two photon interference between successively emitted photons from the same source with a delay of 12.5 ns, yielding a V
HOM of 0.56 ± 0.11. (e) PL intensity versus excitation power for hBN SPEs, comparing uncoupled (blue) and coupled (red) configurations using a metallo-dielectric antenna setup. The coupled system achieves a near-unity photon collection efficiency of 98 %, compared to 13 % for the uncoupled case. The inset shows the emitter intensity over time, demonstrating excellent temporal stability without blinking. (f) High purity hBN SPEs with
g
2
0
${g}^{\left(2\right)}\left(0\right)$
equals to 0.0064 under pulsed excitation. Adapted with permission from: (a), ref. [46], AIP Publishing; (b), ref. [67], Copyright 2016, American Chemical Society; (c), ref. [185], Springer Nature; (d), ref. [186], American Physical Society; (e), ref. [93], Copyright 2019, American Chemical Society; (f), ref. [107], American Physical Society.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_006.jpg)
Structure, defects, and performance of hBN SPEs. (a) Schematic of a hBN lattice structure highlighting various types of defects it can host. The lattice is composed of boron (B, yellow) and nitrogen (N, blue) atoms. Common defects include nitrogen vacancies (VN), boron vacancies (VB), oxygen substituting nitrogen or boron (ON, OB), carbon substituting nitrogen or boron (CN, CB), and complex vacancies with multiple atoms missing or substituted (e.g., VB3H, VB2O). (b) Schematic demonstrating the separation of excitation and emission light using a dichroic mirror (DM) during the optical characterization of hBN. The excitation (Exc.) light is directed onto the sample via an objective (Obj.), while the emitted (Emi.) light is reflected by the DM and collected for analysis. (c) Low-temperature spectra of the eight hBN SPEs, labeled 1 through 8. The ZPL for all emitters is reproducible, centered around 436 ± 0.7 nm. (d) Two photon interference between successively emitted photons from the same source with a delay of 12.5 ns, yielding a V
HOM of 0.56 ± 0.11. (e) PL intensity versus excitation power for hBN SPEs, comparing uncoupled (blue) and coupled (red) configurations using a metallo-dielectric antenna setup. The coupled system achieves a near-unity photon collection efficiency of 98 %, compared to 13 % for the uncoupled case. The inset shows the emitter intensity over time, demonstrating excellent temporal stability without blinking. (f) High purity hBN SPEs with
Tran et al. [53], [54] first demonstrated single-photon emission from monolayer and multilayer hBN SPEs, with stable emission at 623 nm over 10 min and
Compared with many other platforms, hBN SPEs show reproducible emission wavelengths. Fournier et al. [185] demonstrated hBN SPEs operating at 436 nm that were created at controlled locations using electron beam irradiation. The local irradiation process activates SPE ensembles with submicron precision, leading to ZPLs consistently centered at 436 ± 1 nm (Figure 6c). Horder et al. [200] employed resonant excitation to characterize the emission line shape, demonstrating coherence of optical transitions through the observation of Rabi oscillations. Fournier et al. [186] conducted HOM experiments on the blue hBN SPEs, demonstrating corrected HOM visibility of 0.56 ± 0.11 at cryogenic temperatures (Figure 6d). The blue hBN SPEs exhibit spectral stability, room-temperature operation, ultra-narrow linewidth, and high I under nonresonant excitation, making them favorable for quantum frequency conversion (QFC) to telecommunications wavelengths.
Li et al. [93] demonstrated hBN SPEs coupled to metal-dielectric antennas, achieving near-unity light collection efficiency (98 %) and a QY of 12 % (Figure 6e). The SPEs exhibited high
B
(107 cps at 101 mW excitation power) and maintained single-photon emission under excitation powers up to 8 mW at room temperature. Vogl et al. [107] demonstrated hBN SPEs integrated with a tunable microcavity [201] consisting of a hemispherical and flat mirror, achieving corrected
These properties make hBN SPEs well-suited for several quantum applications. White et al. [202] demonstrated quantum random number generation (QRNG) using hBN SPEs coupled to an on-chip photonic waveguide structure at room temperature. Samaner et al. [115] integrated a hBN SPE into a QKD system based on the B92 protocol, achieving a sifted key rate of 238 bps with a quantum bit error rate of 8.95 % at a 1 MHz clock rate. Scognamiglio et al. [203] demonstrated that hBN SPEs operating at 417 nm show promise for underwater quantum communications.
One of hBN SPE’s most unique properties is its compatibility with spin-based quantum sensing [204]. Optically active spin defects, such as nitrogen vacancies [205], [206], exhibit a ground-state spin that can be optically addressed and manipulated. The manipulation of the spin states through external magnetic fields [207], temperature variation [208], [209], or strain [208], [210] forms the basis for their application in quantum sensing [211], [212]. These defects have the potential to detect minute changes in environmental parameters like magnetic fields or temperature at the nanoscale, leading to applications in high-sensitivity quantum sensing devices (Figure 7a). One of the most powerful techniques for characterizing and manipulating the spin properties of defects in quantum materials is Optically Detected Magnetic Resonance (ODMR) [216]. ODMR enables the detection of spin transitions in a defect’s electronic structure by combining microwave excitation and optical readout (Figure 7b). In the case of hBN spin defects, this method involves monitoring the fluorescence from the defect centers while applying microwave radiation to induce spin transitions between different energy levels.
![Figure 7:
Quantum spin sensing based on hBN defects. (a) Wide-field imaging of magnetization of an exfoliated Fe3GeTe2 flake by VB
− spin defects in hBN. (b) Dependence of ODMR frequencies on the magnetic field. Experimental data (red) and fit (blue line) with parameters D/h = 3.48 GHz, E/h = 50 MHz and g = 2.000. (c) Simplified VB
− energy-level diagram and the transitions among the ground state (3A2′), the excited state (3E′), and the metastable state (1E′, 1E′′). (d)
g
2
τ
${g}^{\left(2\right)}\left(\tau \right)$
of carbon-related defects in hBN. Inset indicates the fitted
g
2
0
${g}^{\left(2\right)}\left(0\right)$
= 0.25 ± 0.02. (e) The ODMR spectrum of a single defect in hBN measured in the absence of magnetic field. The top-right inset shows a confocal image of the PL intensity of the hBN device under 532 nm laser illumination. The bottom-right inset shows the pulse sequences used in the measurement. (f) A schematic illustration of quantum microscopy with spin defects in hBN. The setup includes a quantum active hBN flake and a sample to be imaged. The laser is used for excitation, and PL is collected for imaging. The microwave (MW) input enables control of the spin defects in the hBN for quantum sensing applications. Adapted with permission from: (a), ref. [211], Springer Nature; (b), ref. [207], Springer Nature; (c), ref. [208], Copyright 2021, American Chemical Society; (d), ref. [213], Springer Nature; (e), ref. [214], Springer Nature; (f), ref. [215], Springer Nature.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_007.jpg)
Quantum spin sensing based on hBN defects. (a) Wide-field imaging of magnetization of an exfoliated Fe3GeTe2 flake by VB
− spin defects in hBN. (b) Dependence of ODMR frequencies on the magnetic field. Experimental data (red) and fit (blue line) with parameters D/h = 3.48 GHz, E/h = 50 MHz and g = 2.000. (c) Simplified VB
− energy-level diagram and the transitions among the ground state (3A2′), the excited state (3E′), and the metastable state (1E′, 1E′′). (d)
The first spin signature for hBN defects was observed in boron vacancies (VB −), which exhibit a broad optical emission spectrum centered at ∼800 nm [207], [210]. These defects, which exhibit a broad optical emission spectrum centered around 800 nm, are characterized by a ground-state spin triplet (S = 1). The spin axis aligns along the c-axis of the hBN crystal, with a zero-field splitting of D ≈ 3.45 GHz separating the spin |0> and spin |±1> sublevels [208], [217] (Figure 7c). Through careful laser pulsing, the VB − defect’s spin states can be initialized, manipulated, and optically read out. However, due to their relatively low optical quantum efficiency, ODMR measurements have been limited to ensemble-level experiments rather than single-spin resolution.
In contrast, carbon-related spin defects in hBN [218], [219], some also with a spin triplet (S = 1), exhibit advantages for single-photon emission (Figure 7d). These defects, which emit light in the visible spectrum range, demonstrate bright room-temperature emission with more than 80 % of the emission occurring from the ZPL [220], [221]. They also show high ODMR contrast over 30 % (Figure 7e) and a long dephasing time, exceeding 100 ns [213], [222]. These features make them highly suitable for practical quantum sensing applications. Notably, these carbon-based defects achieve an estimated DC magnetic field sensitivity of approximately 3 μT/
Optically active spin defects in hBN bring two primary advantages over traditional NV centers in diamond, particularly for quantum sensing applications. First, quantum sensors in hBN have higher photon extraction efficiency: The 2D nature of hBN provides superior photon extraction efficiency compared to diamond. In hBN, the emission originates from the surface, minimizing internal photon losses caused by total internal reflection – a problem commonly encountered in bulk diamond. This enhanced efficiency can improve the signal-to-noise ratio in quantum measurements, making hBN a promising alternative for quantum sensing. Second, the quantum sensors in hBN could potentially offer better sensitivity and spatial resolution. Due to its atomically thin, planar structure, hBN quantum sensors can be positioned just a few ångströms away from the target object, leading to unprecedented sensitivity and spatial resolution. This capability is particularly important for applications such as imaging magnetic domains in 2D materials, where proximity to the sample is critical. The atomically smooth surface of hBN further enhances the sensing potential by reducing signal interference from surface roughness. The unique properties of hBN defects have already been demonstrated in experimental setups, including quantum microscopes [215], [225] and fiber-integrated devices [226] (Figure 7f). As research progresses, hBN is likely to play a pivotal role in the future of quantum technology, providing a versatile, high-performance platform for both fundamental studies and practical quantum devices.
Several challenges hinder the broader application of hBN SPEs. The uniformity of hBN SPEs is limited due to inherent challenges in material manipulation and control at the nanoscale that prevent the consistent creation of identical defects and often result in the formation of unintended defect types. Most hBN SPE ZPLs exist at wavelengths of 400–800 nm, with telecom wavelength emission yet to be demonstrated, limiting integration into optical fiber applications. Both the QY (current record of 12 %) [93] and I (current record of 56 %) [186] of hBN SPEs must be improved to realize technologically relevant hBN-based quantum photonic applications.
7 Few-layer transition metal dichalcogenides
The TMDCs have a layered MX2 structure with hexagonally arranged X-M-X units [227] (Figure 8a). The 2D TMDCs offer strong light–matter interaction [233], [234], direct bandgaps, large exciton binding energies (0.5–1 eV) [235], [236], and valley degrees of freedom [237], allowing for circularly polarized excitonic optical transitions and efficient tuning via magnetic field. The 2D TMDC-based SPEs offer high photon extraction efficiency, ease of coupling with external fields, and seamless integration into photonic circuits [38], [39]. Single-photon emission in TMDCs originates from excitons trapped by localized strain [228] or defects [238]. Strain is introduced using bubbles [239], [240], patterned nanostructures [241], [242], nanopillars [75], [230], or atomic force microscopy (AFM) tips [243], [244] to funnel excitons into localized regions (Figure 8b). Defects are introduced by material growth [55], [56], [57], [58], electron beam [245], [246], and ion beam irradiation [240].
![Figure 8:
Structure, strain, and performance of layered TMDC SPEs. (a) Atomic structures of two crystallographic phases of TMDCs. The 2H phase (top) features a trigonal prismatic coordination of metal atoms (M) and chalcogen atoms (X), with an A-B-A stacking sequence. The 1T phase (bottom) is characterized by octahedral coordination and a C-B-A stacking sequence. The side and top views highlight the differences in atomic arrangements between the two phases. (b) Illustration of the WSe2 monolayer transferred over gold nanorods. The strain induced by folds and wrinkles formed during the transfer process, particularly over the gaps between nanorods, leads to the localization of SPEs. (c) PL spectra from MoSe2/WSe2 heterobilayers with twist angles of 2° (green) and 20° (blue; intensity scaled by 130×). The twist angle significantly impacts the PL characteristics, with the 2° sample showing a strong peak near 1.3 eV and the 20° sample exhibiting multiple peaks around 1.6 eV. (d) PL intensity of WSe2 SPEs grown via flux and CVT methods, both before and after coupling to an optical cavity. Flux-grown SPEs show a quantum yield of up to 65.2 % after cavity coupling (red), compared to 16.5 % without coupling (blue). CVT-grown SPEs achieve a quantum yield of 12.6 % with coupling (green) and 1.5 % without coupling (black). (e) Second-order autocorrelation function
g
2
τ
${g}^{\left(2\right)}\left(\tau \right)$
for SPEs in a WSe2 monolayer under pulsed quasi-resonant excitation. The blue data points represent experimental measurements, while the red curve is a fit. The pronounced antibunching at zero delay time, with
g
2
0
${g}^{\left(2\right)}\left(0\right)$
equals to 0.036 ± 0.004. (f) HOM interference visibility V
HOM as a function of the temporal postselection window size for SPEs in a WSe2 monolayer coupled to a tunable open optical cavity. The blue line represents the measured visibility, with error bounds in gray. Visibility decreases with increasing integration window size, yielding V
HOM equals to 0.02. Adapted with permission from: (a), ref. [227], Licensee MDPI, Basel, Switzerland; (b), ref. [228], John Wiley and Sons; (c), ref. [229], Springer Nature; (d), ref. [230], Springer Nature; €, ref. [231], IOP Publishing; (f), ref. [232], American Chemical Society.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_008.jpg)
Structure, strain, and performance of layered TMDC SPEs. (a) Atomic structures of two crystallographic phases of TMDCs. The 2H phase (top) features a trigonal prismatic coordination of metal atoms (M) and chalcogen atoms (X), with an A-B-A stacking sequence. The 1T phase (bottom) is characterized by octahedral coordination and a C-B-A stacking sequence. The side and top views highlight the differences in atomic arrangements between the two phases. (b) Illustration of the WSe2 monolayer transferred over gold nanorods. The strain induced by folds and wrinkles formed during the transfer process, particularly over the gaps between nanorods, leads to the localization of SPEs. (c) PL spectra from MoSe2/WSe2 heterobilayers with twist angles of 2° (green) and 20° (blue; intensity scaled by 130×). The twist angle significantly impacts the PL characteristics, with the 2° sample showing a strong peak near 1.3 eV and the 20° sample exhibiting multiple peaks around 1.6 eV. (d) PL intensity of WSe2 SPEs grown via flux and CVT methods, both before and after coupling to an optical cavity. Flux-grown SPEs show a quantum yield of up to 65.2 % after cavity coupling (red), compared to 16.5 % without coupling (blue). CVT-grown SPEs achieve a quantum yield of 12.6 % with coupling (green) and 1.5 % without coupling (black). (e) Second-order autocorrelation function
Over the past decade, SPEs have been demonstrated in TMDCs such as WSe2 [55], [56], [57], [58], WS2 [241], [247], MoS2 [248], [249], [250], MoSe2 [251], [252], and MoTe2 [38]. Palacios-Berraquero et al. [247] demonstrated electrically excited SPEs in layered WSe2 and WS2. The SPEs in WSe2 emitted at 760 nm with
The weak van der Waals (vdW) interactions between TMDC layers allow for incommensurate stacking, which can result from relative rotation between the layers or, in heterobilayers, from lattice mismatch. The interplay between lattice mismatch and interlayer electronic coupling leads to the formation of interlayer excitons (IXs), where the exciton’s wavefunction spans both layers. For type-II band alignment, IXs possess out-of-plane electric dipole moments, enabling Stark tuning over a broad spectral range. Compared to intralayer excitons, IXs exhibit extended radiative lifetimes of order hundreds of nanoseconds and microsecond-scale valley lifetimes due to reduced electron–hole overlap. Combined with localized strain fields and in-gap defect states, such IXs lead to single-photon emission. Zhao et al. [253] demonstrated SPEs based on Γ – defect IXs in MoS2/WSe2 heterobilayers and heterotrilayers using nanopillars with a gold substrate. The nanopillars were used to create strain fields and defects, while the gold substrate quenched PL from the homogeneous region. The SPEs emitted at 855–1,078 nm with
Substantial work has gone into cavity enhancement of TMDC SPEs. For instance, Luo et al. [230] demonstrated WSe2 SPEs coupled to plasmonic cavities, achieving a Purcell factor up to 551 and an enhanced QY of up to 65 % (average 44 %) (Figure 8d). Sortino et al. [257] coupled WSe2 SPEs to GaP dielectric nano-antennas, demonstrating a 102–104 enhancement of the PL intensity. Von Helversen et al. [231] demonstrated WSe2 SPEs with
There is currently limited literature describing TMDCs integrated into functional quantum devices. For instance, Gao et al. [258] employed WSe2 SPEs with
8 Spectral tuning of SPEs
The reproducibility of LD-SPEs is limited by fabrication variability and material inhomogeneity. Spectral tuning can be achieved by strain engineering, the Stark effect, chemical functionalization, or combinations of these approaches. Spectral tuning via strain engineering was demonstrated in SPEs within 2D materials like hBN and TMDCs due to their high Young’s modulus (150–400 GPa). Grosso et al. [260] demonstrated that hBN SPEs can achieve spectral tunability of up to 6 meV by strain control using flexible polycarbonate (PC) beams. Xue et al. [261] measured the PL lines of hBN SPEs under varying hydrostatic pressure, demonstrating pressure coefficients of ∼15 meV/GPa at 5 K (Figure 9a and b). The SPEs exhibited a flexible bi-directional shift, showing both redshifts and blueshifts in response to pressure applied from different directions. Iff et al. [262] demonstrated a reversible tuning range of up to 18 meV in WSe2 SPEs using piezoelectric actuators (Figure 9c). They observed an energy shift of 5.4 μeV/V by sweeping the electric field applied to the piezoelectric actuator from −20 kV/cm to 20 kV/cm (Figure 9d).
![Figure 9:
Spectral tuning of SPEs. (a) Schematic representation of a 2D hBN flake under strain. The strain components ε
11 and ε
22 are applied along two principal axes of the lattice, demonstrating the uniaxial or biaxial strain induced in the material. The crystallographic orientation is indicated by the axes a
1, a
2, and a
3. (b) PL spectra of SPEs in the hBN flake under varying pressures, from 0.58 GPa to 3.20 GPa. The PL peak shows a redshift of ∼5 nm as the applied pressure increases, indicating strain-dependent spectral tuning. (c) Schematic representation of the experimental setup, featuring a WSe2 monolayer placed on a 200 µm piezoelectric substrate. Gold (Au) electrodes are used for electrical contact, and an external voltage is applied across the piezoelectric device to induce strain in the WSe2 layer, enabling spectral tuning. (d) PL spectra of the SPEs in WSe2 monolayer under different applied electric fields: +20 kV/cm (blue), 0 kV/cm (black), and −20 kV/cm (red). The shift in the PL peak energy with varying electric field demonstrates field-dependent control of emission properties. (e) Cross-sectional schematic of a heterostructure device comprising TMDC layers encapsulated by h-BN. The applied gate voltages (V
HS, V
TG, and V
BG) generate an electric field across the TMDC heterostructure, enabling electrical tuning of interlayer excitons (IX) via the DC Stark effect. A laser excites the system, creating interlayer excitons, which are influenced by both the electric field and potential strain (indicated by F
P
) applied to the device. (f) Illustration of the chemomechanical modification process for SPEs in monolayer WSe2 using aryl diazonium chemistry. Treatment with 4-NBD results in the physisorption of a nitrophenyl (NPh) oligomer layer, consisting of 2-ring and 3-ring structures, onto the WSe2 surface. This functionalization suppresses strain-induced defect emissions, enabling the formation of spectrally isolated SPEs. Nitrogen gas (N2) is released as a by-product of the reaction. Adapted with permission from: (a), (b), ref. [261], Copyright 2018, American Chemical Society; (c), (d), ref. [262], Copyright 2019, American Chemical Society; (e), ref. [263], Springer Nature; (f), ref. [264], Springer Nature.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_009.jpg)
Spectral tuning of SPEs. (a) Schematic representation of a 2D hBN flake under strain. The strain components ε 11 and ε 22 are applied along two principal axes of the lattice, demonstrating the uniaxial or biaxial strain induced in the material. The crystallographic orientation is indicated by the axes a 1, a 2, and a 3. (b) PL spectra of SPEs in the hBN flake under varying pressures, from 0.58 GPa to 3.20 GPa. The PL peak shows a redshift of ∼5 nm as the applied pressure increases, indicating strain-dependent spectral tuning. (c) Schematic representation of the experimental setup, featuring a WSe2 monolayer placed on a 200 µm piezoelectric substrate. Gold (Au) electrodes are used for electrical contact, and an external voltage is applied across the piezoelectric device to induce strain in the WSe2 layer, enabling spectral tuning. (d) PL spectra of the SPEs in WSe2 monolayer under different applied electric fields: +20 kV/cm (blue), 0 kV/cm (black), and −20 kV/cm (red). The shift in the PL peak energy with varying electric field demonstrates field-dependent control of emission properties. (e) Cross-sectional schematic of a heterostructure device comprising TMDC layers encapsulated by h-BN. The applied gate voltages (V HS, V TG, and V BG) generate an electric field across the TMDC heterostructure, enabling electrical tuning of interlayer excitons (IX) via the DC Stark effect. A laser excites the system, creating interlayer excitons, which are influenced by both the electric field and potential strain (indicated by F P ) applied to the device. (f) Illustration of the chemomechanical modification process for SPEs in monolayer WSe2 using aryl diazonium chemistry. Treatment with 4-NBD results in the physisorption of a nitrophenyl (NPh) oligomer layer, consisting of 2-ring and 3-ring structures, onto the WSe2 surface. This functionalization suppresses strain-induced defect emissions, enabling the formation of spectrally isolated SPEs. Nitrogen gas (N2) is released as a by-product of the reaction. Adapted with permission from: (a), (b), ref. [261], Copyright 2018, American Chemical Society; (c), (d), ref. [262], Copyright 2019, American Chemical Society; (e), ref. [263], Springer Nature; (f), ref. [264], Springer Nature.
Spectral tuning via the linear and quadratic Stark effect has been demonstrated in SPEs using hBN, TMDCs, and their heterostructures. The Stark effect is characterized by an energy shift Δℏω, which depends on the dipole moment
Spectral tuning via chemical functionalization was demonstrated in SPEs within SWCNTs and TMDCs. He et al. [74] demonstrated tunable wavelengths ranging from 1,280 nm to 1,550 nm in SWCNT SPEs at room temperature by varying the chiral index and aryl functionalization. Zhao et al. [33] demonstrated a wavelength shift in GQD SPEs, from approximately 650 nm to 760 nm after functionalization with chlorine atoms. Utama et al. [264] demonstrated a chemomechanical approach to modify the spectra of WSe2 SPEs (Figure 9f). They applied surface modification to strained monolayer WSe2 using 4-nitrobenzenediazonium (4-NBD) tetrafluoroborate. This process quenched most strain-induced defect emission, resulting in sharp SPEs with
9 Cavity coupling for SPE optimization
Coupling LD-SPEs to optical cavities enhances the
B
and
I
by introducing additional decay pathways and strong light–matter interactions due to electromagnetic confinement. As shown in Figure 10a, a two-level emitter coupled to an optical cavity mode can dissipate energy into the optical cavity via a Jaynes–Cummings type interaction [273], characterized by a coherent coupling rate (g). The cavity then releases the energy into the environment at a rate determined by the cavity linewidth (κ). The emission enhancement is quantified by the Purcell factor (F
P
), which is defined as
![Figure 10:
SPEs couple with cavities. (a) Schematic of a two-level emitter coupled to an optical cavity mode. The emitter experiences dissipation (γ) and interacts with the cavity mode via coupling strength g. Photons escape the cavity with decay rate κ, while γ∗ represents additional dissipation channels. (b) Lifetime measurements of 32 InAs/InP QDs from 20 different cavities, showing the impact of detuning on the QD lifetime. The shaded region represents the cavity effect, with the red dashed line indicating the lifetime of a bulk dot. (c) Schematic of a SWCNT coupled to a tunable microcavity. The cavity length is adjustable (indicated by the red arrows), allowing control over the optical coupling with the SPEs in the SWCNT. (d) Schematic of a hBN flake placed on a plasmonic nanoparticle array, covered by a poly (methyl methacrylate) (PMMA) film. (e) Fabrication process for integrating CVD-grown hBN with one-dimensional photonic crystal cavities. The steps include CVD growth, transfer, resist coating, electron beam lithography (EBL), and reactive ion etching (RIE) with undercutting. (f) Schematic of the experimental setup used to measure PL intensity and spontaneous emission rate from strain-induced SPEs in a WSe2 monolayer. Left: the emitters are positioned on gold pillars before the formation of the plasmonic nanocavity. Right: after the formation of the plasmonic nanocavity by flipping the material on a planar gold (Au) mirror, leading to enhanced emission properties. Adapted with permission from: (a), ref. [269], Copyright 2015 American Physical Society; (b), ref. [270], Copyright 2016 Optical Society of America; (c), ref. [271], Copyright 2017 American Chemical Society; (d), ref. [272], Copyright 2017, American Chemical Society; (e), ref. [221], Copyright 2020, American Chemical Society; (f), ref. [230], Springer Nature.](/document/doi/10.1515/nanoph-2024-0569/asset/graphic/j_nanoph-2024-0569_fig_010.jpg)
SPEs couple with cavities. (a) Schematic of a two-level emitter coupled to an optical cavity mode. The emitter experiences dissipation (γ) and interacts with the cavity mode via coupling strength g. Photons escape the cavity with decay rate κ, while γ∗ represents additional dissipation channels. (b) Lifetime measurements of 32 InAs/InP QDs from 20 different cavities, showing the impact of detuning on the QD lifetime. The shaded region represents the cavity effect, with the red dashed line indicating the lifetime of a bulk dot. (c) Schematic of a SWCNT coupled to a tunable microcavity. The cavity length is adjustable (indicated by the red arrows), allowing control over the optical coupling with the SPEs in the SWCNT. (d) Schematic of a hBN flake placed on a plasmonic nanoparticle array, covered by a poly (methyl methacrylate) (PMMA) film. (e) Fabrication process for integrating CVD-grown hBN with one-dimensional photonic crystal cavities. The steps include CVD growth, transfer, resist coating, electron beam lithography (EBL), and reactive ion etching (RIE) with undercutting. (f) Schematic of the experimental setup used to measure PL intensity and spontaneous emission rate from strain-induced SPEs in a WSe2 monolayer. Left: the emitters are positioned on gold pillars before the formation of the plasmonic nanocavity. Right: after the formation of the plasmonic nanocavity by flipping the material on a planar gold (Au) mirror, leading to enhanced emission properties. Adapted with permission from: (a), ref. [269], Copyright 2015 American Physical Society; (b), ref. [270], Copyright 2016 Optical Society of America; (c), ref. [271], Copyright 2017 American Chemical Society; (d), ref. [272], Copyright 2017, American Chemical Society; (e), ref. [221], Copyright 2020, American Chemical Society; (f), ref. [230], Springer Nature.
A growing literature has focused on optimizing
P
and
I
in 0D QD SPEs and 1D SWCNT SPEs through integration with optical cavities. For instance, Kim et al. [270] coupled InAs/InP QD SPEs to nanophotonic cavities, demonstrating a visibility of 67 % after postselection. They measured the lifetimes of 32 QDs from 20 cavities, observing a lifetime as short as 400 ps for the fastest-emitting QD, corresponding to a F
P
of 4.4 ± 0.5 (Figure 10b). Tomm et al. [96] coupled InGaAs EQD SPEs to a tunable Fabry–Pérot cavity, featuring a concave top mirror embedded in a silica substrate and a planar bottom mirror in the semiconductor heterostructure. They achieved
2D SPEs can conform closely with optical cavities, leading to a high F
P
. Tran et al. [272] demonstrated the deterministic coupling of hBN SPEs to plasmonic nanocavity arrays (Figure 10d), exhibiting 2.6-fold PL enhancement and
10 Conclusions and perspectives
LD-SPEs offer high BPI values, various levels of scalability, room-temperature operation, telecom emission, electrically and optically driven emission, and fine spectral tunability. Key hosts for LD-SPEs include QDs, SWCNTs, hBN, and TMDCs, each with unique strengths and challenges based on their synthesis methods and physical properties. CQD SPEs, especially PQD SPEs, demonstrate near-unity QY, room-temperature operation, high P (0.98), and I (0.56) but are limited by low PL intensity due to strong Auger recombination. EQD SPEs show strong BPI values and scalability but are typically restricted to cryogenic temperatures. SWCNT SPEs offer high P (0.69) and I (0.65) at room temperature in the telecom band but are constrained by low QY and strong exciton–exciton annihilation (EEA). hBN SPEs have demonstrated reproducible wavelengths (i.e., 436 nm for blue emitters), high light collection efficiency (98 %), and I (0.56) at room temperature but suffer from low QY and nontelecom emission. TMDC SPEs have achieved high QY (65 %), P (0.95), and integration into photonic heterostructures, but they suffer from low I (2 %) due to high dephasing rates. Resonant excitation schemes [280] have potential to improve the I of hBN and TMDC SPEs. Spectral tuning using strain engineering, the Stark effect, and chemical functionalization is essential to fine tuning emission wavelengths, and LD-SPEs are especially well suited to manipulation by nanoscale photonic cavities.
SPE arrays that can generate many indistinguishable single photons simultaneously are an essential building block for photonic quantum computing. The performance of such systems is affected by factors including the BPI values (Table 1) and the number of photons contributing to multiphoton interference. For example, in Boson sampling, achieving interference with 50 photons and an error threshold of 10 % requires a photon indistinguishability greater than 94.7 % [119]. A critical challenge for scalable quantum technologies based on LD-SPEs is ensuring uniform emitter performance and deterministic emitter placement. However, the literature targeting uniform SPE performance is limited because it is challenging to manipulate the solid-state environment of LD-SPEs at the nanoscale. Blue hBN SPEs exhibit reproducible wavelengths at 436 nm [185], but the P and I vary among emitters. While there is substantial literature exploring deterministic SPE patterning by ion [189]/laser [281] irradiation, these SPEs still exhibit substantial variation from emitter to emitter because of stochastic changes to the defect environment induced by the patterning [191]. Beyond defects and strain fields, IXs confined in the periodic moiré potential across TMDC bilayers present a promising avenue for SPE arrays [259] enabled by the intrinsic spatial reproducibility of the moiré lattice; improved control of clean 2D interfaces will be essential to the realization of moiré quantum photonic platforms.
Many new materials are emerging as candidates for LD-SPEs. 2D metal monochalcogenides (MMCs), such as InSe [282], [283], [284], SnS, GaSe [285], [286], and GeSe [79], [287], exhibit a direct bandgap in both multilayer and bulk forms [288], [289], [290], enabling realizations of SPEs via strain or defect engineering that are compatible integrated photonic platforms [291]. Mixed-dimensional materials [292] combine the features of different dimensionalities, such as the scalability of 0D QDs with the valleytronics of 2D systems [293]. SPEs from such mixed-dimensional heterostructures offer potential for high P and scalability. Atomically thin perovskites [294] exhibit remarkable optoelectronic properties, including high photoluminescence efficiency and tunable emission wavelengths. SPEs can be realized in such materials via techniques used on 2D materials such as electron-beam and femtosecond-laser irradiation. Recently, heterostructures have been shown to generate chiral single-photon emission [295] and improve P , offering new opportunities in quantum applications.
Despite substantial research efforts over the past decade and early demonstrations of quantum metrology [6], quantum computing [5], and quantum networking [296], most LD-SPEs have not yet reached the technological maturity of approaches based on QDs or heralded single photons from SPDC or SFWM processes. However, EQD SPEs have begun to reach the necessary thresholds in brightness, purity, and indistinguishability, which have led to a successful demonstration of 5-photon Boson sampling [116]. PQD SPEs and SWCNT SPEs also exhibit strong potential with their high I , although their functionality in actual applications has not yet been explicitly demonstrated. Based on recent advanced with these materials, it is reasonable to expect EQD, PQD, and SWCNT SPEs to be integrated into boosted Bell state measurements [297] and repeaters [298] in the near future. The applications of 2D SPEs are more limited due to low I , but the B and P achieved to date make them suitable for QKD applications [4], and further improvements in materials processing could unlock 2D SPEs for a broader range of quantum photonic technologies. The substantial progress in engineering LD spin defects over the past several years has unlocked new opportunities for spin-based quantum sensing, with a particular benefit offered by 2D spin defects in hBN that offer a dual purpose of encapsulating environmentally sensitive materials while hosting sensitive probes of local electric and magnetic fields. However, the literature focused on LD spin defects is still relatively new, and much work remains to determine the fundamental limits of these quantum sensors.
Beyond applications relying on emitted single photons, LD-SPEs may replace traditional solid-state SPEs, such as QD and color centers, in cavity-QED systems and enable novel designs that can be used as quantum memories [299], [300], [301], [302]. From an engineering perspective, LD-SPEs may be more compatible with integrated photonic cavities, but control over the SPE properties during photonic integration remains a technical challenge in many cases. The future development of LD-SPEs faces challenges in achieving desirable scalability, on-chip integration, and robustness. Ongoing advancements in theoretical research and fabrication techniques offer the potential to optimize LD-SPEs. As innovations in these areas keep enhancing efficiency and functionality, LD-SPEs are expected to unlock new possibilities across a broad range of applications, paving the way for transformative breakthroughs in quantum technologies and beyond.
Funding source: Army Research Office
Award Identifier / Grant number: W911NF2410080
Acknowledgment
JC, XY, SG, and ME acknowledge the support by Army Research Office under grant No. W911NF2410080. This work was supported by the Center for Nanophase Materials Sciences (CNMS), which is the U.S. Department of Energy, Office of Science User Facility at Oak Ridge National Laboratory. HZ was supported by the Wigner Distinguished Staff Fellowship at the Oak Ridge National Laboratory. BL was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division.
-
Research funding: JC, XY, SG, and ME acknowledge the support by Army Research Office under grant No. W911NF2410080. HZ was supported by the Wigner Distinguished Staff Fellowship at the Oak Ridge National Laboratory.
-
Author contributions: XY and HZ planned the structure of this article. XY, JC, and CC researched data for the article. XY, HZ, BL, SG, and ME contributed substantially to discussion of the content. JC, CC, HZ, and XY wrote the article. Each author (JC, CC, BL, YX, SG, ME, HZ, XY) reviewed and edited the article before submission. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Conflict of interest: Authors state no conflicts of interest.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.
References
[1] J. L. O’Brien, A. Furusawa, and J. Vučković, “Photonic quantum technologies,” Nat. Photonics, vol. 3, no. 12, pp. 687–695, 2009. https://doi.org/10.1038/nphoton.2009.229.Suche in Google Scholar
[2] P. Lodahl, S. Mahmoodian, and S. Stobbe, “Interfacing single photons and single quantum dots with photonic nanostructures,” Rev. Mod. Phys., vol. 87, no. 2, pp. 347–400, 2015. https://doi.org/10.1103/RevModPhys.87.347.Suche in Google Scholar
[3] J. Lee, et al.., “Integrated single photon emitters,” AVS Quantum Sci., vol. 2, no. 3, p. 031701, 2020. https://doi.org/10.1116/5.0011316.Suche in Google Scholar
[4] A. B. D. Shaik and P. Palla, “Optical quantum technologies with hexagonal boron nitride single photon sources,” Sci. Rep., vol. 11, no. 1, p. 12285, 2021. https://doi.org/10.1038/s41598-021-90804-4.Suche in Google Scholar PubMed PubMed Central
[5] M. Gimeno-Segovia, et al.., “From three-photon Greenberger-Horne-Zeilinger states to ballistic universal quantum computation,” Phys. Rev. Lett., vol. 115, no. 2, p. 20502, 2015. https://doi.org/10.1103/PhysRevLett.115.020502.Suche in Google Scholar PubMed
[6] P. Lombardi, et al.., “Advances in quantum metrology with dielectrically structured single photon sources based on molecules,” Adv. Quantum Technol., no. 10, p. 2400107, 2024. https://doi.org/10.1002/qute.202400107.Suche in Google Scholar
[7] M. Müller, et al.., “Quantum-dot single-photon sources for entanglement enhanced interferometry,” Phys. Rev. Lett., vol. 118, no. 25, p. 257402, 2017. https://doi.org/10.1103/PhysRevLett.118.257402.Suche in Google Scholar PubMed
[8] N. Gisin and R. Thew, “Quantum communication,” Nat. Photonics, vol. 1, no. 3, pp. 165–171, 2007. https://doi.org/10.1038/nphoton.2007.22.Suche in Google Scholar
[9] C. Kurtsiefer, et al.., “Stable solid-state source of single photons,” Phys. Rev. Lett., vol. 85, no. 2, 2000, https://doi.org/10.1103/physrevlett.85.290.Suche in Google Scholar PubMed
[10] I. Aharonovich, D. Englund, and M. Toth, “Solid-state single-photon emitters,” Nat. Photonics, vol. 10, no. 10, pp. 631–641, 2016. https://doi.org/10.1038/nphoton.2016.186.Suche in Google Scholar
[11] M. Esmann, S. C. Wein, and C. Antón-Solanas, “Solid-state single-photon sources: recent advances for novel quantum materials,” Adv. Funct. Mater., no. 30, p. 2315936, 2024. https://doi.org/10.1002/adfm.202315936.Suche in Google Scholar
[12] X. Ma, et al.., “Experimental generation of single photons via active multiplexing,” Phys. Rev. A, vol. 83, no. 4, p. 43814, 2011. https://doi.org/10.1103/PhysRevA.83.043814.Suche in Google Scholar
[13] J.-W. Pan, et al.., “Multiphoton entanglement and interferometry,” Rev. Mod. Phys., vol. 84, no. 2, pp. 777–838, 2012. https://doi.org/10.1103/RevModPhys.84.777.Suche in Google Scholar
[14] M. Takeoka, R.-B. Jin, and M. Sasaki, “Full analysis of multi-photon pair effects in spontaneous parametric down conversion based photonic quantum information processing,” New J. Phys., vol. 17, no. 4, p. 43030, 2015. https://doi.org/10.1088/1367-2630/17/4/043030.Suche in Google Scholar
[15] A. Chopin, et al.., “Ultra-efficient generation of time-energy entangled photon pairs in an InGaP photonic crystal cavity,” Commun. Phys., vol. 6, no. 1, p. 77, 2023. https://doi.org/10.1038/s42005-023-01189-x.Suche in Google Scholar
[16] M. J. Holmes, et al.., “Room-temperature triggered single photon emission from a III-nitride site-controlled nanowire quantum dot,” Nano Lett., vol. 14, no. 2, pp. 982–986, 2014. https://doi.org/10.1021/nl404400d.Suche in Google Scholar PubMed
[17] Y. Arakawa and M. J. Holmes, “Progress in quantum-dot single photon sources for quantum information technologies: a broad spectrum overview,” Appl. Phys. Rev., vol. 7, no. 2, p. 21309, 2020. https://doi.org/10.1063/5.0010193.Suche in Google Scholar
[18] T. Zhong, et al.., “Optically addressing single rare-earth ions in a nanophotonic cavity,” Phys. Rev. Lett., vol. 121, no. 18, p. 183603, 2018. https://doi.org/10.1103/PhysRevLett.121.183603.Suche in Google Scholar PubMed
[19] J. M. Kindem, et al.., “Control and single-shot readout of an ion embedded in a nanophotonic cavity,” Nature, vol. 580, no. 7802, pp. 201–204, 2020. https://doi.org/10.1038/s41586-020-2160-9.Suche in Google Scholar PubMed
[20] S. Ourari, et al.., “Indistinguishable telecom band photons from a single Er ion in the solid state,” Nature, vol. 620, no. 7976, pp. 977–981, 2023. https://doi.org/10.1038/s41586-023-06281-4.Suche in Google Scholar PubMed
[21] A. Dräbenstedt, et al.., “Low-temperature microscopy and spectroscopy on single defect centers in diamond,” Phys. Rev. B, vol. 60, no. 16, pp. 11503–11508, 1999. https://doi.org/10.1103/PhysRevB.60.11503.Suche in Google Scholar
[22] N. B. Manson and J. P. Harrison, “Photo-ionization of the nitrogen-vacancy center in diamond,” Diamond Relat. Mater., vol. 14, no. 10, pp. 1705–1710, 2005. https://doi.org/10.1016/j.diamond.2005.06.027.Suche in Google Scholar
[23] F. Treussart, et al.., “Photoluminescence of single colour defects in 50 nm diamond nanocrystals,” Phys. B Condens. Matter, vols. 376–377, pp. 926–929, 2006, https://doi.org/10.1016/j.physb.2005.12.232.Suche in Google Scholar
[24] M. W. Doherty, et al.., “The nitrogen-vacancy colour centre in diamond,” Phys. Rep., vol. 528, no. 1, pp. 1–45, 2013. https://doi.org/10.1016/j.physrep.2013.02.001.Suche in Google Scholar
[25] R. Schirhagl, et al.., “Nitrogen-vacancy centers in diamond: nanoscale sensors for physics and biology,” Annu. Rev. Phys. Chem., vol. 65, no. 1, pp. 83–105, 2014. https://doi.org/10.1146/annurev-physchem-040513-103659.Suche in Google Scholar PubMed
[26] P. Udvarhelyi, et al.., “Identification of a telecom wavelength single photon emitter in silicon,” Phys. Rev. Lett., vol. 127, no. 19, p. 196402, 2021. https://doi.org/10.1103/PhysRevLett.127.196402.Suche in Google Scholar PubMed
[27] M. Hollenbach, et al.., “Wafer-scale nanofabrication of telecom single-photon emitters in silicon,” Nat. Commun., vol. 13, no. 1, p. 7683, 2022. https://doi.org/10.1038/s41467-022-35051-5.Suche in Google Scholar PubMed PubMed Central
[28] N. H. Wan, et al.., “Efficient extraction of light from a nitrogen-vacancy center in a diamond parabolic reflector,” Nano Lett., vol. 18, no. 5, pp. 2787–2793, 2018. https://doi.org/10.1021/acs.nanolett.7b04684.Suche in Google Scholar PubMed
[29] X. Liu and M. C. Hersam, “2D materials for quantum information science,” Nat. Rev. Mater., vol. 4, no. 10, pp. 669–684, 2019. https://doi.org/10.1038/s41578-019-0136-x.Suche in Google Scholar
[30] T. M. Babinec, et al.., “A diamond nanowire single-photon source,” Nat. Nanotechnol., vol. 5, no. 3, pp. 195–199, 2010. https://doi.org/10.1038/nnano.2010.6.Suche in Google Scholar PubMed
[31] I. Aharonovich, et al.., “Diamond-based single-photon emitters,” Rep. Prog. Phys., vol. 74, no. 7, p. 076501, 2011. https://doi.org/10.1088/0034-4885/74/7/076501.Suche in Google Scholar
[32] J. Almutlaq, et al.., “Closed-loop electron-beam-induced spectroscopy and nanofabrication around individual quantum emitters,” Nanophotonics, vol. 13, no. 12, pp. 2251–2258, 2024. https://doi.org/10.1515/nanoph-2023-0877.Suche in Google Scholar PubMed PubMed Central
[33] S. Zhao, et al.., “Single photon emission from graphene quantum dots at room temperature,” Nat. Commun., vol. 9, no. 1, p. 3470, 2018. https://doi.org/10.1038/s41467-018-05888-w].10.1038/s41467-018-05888-wSuche in Google Scholar PubMed PubMed Central
[34] M. Toth and I. Aharonovich, “Single photon sources in atomically thin materials,” Annu. Rev. Phys. Chem., vol. 70, no. 1, pp. 123–142, 2019. https://doi.org/10.1146/annurev-physchem-042018-052628.Suche in Google Scholar PubMed
[35] S. Gupta, et al.., “Single-photon emission from two-dimensional materials, to a brighter future,” J. Phys. Chem. Lett., vol. 14, no. 13, pp. 3274–3284, 2023. https://doi.org/10.1021/acs.jpclett.2c03674.Suche in Google Scholar PubMed
[36] S. Jesse, et al.., “Direct atomic fabrication and dopant positioning in Si using electron beams with active real-time image-based feedback,” Nanotechnology, vol. 29, no. 25, p. 255303, 2018. https://doi.org/10.1088/1361-6528/aabb79.Suche in Google Scholar PubMed
[37] O. Dyck, et al.., “Atom-by-atom fabrication with electron beams,” Nat. Rev. Mater., vol. 4, no. 7, pp. 497–507, 2019. https://doi.org/10.1038/s41578-019-0118-z.Suche in Google Scholar
[38] H. Zhao, et al.., “Site-controlled telecom-wavelength single-photon emitters in atomically-thin MoTe2,” Nat. Commun., vol. 12, no. 1, p. 6753, 2021. https://doi.org/10.1038/s41467-021-27033-w.Suche in Google Scholar PubMed PubMed Central
[39] S. Häußler, et al.., “Tunable fiber-cavity enhanced photon emission from defect centers in hBN,” Adv. Opt. Mater., vol. 9, no. 17, p. 2002218, 2021. https://doi.org/10.1002/adom.202002218.Suche in Google Scholar
[40] X. Xu, et al.., “Spin and pseudospins in layered transition metal dichalcogenides,” Nat. Phys., vol. 10, no. 5, pp. 343–350, 2014. https://doi.org/10.1038/nphys2942.Suche in Google Scholar
[41] Y. J. Bae, et al.., “Exciton-coupled coherent magnons in a 2D semiconductor,” Nature, vol. 609, no. 7926, pp. 282–286, 2022. https://doi.org/10.1038/s41586-022-05024-1.Suche in Google Scholar PubMed
[42] N. J. Brennan, et al.., “Important elements of spin-exciton and magnon-exciton coupling,” ACS Phys. Chem. Au, vol. 4, no. 4, pp. 322–327, 2024. https://doi.org/10.1021/acsphyschemau.4c00010.Suche in Google Scholar PubMed PubMed Central
[43] K. F. Mak, et al.., “Control of valley polarization in monolayer MoS2 by optical helicity,” Nat. Nanotechnol., vol. 7, no. 8, pp. 494–498, 2012. https://doi.org/10.1038/nnano.2012.96.Suche in Google Scholar PubMed
[44] J. Yan, et al.., “Double-pulse generation of indistinguishable single photons with optically controlled polarization,” Nano Lett., vol. 22, no. 4, pp. 1483–1490, 2022. https://doi.org/10.1021/acs.nanolett.1c03543].10.1021/acs.nanolett.1c03543Suche in Google Scholar PubMed
[45] S. Ditalia Tchernij, et al.., “Single-photon emitters in lead-implanted single-crystal diamond,” ACS Photonics, vol. 5, no. 12, pp. 4864–4871, 2018. https://doi.org/10.1021/acsphotonics.8b01013.Suche in Google Scholar
[46] G. Zhang, et al.., “Material platforms for defect qubits and single-photon emitters,” Appl. Phys. Rev., vol. 7, no. 3, p. 031308, 2020. https://doi.org/10.1063/5.0006075.Suche in Google Scholar
[47] J. Wang, et al.., “Bright room temperature single photon source at telecom range in cubic silicon carbide,” Nat. Commun., vol. 9, no. 1, p. 4106, 2018. https://doi.org/10.1038/s41467-018-06605-3.Suche in Google Scholar PubMed PubMed Central
[48] A. Senichev, et al.., “Room-temperature single-photon emitters in silicon nitride,” Sci. Adv., vol. 7, no. 50, p. eabj0627, 2021. https://doi.org/10.1126/sciadv.abj0627.Suche in Google Scholar PubMed PubMed Central
[49] S. G. Bishop, et al.., “Room-temperature quantum emitter in aluminum nitride,” ACS Photonics, vol. 7, no. 7, pp. 1636–1641, 2020. https://doi.org/10.1021/acsphotonics.0c00528.Suche in Google Scholar PubMed PubMed Central
[50] Y. Xue, et al.., “Single-photon emission from point defects in aluminum nitride films,” J. Phys. Chem. Lett., vol. 11, no. 7, pp. 2689–2694, 2020. https://doi.org/10.1021/acs.jpclett.0c00511.Suche in Google Scholar PubMed
[51] A. M. Berhane, et al.., “Bright room-temperature single-photon emission from defects in gallium nitride,” Adv. Mater., vol. 29, no. 12, p. 1605092, 2017. https://doi.org/10.1002/adma.201605092.Suche in Google Scholar PubMed
[52] Y. Zhou, et al.., “Room temperature solid-state quantum emitters in the telecom range,” Sci. Adv., vol. 4, no. 3, p. eaar3580, 2018. https://doi.org/10.1126/sciadv.aar3580.Suche in Google Scholar PubMed PubMed Central
[53] T. T. Tran, et al.., “Quantum emission from defects in single-crystalline hexagonal boron nitride,” Phys. Rev. Appl., vol. 5, no. 3, p. 034005, 2016. https://doi.org/10.1103/PhysRevApplied.5.034005.Suche in Google Scholar
[54] T. T. Tran, et al.., “Quantum emission from hexagonal boron nitride monolayers,” Nat. Nanotech., vol. 11, no. 1, pp. 37–41, 2016. https://doi.org/10.1038/nnano.2015.242.Suche in Google Scholar PubMed
[55] A. Srivastava, et al.., “Optically active quantum dots in monolayer WSe2,” Nat. Nanotech., vol. 10, no. 6, pp. 491–496, 2015. https://doi.org/10.1038/nnano.2015.60.Suche in Google Scholar PubMed
[56] M. Koperski, et al.., “Single photon emitters in exfoliated WSe2 structures,” Nat. Nanotech., vol. 10, no. 6, pp. 503–506, 2015. https://doi.org/10.1038/nnano.2015.67.Suche in Google Scholar PubMed
[57] Y.-M. He, et al.., “Single quantum emitters in monolayer semiconductors,” Nat. Nanotech., vol. 10, no. 6, pp. 497–502, 2015. https://doi.org/10.1038/nnano.2015.75.Suche in Google Scholar PubMed
[58] C. Chakraborty, et al.., “Voltage-controlled quantum light from an atomically thin semiconductor,” Nat. Nanotech., vol. 10, no. 6, pp. 507–511, 2015. https://doi.org/10.1038/nnano.2015.79.Suche in Google Scholar PubMed
[59] X. Ma, et al.., “Room-temperature single-photon generation from solitary dopants of carbon nanotubes,” Nat. Nanotechnol., vol. 10, no. 8, pp. 671–675, 2015. https://doi.org/10.1038/nnano.2015.136.Suche in Google Scholar PubMed
[60] F. Pyatkov, et al.., “Cavity-enhanced light emission from electrically driven carbon nanotubes,” Nat. Photonics, vol. 10, no. 6, pp. 420–427, 2016. https://doi.org/10.1038/nphoton.2016.70.Suche in Google Scholar
[61] M. E. Reimer, et al.., “Bright single-photon sources in bottom-up tailored nanowires,” Nat. Commun., vol. 3, no. 1, p. 737, 2012. https://doi.org/10.1038/ncomms1746.Suche in Google Scholar PubMed PubMed Central
[62] T. B. Hoang, G. M. Akselrod, and M. H. Mikkelsen, “Ultrafast room-temperature single photon emission from quantum dots coupled to plasmonic nanocavities,” Nano Lett., vol. 16, no. 1, pp. 270–275, 2016. https://doi.org/10.1021/acs.nanolett.5b03724.Suche in Google Scholar PubMed
[63] Z. Yuan, et al.., “Electrically driven single-photon source,” Science, vol. 295, no. 5552, pp. 102–105, 2002. https://doi.org/10.1126/science.1066790.Suche in Google Scholar PubMed
[64] C. Santori, et al.., “Indistinguishable photons from a single-photon device,” Nature, vol. 419, no. 6907, pp. 594–597, 2002. https://doi.org/10.1038/nature01086.Suche in Google Scholar PubMed
[65] M. A. Feldman, et al.., “Evidence of photochromism in a hexagonal boron nitride single-photon emitter,” Optica, vol. 8, no. 1, pp. 1–5, 2021. https://doi.org/10.1364/OPTICA.406184.Suche in Google Scholar
[66] A. Sajid, M. J. Ford, and J. R. Reimers, “Single-photon emitters in hexagonal boron nitride: a review of progress,” Rep. Prog. Phys., vol. 83, no. 4, p. 044501, 2020. https://doi.org/10.1088/1361-6633/ab6310.Suche in Google Scholar PubMed
[67] T. T. Tran, et al.., “Robust multicolor single photon emission from point defects in hexagonal boron nitride,” ACS Nano, vol. 10, no. 8, pp. 7331–7338, 2016. https://doi.org/10.1021/acsnano.6b03602.Suche in Google Scholar PubMed
[68] L. Gan, et al.., “Large-scale, high-yield laser fabrication of bright and pure single-photon emitters at room temperature in hexagonal boron nitride,” ACS Nano, vol. 16, no. 9, pp. 14254–14261, 2022. https://doi.org/10.1021/acsnano.2c04386.Suche in Google Scholar PubMed
[69] N. Chejanovsky, et al.., “Structural attributes and photodynamics of visible spectrum quantum emitters in hexagonal boron nitride,” Nano Lett., vol. 16, no. 11, pp. 7037–7045, 2016. https://doi.org/10.1021/acs.nanolett.6b03268.Suche in Google Scholar PubMed
[70] M. Nonahal, et al.., “Deterministic fabrication of a coupled cavity–emitter system in hexagonal boron nitride,” Nano Lett., vol. 23, no. 14, pp. 6645–6650, 2023. https://doi.org/10.1021/acs.nanolett.3c01836.Suche in Google Scholar PubMed
[71] M. Nguyen, et al.., “Nanoassembly of quantum emitters in hexagonal boron nitride and gold nanospheres,” Nanoscale, vol. 10, no. 5, pp. 2267–2274, 2018. https://doi.org/10.1039/C7NR08249E.Suche in Google Scholar PubMed
[72] T. Vogl, et al.., “Fabrication and deterministic transfer of high-quality quantum emitters in hexagonal boron nitride,” ACS Photonics, vol. 5, no. 6, pp. 2305–2312, 2018. https://doi.org/10.1021/acsphotonics.8b00127.Suche in Google Scholar
[73] Z.-Q. Xu, et al.., “Single photon emission from plasma treated 2D hexagonal boron nitride,” Nanoscale, vol. 10, no. 17, pp. 7957–7965, 2018. https://doi.org/10.1039/C7NR08222C.Suche in Google Scholar PubMed
[74] X. He, et al.., “Tunable room-temperature single-photon emission at telecom wavelengths from sp3 defects in carbon nanotubes,” Nat. Photonics, vol. 11, no. 9, pp. 577–582, 2017. https://doi.org/10.1038/nphoton.2017.119.Suche in Google Scholar
[75] A. Branny, et al.., “Deterministic strain-induced arrays of quantum emitters in a two-dimensional semiconductor,” Nat. Commun., vol. 8, no. 1, p. 15053, 2017. https://doi.org/10.1038/ncomms15053.Suche in Google Scholar PubMed PubMed Central
[76] C. Chakraborty, N. Vamivakas, and D. Englund, “Advances in quantum light emission from 2D materials,” Nanophotonics, vol. 8, no. 11, pp. 2017–2032, 2019. https://doi.org/10.1515/nanoph-2019-0140.Suche in Google Scholar
[77] F. Peyskens, et al.., “Integration of single photon emitters in 2D layered materials with a silicon nitride photonic chip,” Nat. Commun., vol. 10, no. 1, p. 4435, 2019. https://doi.org/10.1038/s41467-019-12421-0.Suche in Google Scholar PubMed PubMed Central
[78] A. Saha, et al.., “Narrow-band single-photon emission through selective aryl functionalization of zigzag carbon nanotubes,” Nat. Chem., vol. 10, no. 11, pp. 1089–1095, 2018. https://doi.org/10.1038/s41557-018-0126-4.Suche in Google Scholar PubMed
[79] P. Tonndorf, et al.., “Single-photon emitters in GaSe,” 2D Mater., vol. 4, no. 2, p. 021010, 2017. https://doi.org/10.1088/2053-1583/aa525b.Suche in Google Scholar
[80] N. R. Jungwirth, et al.., “Temperature dependence of wavelength selectable zero-phonon emission from single defects in hexagonal boron nitride,” Nano Lett., vol. 16, no. 10, pp. 6052–6057, 2016. https://doi.org/10.1021/acs.nanolett.6b01987.Suche in Google Scholar PubMed
[81] G. Clark, et al.., “Single defect light-emitting diode in a van der Waals heterostructure,” Nano Lett., vol. 16, no. 6, pp. 3944–3948, 2016. https://doi.org/10.1021/acs.nanolett.6b01580.Suche in Google Scholar PubMed
[82] W. Luo, et al.., “Imaging strain-localized single-photon emitters in layered GaSe below the diffraction limit,” ACS Nano, vol. 17, no. 23, pp. 23455–23465, 2023. https://doi.org/10.1021/acsnano.3c05250.Suche in Google Scholar PubMed
[83] M. A. Feldman, et al.., “Colossal photon bunching in quasiparticle-mediated nanodiamond cathodoluminescence,” Phys. Rev. B, vol. 97, no. 8, p. 081404, 2018. https://doi.org/10.1103/PhysRevB.97.081404.Suche in Google Scholar
[84] V. Iyer, et al.., “Photon bunching in cathodoluminescence induced by indirect electron excitation,” Nanoscale, vol. 15, no. 22, pp. 9738–9744, 2023. https://doi.org/10.1039/D3NR00376K.Suche in Google Scholar
[85] D. Curie, et al.., “Correlative nanoscale imaging of strained hBN spin defects,” ACS Appl. Mater. Interfaces, vol. 14, no. 36, pp. 41361–41368, 2022. https://doi.org/10.1021/acsami.2c11886.Suche in Google Scholar PubMed
[86] S. Roux, et al.., “Cathodoluminescence monitoring of quantum emitter activation in hexagonal boron nitride,” Appl. Phys. Lett., vol. 121, no. 18, p. 184002, 2022. https://doi.org/10.1063/5.0126357.Suche in Google Scholar
[87] R. Bourrellier, et al.., “Bright UV single photon emission at point defects in h -BN,” Nano Lett., vol. 16, no. 7, pp. 4317–4321, 2016. https://doi.org/10.1021/acs.nanolett.6b01368.Suche in Google Scholar PubMed
[88] L. H. G. Tizei and M. Kociak, “Spatially resolved quantum nano-optics of single photons using an electron microscope,” Phys. Rev. Lett., vol. 110, no. 15, p. 153604, 2013. https://doi.org/10.1103/PhysRevLett.110.153604.Suche in Google Scholar PubMed
[89] A. Gale, et al.., “Site-specific fabrication of blue quantum emitters in hexagonal boron nitride,” ACS Photonics, vol. 9, no. 6, pp. 2170–2177, 2022. https://doi.org/10.1021/acsphotonics.2c00631.Suche in Google Scholar
[90] S. Meuret, et al.., “Photon bunching in cathodoluminescence,” Phys. Rev. Lett., vol. 114, no. 19, p. 197401, 2015. https://doi.org/10.1103/PhysRevLett.114.197401.Suche in Google Scholar PubMed
[91] M. Solà-Garcia, et al.., “Photon statistics of incoherent cathodoluminescence with continuous and pulsed electron beams,” ACS Photonics, vol. 8, no. 3, pp. 916–925, 2021. https://doi.org/10.1021/acsphotonics.0c01939.Suche in Google Scholar PubMed PubMed Central
[92] S. Fiedler, et al.., “Sub-to-super-poissonian photon statistics in cathodoluminescence of color center ensembles in isolated diamond crystals,” Nanophotonics, vol. 12, no. 12, pp. 2231–2237, 2023. https://doi.org/10.1515/nanoph-2023-0204.Suche in Google Scholar PubMed PubMed Central
[93] X. Li, et al.., “Near-unity light collection efficiency from quantum emitters in boron nitride by coupling to metallo-dielectric antennas,” ACS Nano, vol. 13, no. 6, pp. 6992–6997, 2019. https://doi.org/10.1021/acsnano.9b01996.Suche in Google Scholar PubMed
[94] B. D. Mangum, et al.., “Disentangling the effects of clustering and multi-exciton emission in second-order photon correlation experiments,” Opt. Express, vol. 21, no. 6, p. 7419, 2013. https://doi.org/10.1364/OE.21.007419.Suche in Google Scholar PubMed PubMed Central
[95] H. Wang, et al.., “Towards optimal single-photon sources from polarized microcavities,” Nat. Photonics, vol. 13, no. 11, pp. 770–775, 2019. https://doi.org/10.1038/s41566-019-0494-3.Suche in Google Scholar
[96] N. Tomm, et al.., “A bright and fast source of coherent single photons,” Nat. Nanotechnol., vol. 16, no. 4, pp. 399–403, 2021. https://doi.org/10.1038/s41565-020-00831-x.Suche in Google Scholar PubMed
[97] T. T. Tran, et al.., “Resonant excitation of quantum emitters in hexagonal boron nitride,” ACS Photonics, vol. 5, no. 2, pp. 295–300, 2018. https://doi.org/10.1021/acsphotonics.7b00977.Suche in Google Scholar
[98] J. M. Raimond, M. Brune, and S. Haroche, “Colloquium: manipulating quantum entanglement with atoms and photons in a cavity,” Rev. Mod. Phys., vol. 73, no. 3, 2001, https://doi.org/10.1103/revmodphys.73.565.Suche in Google Scholar
[99] H. Jayakumar, et al.., “Deterministic photon pairs and coherent optical control of a single quantum dot,” Phys. Rev. Lett., vol. 110, no. 13, p. 135505, 2013. https://doi.org/10.1103/PhysRevLett.110.135505.Suche in Google Scholar PubMed
[100] Y. Karli, et al.., “SUPER scheme in action: experimental demonstration of red-detuned excitation of a quantum emitter,” Nano Lett., vol. 22, no. 16, pp. 6567–6572, 2022. https://doi.org/10.1021/acs.nanolett.2c01783.Suche in Google Scholar PubMed PubMed Central
[101] F. Sbresny, et al.., “Stimulated generation of indistinguishable single photons from a quantum ladder system,” Phys. Rev. Lett., vol. 128, no. 9, p. 93603, 2022. https://doi.org/10.1103/PhysRevLett.128.093603.Suche in Google Scholar PubMed
[102] Y. Wei, et al.., “Tailoring solid-state single-photon sources with stimulated emissions,” Nat. Nanotechnol., vol. 17, no. 5, pp. 470–476, 2022. https://doi.org/10.1038/s41565-022-01092-6.Suche in Google Scholar PubMed
[103] L. Husel, et al.., “Cavity-enhanced photon indistinguishability at room temperature and telecom wavelengths,” Nat. Commun., vol. 15, no. 1, p. 3989, 2024. https://doi.org/10.1038/s41467-024-48119-1.Suche in Google Scholar PubMed PubMed Central
[104] M. Fox, Quantum Optics: An Introduction, Oxford, New York, Oxford University Press, 2006.Suche in Google Scholar
[105] C. Jones, et al.., “Time-dependent Mandel Q parameter analysis for a hexagonal boron nitride single photon source,” Opt. Express, vol. 31, no. 6, p. 10794, 2023. https://doi.org/10.1364/OE.485216.Suche in Google Scholar PubMed
[106] R. H. Brown and R. Q. Twiss, “Correlation between photons in two coherent beams of light,” Nature, vol. 177, no. 4497, pp. 27–29, 1956. https://doi.org/10.1038/177027a0.Suche in Google Scholar
[107] T. Vogl, et al.., “Sensitive single-photon test of extended quantum theory with two-dimensional hexagonal boron nitride,” Phys. Rev. Res., vol. 3, no. 1, p. 13296, 2021. https://doi.org/10.1103/PhysRevResearch.3.013296.Suche in Google Scholar
[108] C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett., vol. 59, no. 18, pp. 2044–2046, 1987. https://doi.org/10.1103/PhysRevLett.59.2044.Suche in Google Scholar PubMed
[109] P. J. Mosley, et al.., “Heralded generation of ultrafast single photons in pure quantum states,” Phys. Rev. Lett., vol. 100, no. 13, p. 133601, 2008. https://doi.org/10.1103/PhysRevLett.100.133601.Suche in Google Scholar PubMed
[110] 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.Suche in Google Scholar
[111] P. Borri, et al.., “Ultralong dephasing time in InGaAs quantum dots,” Phys. Rev. Lett., vol. 87, no. 15, p. 157401, 2001. https://doi.org/10.1103/PhysRevLett.87.157401.Suche in Google Scholar PubMed
[112] H. Utzat, et al.., “Coherent single-photon emission from colloidal lead halide perovskite quantum dots,” Science, vol. 363, no. 6431, pp. 1068–1072, 2019. https://doi.org/10.1126/science.aau7392.Suche in Google Scholar PubMed
[113] K. Takemoto, et al.., “Quantum key distribution over 120 km using ultrahigh purity single-photon source and superconducting single-photon detectors,” Sci. Rep., vol. 5, no. 1, p. 14383, 2015. https://doi.org/10.1038/srep14383.Suche in Google Scholar PubMed PubMed Central
[114] H.-K. Lo, M. Curty, and K. Tamaki, “Secure quantum key distribution,” Nat. Photonics, vol. 8, no. 8, pp. 595–604, 2014. https://doi.org/10.1038/nphoton.2014.149.Suche in Google Scholar
[115] Ç. Samaner, et al.., “Free-space quantum key distribution with single photons from defects in hexagonal boron nitride,” Adv. Quantum Technol., vol. 5, no. 9, p. 2200059, 2022. https://doi.org/10.1002/qute.202200059.Suche in Google Scholar
[116] H. Wang, et al.., “High-efficiency multiphoton boson sampling,” Nat. Photonics, vol. 11, no. 6, pp. 361–365, 2017. https://doi.org/10.1038/nphoton.2017.63.Suche in Google Scholar
[117] R. García-Patrón, J. J. Renema, and V. Shchesnovich, “Simulating boson sampling in lossy architectures,” Quantum, vol. 3, p. 169, 2019, https://doi.org/10.22331/q-2019-08-05-169.Suche in Google Scholar
[118] N. Maring, et al.., “A versatile single-photon-based quantum computing platform,” Nat. Photonics, vol. 18, no. 6, pp. 603–609, 2024. https://doi.org/10.1038/s41566-024-01403-4.Suche in Google Scholar
[119] J. J. Renema, et al.., “Efficient classical algorithm for boson sampling with partially distinguishable photons,” Phys. Rev. Lett., vol. 120, no. 22, p. 220502, 2018. https://doi.org/10.1103/PhysRevLett.120.220502.Suche in Google Scholar PubMed
[120] M. Reindl, et al.., “All-photonic quantum teleportation using on-demand solid-state quantum emitters,” Sci. Adv., vol. 4, no. 12, p. eaau1255, 2018. https://doi.org/10.1126/sciadv.aau1255.Suche in Google Scholar PubMed PubMed Central
[121] M. Anderson, et al.., “Quantum teleportation using highly coherent emission from telecom C-band quantum dots,” Npj Quantum Inf., vol. 6, no. 1, p. 14, 2020. https://doi.org/10.1038/s41534-020-0249-5.Suche in Google Scholar
[122] D. Istrati, et al.., “Sequential generation of linear cluster states from a single photon emitter,” Nat. Commun., vol. 11, no. 1, p. 5501, 2020. https://doi.org/10.1038/s41467-020-19341-4.Suche in Google Scholar PubMed PubMed Central
[123] D. Cogan, et al.., “Deterministic generation of indistinguishable photons in a cluster state,” Nat. Photonics, vol. 17, no. 4, pp. 324–329, 2023. https://doi.org/10.1038/s41566-022-01152-2.Suche in Google Scholar PubMed PubMed Central
[124] N. Coste, et al.., “High-rate entanglement between a semiconductor spin and indistinguishable photons,” Nat. Photonics, vol. 17, no. 7, pp. 582–587, 2023. https://doi.org/10.1038/s41566-023-01186-0.Suche in Google Scholar
[125] M. H. Appel, et al.., “Entangling a hole spin with a time-bin photon: a waveguide approach for quantum dot sources of multiphoton entanglement,” Phys. Rev. Lett., vol. 128, no. 23, p. 233602, 2022. https://doi.org/10.1103/PhysRevLett.128.233602.Suche in Google Scholar PubMed
[126] K. Tiurev, et al.., “High-fidelity multiphoton-entangled cluster state with solid-state quantum emitters in photonic nanostructures,” Phys. Rev. A, vol. 105, no. 3, p. L030601, 2022. https://doi.org/10.1103/PhysRevA.105.L030601.Suche in Google Scholar
[127] H. Cao, et al.., “Photonic source of heralded greenberger-horne-zeilinger states,” Phys. Rev. Lett., vol. 132, no. 13, p. 130604, 2024. https://doi.org/10.1103/PhysRevLett.132.130604.Suche in Google Scholar PubMed
[128] C. H. Bennett and G. Brassard, “Quantum cryptography: public key distribution and coin tossing,” Theor. Comput. Sci., vol. 560, pp. 7–11, 2014, https://doi.org/10.1016/j.tcs.2014.05.025.Suche in Google Scholar
[129] Y. Lin, Y. Ye, and W. Fang, “Electrically driven single-photon sources,” J. Semicond., vol. 40, no. 7, p. 071904, 2019. https://doi.org/10.1088/1674-4926/40/7/071904.Suche in Google Scholar
[130] M. J. Holmes, et al.., “Single photons from a hot solid-state emitter at 350 K,” ACS Photonics, vol. 3, no. 4, pp. 543–546, 2016. https://doi.org/10.1021/acsphotonics.6b00112.Suche in Google Scholar
[131] M. Munsch, et al.., “Dielectric GaAs antenna ensuring an efficient broadband coupling between an InAs quantum dot and a Gaussian optical beam,” Phys. Rev. Lett., vol. 110, no. 17, p. 177402, 2013. https://doi.org/10.1103/PhysRevLett.110.177402.Suche in Google Scholar PubMed
[132] M. Liu, et al.., “Colloidal quantum dot electronics,” Nat. Electron., vol. 4, no. 8, pp. 548–558, 2021. https://doi.org/10.1038/s41928-021-00632-7.Suche in Google Scholar
[133] Y. Yan, et al.., “Recent advances on graphene quantum dots: from chemistry and physics to applications,” Adv. Mater., vol. 31, no. 21, p. 1808283, 2019. https://doi.org/10.1002/adma.201808283.Suche in Google Scholar PubMed
[134] C. B. Murray, D. J. Norris, and M. G. Bawendi, “Synthesis and characterization of nearly monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites,” J. Am. Chem. Soc., vol. 115, no. 19, pp. 8706–8715, 1993. https://doi.org/10.1021/ja00072a025.Suche in Google Scholar
[135] V. Chandrasekaran, et al.., “Nearly blinking-free, high-purity single-photon emission by colloidal InP/ZnSe quantum dots,” Nano Lett., vol. 17, no. 10, pp. 6104–6109, 2017. https://doi.org/10.1021/acs.nanolett.7b02634.Suche in Google Scholar PubMed
[136] A. Jain, et al.., “Atomistic design of CdSe/CdS core–shell quantum dots with suppressed auger recombination,” Nano Lett., vol. 16, no. 10, pp. 6491–6496, 2016. https://doi.org/10.1021/acs.nanolett.6b03059.Suche in Google Scholar PubMed
[137] S. Krishnamurthy, et al.., “PbS/CdS quantum dot room-temperature single-emitter spectroscopy reaches the telecom O and S bands via an engineered stability,” ACS Nano, vol. 15, no. 1, pp. 575–587, 2021. https://doi.org/10.1021/acsnano.0c05907.Suche in Google Scholar PubMed
[138] J. Zou, et al.., “Perovskite quantum dots: synthesis, applications, prospects, and challenges,” J. Appl. Phys., vol. 132, no. 22, p. 220901, 2022. https://doi.org/10.1063/5.0126496.Suche in Google Scholar
[139] L. Protesescu, et al.., “Nanocrystals of cesium lead halide perovskites (CsPbX 3 , X = Cl, Br, and I): novel optoelectronic materials showing bright emission with wide color gamut,” Nano Lett., vol. 15, no. 6, pp. 3692–3696, 2015. https://doi.org/10.1021/nl5048779.Suche in Google Scholar PubMed PubMed Central
[140] J. Pan, et al.., “Highly efficient perovskite-quantum-dot light-emitting diodes by surface engineering,” Adv. Mater., vol. 28, no. 39, pp. 8718–8725, 2016. https://doi.org/10.1002/adma.201600784.Suche in Google Scholar PubMed
[141] F. Zhang, et al.., “Colloidal synthesis of air-stable CH 3 NH 3 PbI 3 quantum dots by gaining chemical insight into the solvent effects,” Chem. Mater., vol. 29, no. 8, pp. 3793–3799, 2017. https://doi.org/10.1021/acs.chemmater.7b01100.Suche in Google Scholar
[142] I. Levchuk, et al.., “Brightly luminescent and color-tunable formamidinium lead halide perovskite FAPbX3 (X = Cl, Br, I) colloidal nanocrystals,” Nano Lett., vol. 17, no. 5, pp. 2765–2770, 2017. https://doi.org/10.1021/acs.nanolett.6b04781.Suche in Google Scholar PubMed
[143] Y. Tong, et al.., “Highly luminescent cesium lead halide perovskite nanocrystals with tunable composition and thickness by ultrasonication,” Angew. Chem. Int. Ed., vol. 55, no. 44, pp. 13887–13892, 2016. https://doi.org/10.1002/anie.201605909.Suche in Google Scholar PubMed
[144] L. Rao, et al.., “Polar-solvent-free synthesis of highly photoluminescent and stable CsPbBr 3 nanocrystals with controlled shape and size by ultrasonication,” Chem. Mater., vol. 31, no. 2, pp. 365–375, 2019. https://doi.org/10.1021/acs.chemmater.8b03298.Suche in Google Scholar
[145] F. Liu, et al.., “Highly luminescent phase-stable CsPbI3 perovskite quantum dots achieving near 100% absolute photoluminescence quantum yield,” ACS Nano, vol. 11, no. 10, pp. 10373–10383, 2017. https://doi.org/10.1021/acsnano.7b05442.Suche in Google Scholar PubMed
[146] X. Tang, et al.., “Single halide perovskite/semiconductor core/shell quantum dots with ultrastability and nonblinking properties,” Adv. Sci., vol. 6, no. 18, p. 1900412, 2019. https://doi.org/10.1002/advs.201900412.Suche in Google Scholar PubMed PubMed Central
[147] 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.Suche in Google Scholar
[148] T. Miyazawa, et al.., “Single-photon emission at 1.5 μm from an InAs/InP quantum dot with highly suppressed multi-photon emission probabilities,” Appl. Phys. Lett., vol. 109, no. 13, p. 132106, 2016. https://doi.org/10.1063/1.4961888.Suche in Google Scholar
[149] C. Zhu, et al.., “Room-temperature, highly pure single-photon sources from all-inorganic lead halide perovskite quantum dots,” Nano Lett., vol. 22, no. 9, pp. 3751–3760, 2022. https://doi.org/10.1021/acs.nanolett.2c00756.Suche in Google Scholar PubMed PubMed Central
[150] A. E. K. Kaplan, et al.., “Hong–Ou–Mandel interference in colloidal CsPbBr3 perovskite nanocrystals,” Nat. Photonics, vol. 17, no. 9, pp. 775–780, 2023. https://doi.org/10.1038/s41566-023-01225-w.Suche in Google Scholar
[151] S. Jun, et al.., “Ultrafast and bright quantum emitters from the cavity-coupled single perovskite nanocrystals,” ACS Nano, vol. 18, no. 2, pp. 1396–1403, 2024. https://doi.org/10.1021/acsnano.3c06760.Suche in Google Scholar PubMed PubMed Central
[152] T. Farrow, et al.., “Ultranarrow line width room-temperature single-photon source from perovskite quantum dot embedded in optical microcavity,” Nano Lett., vol. 23, no. 23, pp. 10667–10673, 2023. https://doi.org/10.1021/acs.nanolett.3c02058.Suche in Google Scholar PubMed PubMed Central
[153] J. Wu, W. Pisula, and K. Müllen, “Graphenes as potential material for electronics,” Chem. Rev., vol. 107, no. 3, pp. 718–747, 2007. https://doi.org/10.1021/cr068010r.Suche in Google Scholar PubMed
[154] H. Arab, S. MohammadNejad, and P. MohammadNejad, “Se-doped NH2-functionalized graphene quantum dot for single-photon emission at free-space quantum communication wavelength,” Quantum Inf. Process., vol. 20, no. 5, p. 184, 2021. https://doi.org/10.1007/s11128-021-03122-z.Suche in Google Scholar
[155] A. Jasik, et al.., “The influence of the growth rate and V/III ratio on the crystal quality of InGaAs/GaAs QW structures grown by MBE and MOCVD methods,” J. Cryst. Growth, vol. 311, no. 19, pp. 4423–4432, 2009. https://doi.org/10.1016/j.jcrysgro.2009.07.032.Suche in Google Scholar
[156] J. Gérard, et al.., “Enhanced spontaneous emission by quantum boxes in a monolithic optical microcavity,” Phys. Rev. Lett., vol. 81, no. 5, pp. 1110–1113, 1998. https://doi.org/10.1103/PhysRevLett.81.1110.Suche in Google Scholar
[157] K. Hennessy, et al.., “Quantum nature of a strongly coupled single quantum dot–cavity system,” Nature, vol. 445, no. 7130, pp. 896–899, 2007. https://doi.org/10.1038/nature05586.Suche in Google Scholar PubMed
[158] M. J. Holmes, M. Arita, and Y. Arakawa, “III-nitride quantum dots as single photon emitters,” Semicond. Sci. Technol., vol. 34, no. 3, p. 33001, 2019. https://doi.org/10.1088/1361-6641/ab02c8.Suche in Google Scholar
[159] S. Bounouar, et al.., “Ultrafast room temperature single-photon source from nanowire-quantum dots,” Nano Lett., vol. 12, no. 6, pp. 2977–2981, 2012. https://doi.org/10.1021/nl300733f.Suche in Google Scholar PubMed
[160] O. Fedorych, et al.., “Room temperature single photon emission from an epitaxially grown quantum dot,” Appl. Phys. Lett., vol. 100, no. 6, p. 61114, 2012. https://doi.org/10.1063/1.3683498.Suche in Google Scholar
[161] W. Quitsch, et al.., “Electrically driven single photon emission from a CdSe/ZnSSe/MgS semiconductor quantum dot,” Phys. Status Solidi C, vol. 11, nos. 7–8, pp. 1256–1259, 2014. https://doi.org/10.1002/pssc.201300627.Suche in Google Scholar
[162] S. Deshpande, et al.., “Electrically driven polarized single-photon emission from an InGaN quantum dot in a GaN nanowire,” Nat. Commun., vol. 4, no. 1, p. 1675, 2013. https://doi.org/10.1038/ncomms2691.Suche in Google Scholar PubMed
[163] S. Deshpande, et al.., “Electrically pumped single-photon emission at room temperature from a single InGaN/GaN quantum dot,” Appl. Phys. Lett., vol. 105, no. 14, p. 141109, 2014. https://doi.org/10.1063/1.4897640.Suche in Google Scholar
[164] J.-H. Cho, et al.., “Strongly coherent single-photon emission from site-controlled InGaN quantum dots embedded in GaN nanopyramids,” ACS Photonics, vol. 5, no. 2, pp. 439–444, 2018. https://doi.org/10.1021/acsphotonics.7b00922.Suche in Google Scholar
[165] Y.-M. He, et al.., “On-demand semiconductor single-photon source with near-unity indistinguishability,” Nat. Nanotechnol., vol. 8, no. 3, pp. 213–217, 2013. https://doi.org/10.1038/nnano.2012.262.Suche in Google Scholar PubMed
[166] L. Schweickert, et al.., “On-demand generation of background-free single photons from a solid-state source,” Appl. Phys. Lett., vol. 112, no. 9, p. 93106, 2018. https://doi.org/10.1063/1.5020038.Suche in Google Scholar
[167] M. Paul, et al.., “Single-photon emission at 1.55 μm from MOVPE-grown InAs quantum dots on InGaAs/GaAs metamorphic buffers,” Appl. Phys. Lett., vol. 111, no. 3, p. 033102, 2017. https://doi.org/10.1063/1.4993935.Suche in Google Scholar
[168] K. Takemoto, et al.., “An optical horn structure for single-photon source using quantum dots at telecommunication wavelength,” J. Appl. Phys., vol. 101, no. 8, p. 81720, 2007. https://doi.org/10.1063/1.2723177.Suche in Google Scholar
[169] T. Müller, et al.., “A quantum light-emitting diode for the standard telecom window around 1,550 nm,” Nat. Commun., vol. 9, no. 1, p. 862, 2018. https://doi.org/10.1038/s41467-018-03251-7.Suche in Google Scholar PubMed PubMed Central
[170] C. Nawrath, et al.., “Coherence and indistinguishability of highly pure single photons from non-resonantly and resonantly excited telecom C-band quantum dots,” Appl. Phys. Lett., vol. 115, no. 2, p. 23103, 2019. https://doi.org/10.1063/1.5095196.Suche in Google Scholar
[171] M. A. Pooley, et al.., “Controlled-NOT gate operating with single photons,” Appl. Phys. Lett., vol. 100, no. 21, p. 211103, 2012. https://doi.org/10.1063/1.4719077.Suche in Google Scholar
[172] D. Golberg, et al.., “Boron nitride nanotubes and nanosheets,” ACS Nano, vol. 4, no. 6, pp. 2979–2993, 2010. https://doi.org/10.1021/nn1006495.Suche in Google Scholar PubMed
[173] N. Chejanovsky, et al.., “Quantum light in curved low dimensional hexagonal boron nitride systems,” Sci. Rep., vol. 7, no. 1, p. 14758, 2017. https://doi.org/10.1038/s41598-017-15398-2.Suche in Google Scholar PubMed PubMed Central
[174] J. Ahn, et al.., “Stable emission and fast optical modulation of quantum emitters in boron nitride nanotubes,” Opt. Lett., vol. 43, no. 15, p. 3778, 2018. https://doi.org/10.1364/OL.43.003778.Suche in Google Scholar PubMed
[175] X. Gao, et al.., “Nanotube spin defects for omnidirectional magnetic field sensing,” Nat. Commun., vol. 15, no. 1, p. 7697, 2024. https://doi.org/10.1038/s41467-024-51941-2.Suche in Google Scholar PubMed PubMed Central
[176] X. He, et al.., “Carbon nanotubes as emerging quantum-light sources,” Nat. Mater., vol. 17, no. 8, pp. 663–670, 2018. https://doi.org/10.1038/s41563-018-0109-2.Suche in Google Scholar PubMed
[177] J. Gao, et al.., “Selective wrapping and supramolecular structures of polyfluorene–carbon nanotube hybrids,” ACS Nano, vol. 5, no. 5, pp. 3993–3999, 2011. https://doi.org/10.1021/nn200564n.Suche in Google Scholar PubMed
[178] S. Khasminskaya, et al.., “Fully integrated quantum photonic circuit with an electrically driven light source,” Nat. Photonics, vol. 10, no. 11, pp. 727–732, 2016. https://doi.org/10.1038/nphoton.2016.178.Suche in Google Scholar
[179] C. Raynaud, et al.., “Superlocalization of excitons in carbon nanotubes at cryogenic temperature,” Nano Lett., vol. 19, no. 10, pp. 7210–7216, 2019. https://doi.org/10.1021/acs.nanolett.9b02816.Suche in Google Scholar PubMed
[180] Z. Li, et al.., “Quantum emission assisted by energy landscape modification in pentacene-decorated carbon nanotubes,” ACS Photonics, vol. 8, no. 8, pp. 2367–2374, 2021. https://doi.org/10.1021/acsphotonics.1c00539.Suche in Google Scholar
[181] A. Högele, et al.., “Photon antibunching in the photoluminescence spectra of a single carbon nanotube,” Phys. Rev. Lett., vol. 100, no. 21, p. 217401, 2008. https://doi.org/10.1103/PhysRevLett.100.217401.Suche in Google Scholar PubMed
[182] Y. Zheng, et al.., “Quantum light emission from coupled defect states in DNA-functionalized carbon nanotubes,” ACS Nano, vol. 15, no. 6, pp. 10406–10414, 2021. https://doi.org/10.1021/acsnano.1c02709.Suche in Google Scholar PubMed
[183] T. Hertel, et al.., “Diffusion limited photoluminescence quantum yields in 1-D semiconductors: single-wall carbon nanotubes,” ACS Nano, vol. 4, no. 12, pp. 7161–7168, 2010. https://doi.org/10.1021/nn101612b.Suche in Google Scholar PubMed
[184] F. Hayee, et al.., “Revealing multiple classes of stable quantum emitters in hexagonal boron nitride with correlated optical and electron microscopy,” Nat. Mater., vol. 19, no. 5, pp. 534–539, 2020. https://doi.org/10.1038/s41563-020-0616-9.Suche in Google Scholar PubMed
[185] C. Fournier, et al.., “Position-controlled quantum emitters with reproducible emission wavelength in hexagonal boron nitride,” Nat. Commun., vol. 12, no. 1, p. 3779, 2021. https://doi.org/10.1038/s41467-021-24019-6.Suche in Google Scholar PubMed PubMed Central
[186] C. Fournier, et al.., “Two-photon interference from a quantum emitter in hexagonal boron nitride,” Phys. Rev. Appl., vol. 19, no. 4, p. L041003, 2023. https://doi.org/10.1103/PhysRevApplied.19.L041003.Suche in Google Scholar
[187] T. T. Tran, et al.., “Room-temperature single-photon emission from oxidized tungsten disulfide multilayers,” Adv. Opt. Mater., vol. 5, no. 5, p. 1600939, 2017. https://doi.org/10.1002/adom.201600939.Suche in Google Scholar
[188] H. Ngoc My Duong, et al.., “Effects of high-energy electron irradiation on quantum emitters in hexagonal boron nitride,” ACS Appl. Mater. Interfaces, vol. 10, no. 29, pp. 24886–24891, 2018. https://doi.org/10.1021/acsami.8b07506.Suche in Google Scholar PubMed
[189] J. Ziegler, et al.., “Deterministic quantum emitter formation in hexagonal boron nitride via controlled edge creation,” Nano Lett., vol. 19, no. 3, pp. 2121–2127, 2019. https://doi.org/10.1021/acs.nanolett.9b00357.Suche in Google Scholar PubMed
[190] N. V. Proscia, et al.., “Near-deterministic activation of room-temperature quantum emitters in hexagonal boron nitride,” Optica, vol. 5, no. 9, p. 1128, 2018. https://doi.org/10.1364/OPTICA.5.001128.Suche in Google Scholar
[191] C. Li, et al.., “Scalable and deterministic fabrication of quantum emitter arrays from hexagonal boron nitride,” Nano Lett., vol. 21, no. 8, pp. 3626–3632, 2021. https://doi.org/10.1021/acs.nanolett.1c00685.Suche in Google Scholar PubMed
[192] S. Castelletto, et al.., “Color centers enabled by direct femto-second laser writing in wide bandgap semiconductors,” Nanomaterials, vol. 11, no. 1, p. 72, 2020. https://doi.org/10.3390/nano11010072.Suche in Google Scholar PubMed PubMed Central
[193] S. A. Tawfik, et al.., “First-principles investigation of quantum emission from hBN defects,” Nanoscale, vol. 9, no. 36, pp. 13575–13582, 2017. https://doi.org/10.1039/C7NR04270A.Suche in Google Scholar PubMed
[194] P. Auburger and A. Gali, “Towards ab initio identification of paramagnetic substitutional carbon defects in hexagonal boron nitride acting as quantum bits,” Phys. Rev. B, vol. 104, no. 7, p. 75410, 2021. https://doi.org/10.1103/PhysRevB.104.075410.Suche in Google Scholar
[195] Q. Tan, et al.., “Donor–acceptor pair quantum emitters in hexagonal boron nitride,” Nano Lett., vol. 22, no. 3, pp. 1331–1337, 2022. https://doi.org/10.1021/acs.nanolett.1c04647.Suche in Google Scholar PubMed
[196] Á. Ganyecz, et al.., “First-principles theory of the nitrogen interstitial in hBN: a plausible model for the blue emitter,” Nanoscale, vol. 16, no. 8, pp. 4125–4139, 2024. https://doi.org/10.1039/D3NR05811E.Suche in Google Scholar
[197] C. Jara, et al.., “First-principles identification of single photon emitters based on carbon clusters in hexagonal boron nitride,” J. Phys. Chem. A, vol. 125, no. 6, pp. 1325–1335, 2021. https://doi.org/10.1021/acs.jpca.0c07339.Suche in Google Scholar PubMed
[198] O. Golami, et al.., “A b i n i t i o and group theoretical study of properties of a carbon trimer defect in hexagonal boron nitride,” Phys. Rev. B, vol. 105, no. 18, p. 184101, 2022. https://doi.org/10.1103/PhysRevB.105.184101.Suche in Google Scholar
[199] S. Li, et al.., “Ultraviolet quantum emitters in hexagonal boron nitride from carbon clusters,” J. Phys. Chem. Lett., vol. 13, no. 14, pp. 3150–3157, 2022. https://doi.org/10.1021/acs.jpclett.2c00665.Suche in Google Scholar PubMed PubMed Central
[200] J. Horder, et al.., “Coherence properties of electron-beam-activated emitters in hexagonal boron nitride under resonant excitation,” Phys. Rev. Appl., vol. 18, no. 6, p. 064021, 2022. https://doi.org/10.1103/PhysRevApplied.18.064021.Suche in Google Scholar
[201] T. Vogl, et al.., “Compact cavity-enhanced single-photon generation with hexagonal boron nitride,” ACS Photonics, vol. 6, no. 8, pp. 1955–1962, 2019. https://doi.org/10.1021/acsphotonics.9b00314.Suche in Google Scholar
[202] S. J. U. White, et al.., “Quantum random number generation using a hexagonal boron nitride single photon emitter,” J. Opt., vol. 23, no. 1, p. 1LT01, 2021. https://doi.org/10.1088/2040-8986/abccff.Suche in Google Scholar
[203] D. Scognamiglio, et al.., “On-demand quantum light sources for underwater communications,” Mater. Quantum Technol., vol. 4, no. 2, p. 25402, 2024. https://doi.org/10.1088/2633-4356/ad46d7.Suche in Google Scholar
[204] S. Vaidya, et al.., “Quantum sensing and imaging with spin defects in hexagonal boron nitride,” Adv. Phys. X, vol. 8, no. 1, p. 2206049, 2023. https://doi.org/10.1080/23746149.2023.2206049.Suche in Google Scholar
[205] V. M. Acosta, et al.., “Temperature dependence of the nitrogen-vacancy magnetic resonance in diamond,” Phys. Rev. Lett., vol. 104, no. 7, p. 70801, 2010. https://doi.org/10.1103/PhysRevLett.104.070801.Suche in Google Scholar PubMed
[206] M. W. Doherty, et al.., “Temperature shifts of the resonances of the NV − center in diamond,” Phys. Rev. B, vol. 90, no. 4, p. 41201, 2014. https://doi.org/10.1103/PhysRevB.90.041201.Suche in Google Scholar
[207] A. Gottscholl, et al.., “Initialization and read-out of intrinsic spin defects in a van der Waals crystal at room temperature,” Nat. Mater., vol. 19, no. 5, pp. 540–545, 2020. https://doi.org/10.1038/s41563-020-0619-6.Suche in Google Scholar PubMed
[208] W. Liu, et al.., “Temperature-dependent energy-level shifts of spin defects in hexagonal boron nitride,” ACS Photonics, vol. 8, no. 7, pp. 1889–1895, 2021. https://doi.org/10.1021/acsphotonics.1c00320.Suche in Google Scholar
[209] N. Mathur, et al.., “Excited-state spin-resonance spectroscopy of VB− defect centers in hexagonal boron nitride,” Nat. Commun., vol. 13, no. 1, p. 3233, 2022. https://doi.org/10.1038/s41467-022-30772-z.Suche in Google Scholar PubMed PubMed Central
[210] A. Gottscholl, et al.., “Spin defects in hBN as promising temperature, pressure and magnetic field quantum sensors,” Nat. Commun., vol. 12, no. 1, p. 4480, 2021. https://doi.org/10.1038/s41467-021-24725-1.Suche in Google Scholar PubMed PubMed Central
[211] M. Huang, et al.., “Wide field imaging of van der Waals ferromagnet Fe3GeTe2 by spin defects in hexagonal boron nitride,” Nat. Commun., vol. 13, no. 1, p. 5369, 2022. https://doi.org/10.1038/s41467-022-33016-2.Suche in Google Scholar PubMed PubMed Central
[212] X. Gao, et al.., “Quantum sensing of paramagnetic spins in liquids with spin qubits in hexagonal boron nitride,” ACS Photonics, vol. 10, no. 8, pp. 2894–2900, 2023. https://doi.org/10.1021/acsphotonics.3c00621.Suche in Google Scholar
[213] N.-J. Guo, et al.., “Coherent control of an ultrabright single spin in hexagonal boron nitride at room temperature,” Nat. Commun., vol. 14, no. 1, p. 2893, 2023. https://doi.org/10.1038/s41467-023-38672-6.Suche in Google Scholar PubMed PubMed Central
[214] H. L. Stern, et al.., “A quantum coherent spin in hexagonal boron nitride at ambient conditions,” Nat. Mater., vol. 23, no. 10, pp. 1379–1385, 2024. https://doi.org/10.1038/s41563-024-01887-z.Suche in Google Scholar PubMed PubMed Central
[215] A. J. Healey, et al.., “Quantum microscopy with van der Waals heterostructures,” Nat. Phys., vol. 19, no. 1, pp. 87–91, 2023. https://doi.org/10.1038/s41567-022-01815-5.Suche in Google Scholar
[216] B. C. Cavenett, “Optically detected magnetic resonance (O.D.M.R.) investigations of recombination processes in semiconductors,” Adv. Phys., vol. 30, no. 4, pp. 475–538, 1981. https://doi.org/10.1080/00018738100101397.Suche in Google Scholar
[217] P. Yu, et al.., “Excited-state spectroscopy of spin defects in hexagonal boron nitride,” Nano Lett., vol. 22, no. 9, pp. 3545–3549, 2022. https://doi.org/10.1021/acs.nanolett.1c04841.Suche in Google Scholar PubMed
[218] P. Khatri, et al.., “Optical gating of photoluminescence from color centers in hexagonal boron nitride,” Nano Lett., vol. 20, no. 6, pp. 4256–4263, 2020. https://doi.org/10.1021/acs.nanolett.0c00751.Suche in Google Scholar PubMed PubMed Central
[219] N. Mendelson, et al.., “Identifying carbon as the source of visible single-photon emission from hexagonal boron nitride,” Nat. Mater., vol. 20, no. 3, pp. 321–328, 2021. https://doi.org/10.1038/s41563-020-00850-y.Suche in Google Scholar PubMed
[220] A. Dietrich, et al.., “Solid-state single photon source with Fourier transform limited lines at room temperature,” Phys. Rev. B, vol. 101, no. 8, p. 081401, 2020. https://doi.org/10.1103/PhysRevB.101.081401.Suche in Google Scholar
[221] J. E. Fröch, et al.., “Coupling hexagonal boron nitride quantum emitters to photonic crystal cavities,” ACS Nano, vol. 14, no. 6, pp. 7085–7091, 2020. https://doi.org/10.1021/acsnano.0c01818.Suche in Google Scholar PubMed
[222] H. L. Stern, et al.., “Room-temperature optically detected magnetic resonance of single defects in hexagonal boron nitride,” Nat. Commun., vol. 13, no. 1, p. 618, 2022. https://doi.org/10.1038/s41467-022-28169-z.Suche in Google Scholar PubMed PubMed Central
[223] N. Chejanovsky, et al.., “Single-spin resonance in a van der Waals embedded paramagnetic defect,” Nat. Mater., vol. 20, no. 8, pp. 1079–1084, 2021. https://doi.org/10.1038/s41563-021-00979-4.Suche in Google Scholar PubMed
[224] S. C. Scholten, et al.., “Multi-species optically addressable spin defects in a van der Waals material,” Nat. Commun., vol. 15, no. 1, p. 6727, 2024. https://doi.org/10.1038/s41467-024-51129-8.Suche in Google Scholar PubMed PubMed Central
[225] K. Sasaki, et al.., “Magnetic field imaging by hBN quantum sensor nanoarray,” Appl. Phys. Lett., vol. 122, no. 24, p. 244003, 2023. https://doi.org/10.1063/5.0147072.Suche in Google Scholar
[226] J. S. Moon, et al.., “Fiber-integrated van der Waals quantum sensor with an optimal cavity interface,” Adv. Opt. Mater., no. 32, p. 2401987, 2024. https://doi.org/10.1002/adom.202401987.Suche in Google Scholar
[227] T. Kosmala, et al.., “Strain induced phase transition of WS2 by local dewetting of Au/mica film upon annealing,” Surfaces, vol. 4, no. 1, pp. 1–8, 2020. https://doi.org/10.3390/surfaces4010001.Suche in Google Scholar
[228] J. Kern, et al.., “Nanoscale positioning of single-photon emitters in atomically thin WSe2,” Adv. Mater., vol. 28, no. 33, pp. 7101–7105, 2016. https://doi.org/10.1002/adma.201600560.Suche in Google Scholar PubMed
[229] K. L. Seyler, et al.., “Signatures of moiré-trapped valley excitons in MoSe2/WSe2 heterobilayers,” Nature, vol. 567, no. 7746, pp. 66–70, 2019. https://doi.org/10.1038/s41586-019-0957-1.Suche in Google Scholar PubMed
[230] Y. Luo, et al.., “Deterministic coupling of site-controlled quantum emitters in monolayer WSe2 to plasmonic nanocavities,” Nat. Nanotech., vol. 13, no. 12, pp. 1137–1142, 2018. https://doi.org/10.1038/s41565-018-0275-z.Suche in Google Scholar PubMed
[231] M. Von Helversen, et al.., “Temperature dependent temporal coherence of metallic-nanoparticle-induced single-photon emitters in a WSe2 monolayer,” 2D Mater., vol. 10, no. 4, p. 045034, 2023. https://doi.org/10.1088/2053-1583/acfb20.Suche in Google Scholar
[232] J.-C. Drawer, et al.., “Monolayer-based single-photon source in a liquid-helium-free open cavity featuring 65% brightness and quantum coherence,” Nano Lett., vol. 23, no. 18, pp. 8683–8689, 2023. https://doi.org/10.1021/acs.nanolett.3c02584.Suche in Google Scholar PubMed PubMed Central
[233] K. F. Mak, et al.., “Atomically thin MoS2: a new direct-gap semiconductor,” Phys. Rev. Lett., vol. 105, no. 13, p. 136805, 2010. https://doi.org/10.1103/PhysRevLett.105.136805.Suche in Google Scholar PubMed
[234] A. Splendiani, et al.., “Emerging photoluminescence in monolayer MoS2,” Nano Lett., vol. 10, no. 4, pp. 1271–1275, 2010. https://doi.org/10.1021/nl903868w.Suche in Google Scholar PubMed
[235] B. Zhu, X. Chen, and X. Cui, “Exciton binding energy of monolayer WS2,” Sci. Rep., vol. 5, no. 1, p. 9218, 2015. https://doi.org/10.1038/srep09218.Suche in Google Scholar PubMed PubMed Central
[236] S. Park, et al.., “Direct determination of monolayer MoS2 and WSe2 exciton binding energies on insulating and metallic substrates,” 2D Mater., vol. 5, no. 2, p. 25003, 2018. https://doi.org/10.1088/2053-1583/aaa4ca.Suche in Google Scholar
[237] X. Lu, et al.., “Optical initialization of a single spin-valley in charged WSe2 quantum dots,” Nat. Nanotechnol., vol. 14, no. 5, pp. 426–431, 2019. https://doi.org/10.1038/s41565-019-0394-1.Suche in Google Scholar PubMed
[238] S. Zhang, et al.., “Defect structure of localized excitons in a WSe2 monolayer,” Phys. Rev. Lett., vol. 119, no. 4, p. 46101, 2017. https://doi.org/10.1103/PhysRevLett.119.046101.Suche in Google Scholar PubMed
[239] G. D. Shepard, et al.., “Nanobubble induced formation of quantum emitters in monolayer semiconductors,” 2D Mater., vol. 4, no. 2, p. 021019, 2017. https://doi.org/10.1088/2053-1583/aa629d.Suche in Google Scholar
[240] S. Cianci, et al.., “Spatially controlled single photon emitters in hBN-capped WS2 domes,” Adv. Opt. Mater., vol. 11, no. 12, p. 2202953, 2023. https://doi.org/10.1002/adom.202202953.Suche in Google Scholar
[241] C. Palacios-Berraquero, et al.., “Large-scale quantum-emitter arrays in atomically thin semiconductors,” Nat. Commun., vol. 8, no. 1, p. 15093, 2017. https://doi.org/10.1038/ncomms15093.Suche in Google Scholar PubMed PubMed Central
[242] W. Wu, et al.., “Locally defined quantum emission from epitaxial few-layer tungsten diselenide,” Appl. Phys. Lett., vol. 114, no. 21, p. 213102, 2019. https://doi.org/10.1063/1.5091779.Suche in Google Scholar
[243] M. R. Rosenberger, et al.., “Quantum calligraphy: writing single-photon emitters in a two-dimensional materials platform,” ACS Nano, vol. 13, no. 1, pp. 904–912, 2019. https://doi.org/10.1021/acsnano.8b08730.Suche in Google Scholar PubMed
[244] J.-P. So, et al.., “Electrically driven strain-induced deterministic single-photon emitters in a van der Waals heterostructure,” Sci. Adv., vol. 7, no. 43, p. eabj3176, 2021. https://doi.org/10.1126/sciadv.abj3176.Suche in Google Scholar PubMed PubMed Central
[245] G. Moody, et al.., “Microsecond Valley lifetime of defect-bound excitons in monolayer WSe2,” Phys. Rev. Lett., vol. 121, no. 5, p. 057403, 2018. https://doi.org/10.1103/PhysRevLett.121.057403.Suche in Google Scholar PubMed
[246] K. Parto, et al.., “Defect and strain engineering of monolayer WSe2 enables site-controlled single-photon emission up to 150 K,” Nat. Commun., vol. 12, no. 1, p. 3585, 2021. https://doi.org/10.1038/s41467-021-23709-5.Suche in Google Scholar PubMed PubMed Central
[247] C. Palacios-Berraquero, et al.., “Atomically thin quantum light-emitting diodes,” Nat. Commun., vol. 7, no. 1, p. 12978, 2016. https://doi.org/10.1038/ncomms12978.Suche in Google Scholar PubMed PubMed Central
[248] J. Klein, et al.., “Site-selectively generated photon emitters in monolayer MoS2 via local helium ion irradiation,” Nat. Commun., vol. 10, no. 1, p. 2755, 2019. https://doi.org/10.1038/s41467-019-10632-z.Suche in Google Scholar PubMed PubMed Central
[249] K. Barthelmi, et al.., “Atomistic defects as single-photon emitters in atomically thin MoS2,” Appl. Phys. Lett., vol. 117, no. 7, p. 070501, 2020. https://doi.org/10.1063/5.0018557.Suche in Google Scholar
[250] J. Klein, et al.., “Engineering the luminescence and generation of individual defect emitters in atomically thin MoS2,” ACS Photonics, vol. 8, no. 2, pp. 669–677, 2021. https://doi.org/10.1021/acsphotonics.0c01907.Suche in Google Scholar
[251] A. Branny, et al.., “Discrete quantum dot like emitters in monolayer MoSe2: spatial mapping, magneto-optics, and charge tuning,” Appl. Phys. Lett., vol. 108, no. 14, p. 142101, 2016. https://doi.org/10.1063/1.4945268.Suche in Google Scholar
[252] L. Yu, et al.., “Site-controlled quantum emitters in monolayer MoSe2,” Nano Lett., vol. 21, no. 6, pp. 2376–2381, 2021. https://doi.org/10.1021/acs.nanolett.0c04282.Suche in Google Scholar PubMed
[253] H. Zhao, et al.., “Manipulating interlayer excitons for near-infrared quantum light generation,” Nano Lett., vol. 23, no. 23, pp. 11006–11012, 2023. https://doi.org/10.1021/acs.nanolett.3c03296.Suche in Google Scholar PubMed
[254] F. He, et al.., “Moiré patterns in 2D materials: a review,” ACS Nano, vol. 15, no. 4, pp. 5944–5958, 2021. https://doi.org/10.1021/acsnano.0c10435.Suche in Google Scholar PubMed
[255] H. Baek, et al.., “Highly energy-tunable quantum light from moiré-trapped excitons,” Sci. Adv., vol. 6, no. 37, p. eaba8526, 2020. https://doi.org/10.1126/sciadv.aba8526.Suche in Google Scholar PubMed PubMed Central
[256] H.-J. Chuang, et al.., “Enhancing single photon emission purity via design of van der Waals heterostructures,” Nano Lett., vol. 24, no. 18, pp. 5529–5535, 2024. https://doi.org/10.1021/acs.nanolett.4c00683.Suche in Google Scholar PubMed
[257] L. Sortino, et al.., “Bright single photon emitters with enhanced quantum efficiency in a two-dimensional semiconductor coupled with dielectric nano-antennas,” Nat. Commun., vol. 12, no. 1, p. 6063, 2021. https://doi.org/10.1038/s41467-021-26262-3.Suche in Google Scholar PubMed PubMed Central
[258] T. Gao, et al.., “Atomically-thin single-photon sources for quantum communication,” Npj 2D Mater. Appl., vol. 7, no. 1, p. 4, 2023. https://doi.org/10.1038/s41699-023-00366-4.Suche in Google Scholar
[259] N. P. Wilson, et al.., “Excitons and emergent quantum phenomena in stacked 2D semiconductors,” Nature, vol. 599, no. 7885, pp. 383–392, 2021. https://doi.org/10.1038/s41586-021-03979-1.Suche in Google Scholar PubMed
[260] G. Grosso, et al.., “Tunable and high-purity room temperature single-photon emission from atomic defects in hexagonal boron nitride,” Nat. Commun., vol. 8, no. 1, p. 705, 2017. https://doi.org/10.1038/s41467-017-00810-2.Suche in Google Scholar PubMed PubMed Central
[261] Y. Xue, et al.., “Anomalous pressure characteristics of defects in hexagonal boron nitride flakes,” ACS Nano, vol. 12, no. 7, pp. 7127–7133, 2018. https://doi.org/10.1021/acsnano.8b02970.Suche in Google Scholar PubMed
[262] O. Iff, et al.., “Strain-tunable single photon sources in WSe2 monolayers,” Nano Lett., vol. 19, no. 10, pp. 6931–6936, 2019. https://doi.org/10.1021/acs.nanolett.9b02221.Suche in Google Scholar PubMed
[263] A. Ciarrocchi, et al.., “Polarization switching and electrical control of interlayer excitons in two-dimensional van der Waals heterostructures,” Nat. Photonics, vol. 13, no. 2, pp. 131–136, 2019. https://doi.org/10.1038/s41566-018-0325-y].10.1038/s41566-018-0325-ySuche in Google Scholar PubMed PubMed Central
[264] M. I. B. Utama, et al.., “Chemomechanical modification of quantum emission in monolayer WSe2,” Nat. Commun., vol. 14, no. 1, p. 2193, 2023. https://doi.org/10.1038/s41467-023-37892-0.Suche in Google Scholar PubMed PubMed Central
[265] G. Noh, et al.., “Stark tuning of single-photon emitters in hexagonal boron nitride,” Nano Lett., vol. 18, no. 8, pp. 4710–4715, 2018. https://doi.org/10.1021/acs.nanolett.8b01030.Suche in Google Scholar PubMed
[266] C. Chakraborty, et al.., “Quantum-confined stark effect of individual defects in a van der Waals heterostructure,” Nano Lett., vol. 17, no. 4, pp. 2253–2258, 2017. https://doi.org/10.1021/acs.nanolett.6b04889.Suche in Google Scholar PubMed
[267] N. Nikolay, et al.., “Very large and reversible stark-shift tuning of single emitters in layered hexagonal boron nitride,” Phys. Rev. Appl., vol. 11, no. 4, p. 41001, 2019. https://doi.org/10.1103/PhysRevApplied.11.041001.Suche in Google Scholar
[268] I. Zhigulin, et al.., “Stark effect of blue quantum emitters in hexagonal boron nitride,” Phys. Rev. Appl., vol. 19, no. 4, p. 44011, 2023. https://doi.org/10.1103/PhysRevApplied.19.044011.Suche in Google Scholar
[269] T. Grange, et al.., “Cavity-funneled generation of indistinguishable single photons from strongly dissipative quantum emitters,” Phys. Rev. Lett., vol. 114, no. 19, p. 193601, 2015. https://doi.org/10.1103/PhysRevLett.114.193601.Suche in Google Scholar PubMed
[270] J.-H. Kim, et al.., “Two-photon interference from a bright single-photon source at telecom wavelengths,” Optica, vol. 3, no. 6, p. 577, 2016. https://doi.org/10.1364/OPTICA.3.000577.Suche in Google Scholar
[271] A. Jeantet, et al.., “Exploiting one-dimensional exciton–phonon coupling for tunable and efficient single-photon generation with a carbon nanotube,” Nano Lett., vol. 17, no. 7, pp. 4184–4188, 2017. https://doi.org/10.1021/acs.nanolett.7b00973.Suche in Google Scholar PubMed
[272] T. T. Tran, et al.., “Deterministic coupling of quantum emitters in 2D materials to plasmonic nanocavity arrays,” Nano Lett., vol. 17, no. 4, pp. 2634–2639, 2017. https://doi.org/10.1021/acs.nanolett.7b00444.Suche in Google Scholar PubMed
[273] E. T. Jaynes and F. W. Cummings, “Comparison of quantum and semiclassical radiation theories with application to the beam maser,” Proc. IEEE, vol. 51, no. 1, pp. 89–109, 1963. https://doi.org/10.1109/PROC.1963.1664.Suche in Google Scholar
[274] C. Zhang, et al.., “Microstructure engineering of hexagonal boron nitride for single-photon emitter applications,” Adv. Opt. Mater., vol. 10, no. 17, p. 2200207, 2022. https://doi.org/10.1002/adom.202200207.Suche in Google Scholar
[275] X. Ding, et al.., “High-efficiency single-photon source above the loss-tolerant threshold for efficient linear optical quantum computing,” arXiv:2311.08347, arXiv, 2023.Suche in Google Scholar
[276] A. Jeantet, et al.., “Widely tunable single-photon source from a carbon nanotube in the Purcell regime,” Phys. Rev. Lett., vol. 116, no. 24, p. 247402, 2016. https://doi.org/10.1103/PhysRevLett.116.247402.Suche in Google Scholar PubMed
[277] D. Hunger, et al.., “A fiber Fabry–Perot cavity with high finesse,” New J. Phys., vol. 12, no. 6, p. 065038, 2010. https://doi.org/10.1088/1367-2630/12/6/065038.Suche in Google Scholar
[278] T. Cai, et al.., “Coupling emission from single localized defects in two-dimensional semiconductor to surface plasmon polaritons,” Nano Lett., vol. 17, no. 11, pp. 6564–6568, 2017. https://doi.org/10.1021/acs.nanolett.7b02222.Suche in Google Scholar PubMed
[279] L. C. Flatten, et al.., “Microcavity enhanced single photon emission from two-dimensional WSe2,” Appl. Phys. Lett., vol. 112, no. 19, p. 191105, 2018. https://doi.org/10.1063/1.5026779.Suche in Google Scholar
[280] C. Errando-Herranz, et al.., “Resonance fluorescence from waveguide-coupled, strain-localized, two-dimensional quantum emitters,” ACS Photonics, vol. 8, no. 4, pp. 1069–1076, 2021. https://doi.org/10.1021/acsphotonics.0c01653.Suche in Google Scholar PubMed PubMed Central
[281] Y.-C. Chen, et al.., “Laser writing of individual nitrogen-vacancy defects in diamond with near-unity yield,” Optica, vol. 6, no. 5, p. 662, 2019. https://doi.org/10.1364/OPTICA.6.000662.Suche in Google Scholar
[282] D. A. Bandurin, et al.., “High electron mobility, quantum Hall effect and anomalous optical response in atomically thin InSe,” Nat. Nanotech., vol. 12, no. 3, pp. 223–227, 2017. https://doi.org/10.1038/nnano.2016.242.Suche in Google Scholar PubMed
[283] T. V. Shubina, et al.., “InSe as a case between 3D and 2D layered crystals for excitons,” Nat. Commun., vol. 10, no. 1, p. 3479, 2019. https://doi.org/10.1038/s41467-019-11487-0.Suche in Google Scholar PubMed PubMed Central
[284] M. Salomone, et al.., “Point defects in two-dimensional indium selenide as tunable single-photon sources,” J. Phys. Chem. Lett., vol. 12, no. 45, pp. 10947–10952, 2021. https://doi.org/10.1021/acs.jpclett.1c02912.Suche in Google Scholar PubMed PubMed Central
[285] W. Luo, et al.., “Deterministic localization of strain-induced single-photon emitters in multilayer GaSe,” ACS Photonics, vol. 10, no. 8, pp. 2530–2539, 2023. https://doi.org/10.1021/acsphotonics.3c00052.Suche in Google Scholar
[286] W. Luo, et al.., “Improving strain-localized GaSe single photon emitters with electrical doping,” Nano Lett., vol. 23, no. 21, pp. 9740–9747, 2023. https://doi.org/10.1021/acs.nanolett.3c02308.Suche in Google Scholar PubMed
[287] D. G. Hopkinson, et al.., “Formation and healing of defects in atomically thin GaSe and InSe,” ACS Nano, vol. 13, no. 5, pp. 5112–5123, 2019. https://doi.org/10.1021/acsnano.8b08253.Suche in Google Scholar PubMed
[288] A. S. Sarkar and E. Stratakis, “Recent advances in 2D metal monochalcogenides,” Adv. Sci., vol. 7, no. 21, p. 2001655, 2020. https://doi.org/10.1002/advs.202001655.Suche in Google Scholar PubMed PubMed Central
[289] Z. Hu, et al.., “Recent progress in 2D group IV–IV monochalcogenides: synthesis, properties and applications,” Proc. SPIE, vol. 30, no. 25, p. 252001, 2019. https://doi.org/10.1088/1361-6528/ab07d9.Suche in Google Scholar PubMed
[290] G. W. Mudd, et al.., “The direct-to-indirect band gap crossover in two-dimensional van der Waals indium selenide crystals,” Sci. Rep., vol. 6, no. 1, p. 39619, 2016. https://doi.org/10.1038/srep39619.Suche in Google Scholar PubMed PubMed Central
[291] P. Tonndorf, et al.., “On-chip waveguide coupling of a layered semiconductor single-photon source,” Nano Lett., vol. 17, no. 9, pp. 5446–5451, 2017. https://doi.org/10.1021/acs.nanolett.7b02092.Suche in Google Scholar PubMed
[292] D. Jariwala, T. J. Marks, and M. C. Hersam, “Mixed-dimensional van der Waals heterostructures,” Nat. Mater., vol. 16, no. 2, pp. 170–181, 2017. https://doi.org/10.1038/nmat4703.Suche in Google Scholar PubMed
[293] O. Lopez-Sanchez, et al.., “Light generation and harvesting in a van der Waals heterostructure,” ACS Nano, vol. 8, no. 3, pp. 3042–3048, 2014. https://doi.org/10.1021/nn500480u.Suche in Google Scholar PubMed PubMed Central
[294] L. Dou, et al.., “Atomically thin two-dimensional organic-inorganic hybrid perovskites,” Science, vol. 349, no. 6255, pp. 1518–1521, 2015. https://doi.org/10.1126/science.aac7660.Suche in Google Scholar PubMed
[295] X. Li, et al.., “Proximity-induced chiral quantum light generation in strain-engineered WSe2/NiPS3 heterostructures,” Nat. Mater., vol. 22, no. 11, pp. 1311–1316, 2023. https://doi.org/10.1038/s41563-023-01645-7.Suche in Google Scholar PubMed
[296] K. Azuma, et al.., “Quantum repeaters: from quantum networks to the quantum internet,” Rev. Mod. Phys., vol. 95, no. 4, p. 045006, 2023. https://doi.org/10.1103/RevModPhys.95.045006.Suche in Google Scholar
[297] F. Ewert and P. van Loock, “$3/4$-efficient bell measurement with passive linear optics and unentangled ancillae,” Phys. Rev. Lett., vol. 113, no. 14, p. 140403, 2014. https://doi.org/10.1103/PhysRevLett.113.140403.Suche in Google Scholar PubMed
[298] E. Kaur, A. Patil, and S. Guha, “Resource-efficient and loss-aware photonic graph state preparation using an array of quantum emitters, and application to all-photonic quantum repeaters,” arXiv.org, 2024. Available at: https://arxiv.org/abs/2402.00731v1 Accessed: Oct. 14, 2024.Suche in Google Scholar
[299] A. Reiserer and G. Rempe, “Cavity-based quantum networks with single atoms and optical photons,” Rev. Mod. Phys., vol. 87, no. 4, pp. 1379–1418, 2015. https://doi.org/10.1103/RevModPhys.87.1379.Suche in Google Scholar
[300] S. Sun, et al.., “A quantum phase switch between a single solid-state spin and a photon,” Nat. Nanotechnol., vol. 11, no. 6, pp. 539–544, 2016. https://doi.org/10.1038/nnano.2015.334.Suche in Google Scholar PubMed
[301] M. K. Bhaskar, et al.., “Experimental demonstration of memory-enhanced quantum communication,” Nature, vol. 580, no. 7801, pp. 60–64, 2020. https://doi.org/10.1038/s41586-020-2103-5.Suche in Google Scholar PubMed
[302] A. Reiserer, “Colloquium: cavity-enhanced quantum network nodes,” Rev. Mod. Phys., vol. 94, no. 4, p. 041003, 2022. https://doi.org/10.1103/RevModPhys.94.041003.Suche in Google Scholar
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Editorial
- Quantum light: creation, integration, and applications
- Reviews
- Low-dimensional solid-state single-photon emitters
- Solid-state single-photon sources operating in the telecom wavelength range
- Quantum super-resolution imaging: a review and perspective
- Perspectives
- New opportunities for creating quantum states of light and matter with intense laser fields
- On-chip frequency-bin quantum photonics
- Building photonic links for microwave quantum processors
- Remote quantum networks based on quantum memories
- Sensing with quantum light: a perspective
- Letter
- Electro-optic frequency shift of single photons from a quantum dot
- Research Articles
- Quantum efficiency of the B-center in hexagonal boron nitride
- Localized exciton emission from monolayer WS2 nanoribbon at cryogenic temperature
- Single-photon emitters in PECVD-grown silicon nitride films: from material growth to photophysical properties
- A fiber-pigtailed quantum dot device generating indistinguishable photons at GHz clock-rates
- Sub-MHz homogeneous linewidth in epitaxial Y2O3: Eu3+ thin film on silicon
- Multimodal Purcell enhancement and optical coherence of Eu3+ ions in a single nanoparticle coupled to a microcavity
- All-optical control of charge-trapping defects in rare-earth doped oxides
- Ultra-broadband and passive stabilization of ultrafast light sources by quantum light injection
- Tunable quantum light by modulated free electrons
- Second-harmonic radiation by on-chip integrable mirror-symmetric nanodimers with sub-nanometric plasmonic gap
- Mie metasurfaces for enhancing photon outcoupling from single embedded quantum emitters
- Design and fabrication of robust hybrid photonic crystal cavities
- Enhanced zero-phonon line emission from an ensemble of W centers in circular and bowtie Bragg grating cavities
- Freeform thin-film lithium niobate mode converter for photon-pair generation
- Luminescence thermometry based on photon emitters in nanophotonic silicon waveguides
- Collective single-photon emission and energy transfer in thin-layer dielectric and plasmonic systems
- Description of ultrastrong light–matter interaction through coupled harmonic oscillator models and their connection with cavity-QED Hamiltonians
- Bound polariton states in the Dicke–Ising model
- Collective multimode strong coupling in plasmonic nanocavities
- Improving quantum metrology protocols with programmable photonic circuits
- Fluorescence enabled phonon counting in an erbium-doped piezo-optomechanical microcavity
- Non-perturbative cathodoluminescence microscopy of beam-sensitive materials
Artikel in diesem Heft
- Frontmatter
- Editorial
- Quantum light: creation, integration, and applications
- Reviews
- Low-dimensional solid-state single-photon emitters
- Solid-state single-photon sources operating in the telecom wavelength range
- Quantum super-resolution imaging: a review and perspective
- Perspectives
- New opportunities for creating quantum states of light and matter with intense laser fields
- On-chip frequency-bin quantum photonics
- Building photonic links for microwave quantum processors
- Remote quantum networks based on quantum memories
- Sensing with quantum light: a perspective
- Letter
- Electro-optic frequency shift of single photons from a quantum dot
- Research Articles
- Quantum efficiency of the B-center in hexagonal boron nitride
- Localized exciton emission from monolayer WS2 nanoribbon at cryogenic temperature
- Single-photon emitters in PECVD-grown silicon nitride films: from material growth to photophysical properties
- A fiber-pigtailed quantum dot device generating indistinguishable photons at GHz clock-rates
- Sub-MHz homogeneous linewidth in epitaxial Y2O3: Eu3+ thin film on silicon
- Multimodal Purcell enhancement and optical coherence of Eu3+ ions in a single nanoparticle coupled to a microcavity
- All-optical control of charge-trapping defects in rare-earth doped oxides
- Ultra-broadband and passive stabilization of ultrafast light sources by quantum light injection
- Tunable quantum light by modulated free electrons
- Second-harmonic radiation by on-chip integrable mirror-symmetric nanodimers with sub-nanometric plasmonic gap
- Mie metasurfaces for enhancing photon outcoupling from single embedded quantum emitters
- Design and fabrication of robust hybrid photonic crystal cavities
- Enhanced zero-phonon line emission from an ensemble of W centers in circular and bowtie Bragg grating cavities
- Freeform thin-film lithium niobate mode converter for photon-pair generation
- Luminescence thermometry based on photon emitters in nanophotonic silicon waveguides
- Collective single-photon emission and energy transfer in thin-layer dielectric and plasmonic systems
- Description of ultrastrong light–matter interaction through coupled harmonic oscillator models and their connection with cavity-QED Hamiltonians
- Bound polariton states in the Dicke–Ising model
- Collective multimode strong coupling in plasmonic nanocavities
- Improving quantum metrology protocols with programmable photonic circuits
- Fluorescence enabled phonon counting in an erbium-doped piezo-optomechanical microcavity
- Non-perturbative cathodoluminescence microscopy of beam-sensitive materials