Electroluminescent vertical tunneling junctions based on WSe2 monolayer quantum emitter arrays: Exploring tunability with electric and magnetic fields

Contributed by Kostya S. Novoselov; received January 26, 2024; accepted April 29, 2024; reviewed by Pawel Hawrylak and Luis Viña
May 31, 2024
121 (23) e2401757121

Significance

Creating and controlling arrays of quantum emitters constitutes a pivotal challenge for the development of devices operating at a quantum level. Here, we realized the formation of arrays of quantum emitters via nanopillar imprinting and helium ion beam irradiation of WSe2 monolayers. The modified monolayer films were incorporated into vertical tunneling light-emitting diodes. Such device geometry enabled electrical excitation of the array and isolation of individual quantum emitters via confocal microscopy. Electrically driven devices provided versatile control knobs for the emission energy via electric fields and the polarization properties linked to the valley degree of freedom via magnetic fields.

Abstract

We experimentally demonstrate the creation of defects in monolayer WSe2 via nanopillar imprinting and helium ion irradiation. Based on the first method, we realize atomically thin vertical tunneling light-emitting diodes based on WSe2 monolayers hosting quantum emitters at deterministically specified locations. We characterize these emitters by investigating the evolution of their emission spectra in external electric and magnetic fields, as well as by inducing electroluminescence at low temperatures. We identify qualitatively different types of quantum emitters and classify them according to the dominant electron–hole recombination paths, determined by the mechanisms of intervalley mixing occurring in fundamental conduction and/or valence subbands.
Creating and controlling light emitting centers in solid-state systems has been actively investigated in materials of reduced dimensionality, such as semiconductor quantum dots (1, 2), monolayer transition metal dichalcogenides (TMDs) (36), or hexagonal boron nitride (7). Recent explorations of van der Waals systems are motivated by their electronic and optoelectronic properties, which are characterized by enhanced light–matter interaction at the nanoscale, rich excitonic optical response, and spin-valley physics while allowing for scalability and device integration. In the monolayer limit, TMDs typically exhibit a direct bandgap at the K point of the Brillouin zone and the optical selection rules are naturally extended to incorporate a valley degree of freedom. The enriched parameter space leads to the formation of diverse types of excitations, particularly when the mechanisms of intervalley mixing are considered. It is evident that the mixing of K+ and K valleys plays a pivotal role in the optical processes involving free and confined excitonic complexes in atomically thin TMDs, e.g., by enabling optical pumping with linearly polarized light (8). A priori, the strength of the intervalley mixing is expected to be enhanced by localization effects due to modifications of the transition energies and/or symmetry breaking by local imperfections of the lattice structure.
The creation of single photon emitters (SPEs), likely originating from defect centers, has been demonstrated experimentally in several TMDs (9), in TMD heterobilayers (10), and in hexagonal boron nitride (hBN) (1115). In most cases, SPEs exhibit a stochastic spatial distribution in the two-dimensional films. In certain scenarios, SPEs have been shown to reside along edges (16), underlying substrate wrinkles (17), and blisters (18); however, those exhibit limited controllability of defect positioning. The two main challenges for the realization of scalable photonic devices based on SPEs in TMD materials are the spatially deterministic creation and electrical control of spectrally bright and stable quantum emitters in these systems. To this end, the creation of ordered arrays of emitters has been achieved via deposition of the TMD monolayers or heterobilayers onto prepatterned nanopillars (9, 19). While such intentionally created emitters exhibit a reasonable yield of quantum emission, such heterostructures do not easily integrate into practical devices.
In this work, we pursue two avenues toward the deterministic creation of SPEs in TMD monolayers and subsequent electrical control. First, we introduce defects into monolayers of WSe2 on hBN via site-selective helium ion irradiation to create light-emitting centers, which can be controlled to the level of single atoms as a function of the incident ions’ type, energy, incidence angle, and dose (ions/area). Second, we deterministically induce SPEs in monolayers of WSe2 by imprinting them with a nanopillar array. After a successful demonstration of quantum emission from the nano-imprinted samples, we proceeded to integrate them into vertical tunneling light-emitting diode (LED) devices. We demonstrate qualitatively different quantum systems, whose distinct optical signatures and magneto-optical properties can be understood by considering various exciton recombination paths enabled by intervalley mixing. Our findings identify the key parameters that determine the specific characteristics of different types of optical spectra of quantum emitters in monolayer TMD in the context of valley physics. The detailed understanding of the recombination processes allows the demonstration of the controllability of quantum states by external fields within technologically relevant electrically driven devices.

Creation of Device-Integrated Arrays of SPEs

First, we focused on the intentional creation of arrays of quantum emitters in monolayer WSe2. Our method of choice was nanopillar imprinting, which created SPEs previously attributed to excitonic transitions within strain-induced localization potentials. However, the length scales of the strain profiles induced by the nanopillars appear to be too large to cause quantization of excitonic states. Therefore, we pursued an alternative hypothesis according to which nano-imprinted SPEs originate through the formation of defect centers. To this end, we compared different approaches to create quantum emitters in monolayer (ML) WSe2 samples: 1) ML-WSe2 deposited onto bulk hBN, which were irradiated with He ions (2027) (Fig. 1A), and ML-WSe2 deposited onto nanopillars (Fig. 1E). The WSe2/hBN structures were irradiated at low He doses—1015to1016ionspercm2. Such small-dose irradiation sites are not detectable in the atomic force microscopy (AFM) topography for depth profile extraction. However, the modifications of the optoelectronic properties are apparent when inspecting the optical response of such irradiated MLs. In the photoluminescence map in Fig. 1B, we saw a significantly quenched intensity of the neutral exciton (X0) at the emission energy 1.73 eV. The spectra were dominated by narrow emission resonances (labeled as XB) over the irradiated flake area, as seen in the map that represents the integrated emission intensity around 1.64 eV in Fig. 1C, which corresponds to the sharp peaks in the PL spectra in Fig. 1D. Such spectra dominated by lower energy narrow resonances obtained from the entire irradiated area, in combination with the decreased X0 emission, suggest that the defects create radiative recombination channels for bound/localized excitons XB (and/or intradefect transitions) and nonradiative recombination channels for free excitons X0. The atomic mass difference between the metal and halogen atoms in TMD monolayers provides selectivity in the helium ion irradiation process, leading to a significantly larger probability of creating halogen vacancies (28). However, the positioning of the emitters is still challenging due to the lateral scattering of light helium ions at the interface between the monolayer film and the substrate. Achieving nanoscale lattice modification requires stringent calibration of the ion dosage, which is dependent on many factors such as the crystal structure and masses of the constituent atoms of the implanted material, the underlying substrate material taking into account its roughness resulting in interface imperfections. The systematic data regarding the calibration of the ion dosage for two different materials (WSe2 and MoS2) and two different substrates (Si/SiO2 and Si/SiO2/hBN) are presented in SI Appendix. Although, ion irradiation has the potential to become a leading method for creating specific SPEs in van der Waals systems with nanoscale positioning, the controllability and reproducibility remain challenging and costly.
Fig. 1.
(A) Optical micrograph of an ML-WSe2/hBN heterostructure irradiated with a low He ion beam dose of 1015 ions.cm‒2. (B) PL map of neutral exciton emission at 1.73 eV. (C) PL map for a selected emission energy of 1.64 eV corresponding to a sharp emission peak in the spectrum in (D). X0 and XB emission peaks are highlighted. (E) Optical micrograph of a WSe2 monolayer deposited onto a nanopillar array. (F) PL map of neutral exciton emission at 1.73 eV and (G) PL map at 1.64 eV showing enhanced emission at the sites previously in touch with the nanopillars. Both maps were obtained after the WSe2 film was taken off the nanopillars and before it was incorporated into the LED heterostructure. (H) Representative spectrum, extracted from the map in (G) measured at the location of the nanopillar featuring X0 and XB emission peaks. [Scale bars: 5 μm in (AC), 20 μm in (EG).]
In the second approach, ML-WSe2 was deposited on top of a rectangular array of nanopillars etched in SiO2 layer. The photoluminescence spectra of WSe2 imprinted by the nanopillars revealed a modification of the optical response of the WSe2 monolayer in the form of an emission intensity enhancement (Fig. 1F) from free excitons X0, which we attributed to a strain-induced local reduction of the single particle band gap, leading to excitons being trapped at the potential wells at the pillar sites through funneling of electrons and holes along strain profile (2934). However, the linewidth and energy of the X0 resonance remain close to the parameters observed in pristine ML-WSe2 film. The free exciton resonance X0 was accompanied by narrow low energy resonances XB, akin to the He ion irradiated samples. The spatial distribution of the XB resonances was ordered and correlated with the nano-pillar pattern. Overall, these spectral characteristics, inspected comparatively between WSe2 monolayers irradiated by focused He ion beams and imprinted by nanopillars arrays, indicate that in both cases the narrow emission lines XB originate from defect-related optical excitations likely involving selenium vacancies.
This hypothesis of defect creation through nanopillar imprinting was further corroborated by the preservation of narrow resonances XB, when the WSe2 monolayer was removed from nanopillar arrays and built into the light-emitting diode device, where AFM imaging shows no visible straining or deformation of the ML. The multistep fabrication process of creating vertical tunneling junction based on Gr/hBN/WSe2/hBN/Gr architecture (Gr—graphene electrodes, hBN—hexagonal boron nitride tunneling barrier, WSe2—imprinted optically active semiconducting monolayer) is pictorially illustrated in Fig. 2AE and described in Materials and Methods. Such configuration enables application of electric field and/or injection of charge carriers in the imprinted WSe2 monolayer by the application of bias between the two graphene electrodes (3541). The fabrication process preserved the narrow emission resonances XB in WSe2 monolayer as demonstrated in Fig. 2F. Those radiative centers demonstrated single photon emission character (36) verified via observation of an antibunching in the second-order photon correlation function shown in Fig. 2G.
Fig. 2.
Schematic representation of the LED fabrication process based on a WSe2 monolayer, which hosts intentionally created quantum emitters (AE). At cryogenic temperature (4.2 K), the photoluminescence spectra display emission resonances in the form of spectrally narrow lines (F), which show an antibunching in the second-order correlation function as illustrated in (G) for the spectrally isolated resonance highlighted by the orange stripe in (F). (H) An optical image of a fabricated LED device with outlined monolayer WSe2, Top and Bottom graphene contacts, electrically contacted with Au electrodes. The thin hBN barrier outlines have been omitted for clarity. The whole device sits on top of a thicker hBN substrate. (Scale bar: 20 μm.)

Optical Properties of Quantum Emitters in Electric and Magnetic Field

We inspected the influence of the bias applied between the graphene electrodes on the optical response of our quantum emitters in LED devices (see Fig. 2H for an optical image of a representative device). When applying the bias in the regime that corresponds to subbandgap electrical excitation, we observe predominantly the effects of the electric field on the emission resonances as shown in Fig. 3A. Generally, we identify two types of emitters: those that display emission energy 1) independent and 2) linearly dependent on the applied bias. For the latter type of emitters, we may assign them a dipole moment that corresponds to the spatial separation of the electron and hole in the range of 30 to 50 pm for various emitters under the assumptions discussed in SI Appendix. Tentatively, we attribute the emergence of the two types of emitters to the deformation of the WSe2 film that is inevitably associated with the process of the creation of the quantum emitters in our method. We may envisage the appearance of wrinkles, folds, and/or other types of distortion of the WSe2 monolayer that could induce spatial separation of charge carriers that form localized excitonic complexes.
Fig. 3.
The low temperature (4.2 K) optoelectrical characterization of the LED devices reveals two classes of quantum emitters based on their response to an electric field, as observed when electrical bias is applied between the two graphene electrodes. (A) The photoluminescence spectra were measured under 514 nm solid-state laser excitation. Two groups of emitters may be identified: those whose energy is independent or linearly dependent on the bias. (B) The latter type of resonances dominates the electroluminescence spectra that are observed upon application of larger bias voltages.
Upon application of the larger bias, we find that our quantum emitters may be electrically excited to raise electroluminescence (EL), as presented in Fig. 3B. We observe a sharp onset at 1.8 V, above which electron and hole tunneling through hBN energy barriers takes place, followed by their radiative recombination in the WSe2 monolayer. That leads to the emergence of narrow emission resonances in the EL spectra, characterized by the linearly dispersive states. This observation suggests that the nondispersive emitters do not give rise to tunneling pathways that favor efficient radiative recombination. Such conditions may be realized, e.g., when the tunneling times are significantly faster than the radiative lifetimes.
Interestingly, a certain duality of our emitters is also revealed by inspecting their magneto-optical response. In a simple view, we identify one class of emission resonances that display exclusively a Zeeman effect and another one when an anticrossing behavior around zero magnetic field is additionally observed. Representative examples of these two types of emitters are presented in Fig. 4 A and B. We interpret this observation by considering plausible recombination processes that may arise due to the unique character of the band structure of monolayer TMDs.
Fig. 4.
Low temperature (4.2 K) magneto-photoluminescence spectra in Faraday configuration with the magnetic field applied perpendicularly to the surface of the WSe2 monolayer are presented. The optical signal is detected in sigma+ polarization. Two types of quantum emitters display (A) exclusively the Zeeman effect and (B) an anticrossing behavior around zero field. We confront the magneto-optical data with (C and D) the calculations of the optical transitions given by the model of an excitonic Hamiltonian involving intervalley mixing as described in the main text. The following parameters were used to obtain the calculated spectra: E0=1.6507 eV, g=10.0, and δ=0 for (C) and E0=1.6371 eV, g=10.4, and δ=1meV for (D). Notably, the weak intensity of the higher energy branch in the latter case for the experimental spectra presented in (B) originates likely from exciton relaxation processes that transfer the excitonic population from the higher energy state to the lower energy one. The eigenstates of the excitonic Hamiltonians used to derive the energy and oscillator strength of the optical transitions may be represented on (E) a valley Bloch sphere, indicating how the excitonic states characterizing the type of quantum emitter demonstrated in (B) are tuned by the magnetic field.
The spin and orbital composition of the fundamental subbands in monolayer WSe2, which define the recombination path of free neutral excitons, is understood well (42). The optical transitions are active in σ+/σ polarization and the helicity of light is coupled to K+/K valleys, respectively. In the presence of a magnetic field, the energy of excitonic resonances is modified by the Zeeman effect with orbital, spin, and valley contributions (43). The linear dependence of the emission energy on the magnetic field for the emitters characterized by narrow resonances is demonstrated in Fig. 4A. Robust spin–orbit effects (44) lead to a spin splitting ΔC=30 meV in the conduction band and ΔV=510 meV in the valence band. The spin-preserving optical transitions related to the ground valence state (commonly referred to as excitons A), involve the electron from the higher energy conduction subband, making the ground state exciton in monolayer WSe2 in principle dark. However, localization effects leading to the formation of quantum emitters may significantly alter this picture. In our samples, we typically observe narrow emission resonances about 60 meV below the free neutral exciton resonance. The difference in the emission energy between free and bound excitons (ΔEPL) may arise from the activation of alternative recombination paths, local modulation of the single particle band-gap and/or Coulomb effects as well as the creation of charged states. Under the assumption that the binding of the excitons occurs predominantly due to selenium vacancies, we can consider two types of localized states: 1) a defect-to-band transition related to the recombination of the electron occupying a Se defect level residing below the conduction band (45) with a free hole and 2) an exciton trapped electrostatically by electrons occupying dangling bond states forming a trion-like excitation (46). These two types of bound states constitute a probable origin of the narrow line emitters in our LED devices. Here, we will focus on developing a minimal model that explains the two types of magneto-optical behavior of our quantum emitters in a semiquantitative fashion.
First, let us note that the key parameters that impact the optical process in our quantum emitters fulfill a condition ΔC<ΔEPL<ΔV. Within such an energy landscape, valley-mixing mechanisms between same-spin states in the conduction band constitute a first-order perturbation to the excitonic state and may be expected to naturally influence the quantum state of our single photon sources. For that reason, the intervalley recombination process of the ground state exciton is characterized by a finite oscillator strength via admixture of a bright exciton defined by a common hole state, as illustrated in Fig. 4B. In the case of the optical transitions within the unperturbed band structure as well as the transitions enabled by conduction band intervalley spin mixing, the quantum emitters display a Zeeman effect with an excitonic g-factor acting as a parameter that distinguishes the two recombination paths. The value of the g-factor in the unperturbed system is equal to −4 based on magneto-luminescence spectra of free neutral excitons in WSe2 monolayers (47). Simple estimations of the g-factors of individual conduction states based on magneto-absorption data for charged excitons (43, 48) demonstrate that the ground state intervalley exciton is characterized by the largest g-factor among all the possible transitions within fundamental subband of WSe2 monolayer with a tentative value of −10. In our samples, we find that statistically our emitters that display exclusively the Zeeman effect are characterized by the g-factors in the range from −4 to −15. Such observation supports the interpretation that intervalley mixing mechanism occurring in the conduction band of WSe2 monolayers plays a significant role in the magneto-optical evolution of quantum emitters. The strength of this mixing is directly reflected in the value of the g-factors of individual resonances so that emitters that display a g-factor of −4 follow a recombination path given by the unperturbed band structure and those emitters with larger g-factor values arise due to states characterized by adequately greater intervalley admixtures. The first-order conduction band perturbations are not sufficient to explain the anticrossing behavior between excitons coupled to K+ and K valleys. To that end, we need to allow the intervalley mixing to occur in the valence band as well, which is unlikely to take place in a pristine WSe2 monolayer lattice. Nevertheless, lattice imperfections that break the inversion symmetry of the WSe2 monolayer constitute a plausible source of higher-order perturbations to localized excitonic states. Selenium vacancies (49, 50) fulfill such requirement, particularly in the presence of an out-of-plane electric field leading to a difference in electrostatic potential between the two selenium planes (51). In such a case, the typical selection rules imposed by the band structure are abolished, in favor of the creation of excitons formed from mixed K+ and K states, as illustrated in Fig. 5C.
Fig. 5.
The schematic representation of the dominant recombination paths of excitons within the fundamental subbands of monolayer WSe2 in the presence of valley mixing mechanisms. The three presented cases correspond to (A) unperturbed band structure, (B) conduction band spin preserving valley mixing and (C) conduction and valence valley mixing induced by a strong perturbation that breaks the inversion symmetry of the WSe2 monolayer crystal structure. The excitons recombining via processes enabled by valley mixing, presented in (B and C) constitute a ground state for the electron–hole pair, unlike the optically active exciton in the unperturbed band structure, where it can be considered as an excited state. Therefore, in low-temperature conditions, the emission resonances due to quantum emitters in the presence of appropriate perturbations are more likely to arise due to the processes depicted in (B and C) than due to the bright exciton recombination presented in (A) because of the enhanced population of the corresponding excitonic states.
We may write a valley exciton Hamiltonian in the presence of a magnetic field that captures the magneto-optical behavior of our quantum emitters. In the basis of an electron–hole pair residing in σ+ and σ valley, corresponding to vectors (1,0) and (0,1), respectively, the excitonic Hamiltonian HX takes the form:
HX=(E012gμBBδ2δ2E0+12gμBB),
[1]
where E0 is the transition energy, g is the excitonic g-factor, μB is the Bohr magneton and δ is the intervalley mixing parameter. The exciton energy, as seen in the magneto-photoluminescence spectra, is given by the eigenvalues E0±12gμBB2+δ2. The eigenvectors are indicative of the mixing strength between σ+ and σ valley and define the oscillator strength of the transitions detected in σ+ and σ polarizations:
φ1=(1,((gμB B)2+δ2)gμB Bδ){(1,1),B=0(1,0),gμBBδφ2=(((gμBB)2+δ2)+gμBBδ,1){(1,1),B=0(0,1),gμBBδ.
[2]
The energy and oscillator strength of the optical transitions active in σ+ polarization with the parameters adjusted to reproduce the experimental spectra are presented in Fig. 4 C and D as color maps. The introduced parameters allow us to distinguish between the three recombination processes presented in Fig. 5AC via the following conditions:
{δ=0,g4forunperturbedbandstructureδ=0,g4forfirst-ordervalleymixingδ>0forhigher-ordervalleymixing.
[3]
These three scenarios can be understood as different levels of preservation of the orbital Bloch functions of the free exciton in the localized state. Electrostatically bound excitons may exhibit wave function close to one originating from the unperturbed band structure, while transitions involving defect states and band carriers lead to valley mixing effects.
Our magneto-photoluminescence spectra detected in σ+ polarization allows a straightforward identification of the type of quantum emitter by recognizing the dominant recombination paths as given by the proposed parameterization. The particulars of the excitonic states displayed by quantum emitters characterized by the excitonic Hamiltonian can be illustrated on the valley Bloch sphere as schematically presented in Fig. 4E. For δ=0, the quantum emitters exist naturally in pure K+ and K states, which reside on the north and south poles of the Bloch sphere. For δ>0, the excitonic states are magnetic field dependent. Without a magnetic field, the excitonic states take the form: 12(|K+±|K), so that they reside on the opposite sides on the equator of the Bloch sphere. Upon application of a magnetic field, the eigenstates are gradually shifted along the meridian, eventually reaching the north and south poles in the limit of gμBBδ.
In conclusion, we have fabricated electrically driven light-emitting diodes that host intentionally created quantum emitters in a monolayer WSe2. By inspecting the evolution of the photoluminescence and electroluminescence spectra in electric and magnetic fields, we identified various types of quantum emitters. We have proposed a minimal model, that allows the classification of our emitters by their dominant recombination processes within the fundamental subbands of monolayer WSe2, which are determined by valley mixing mechanisms. Our findings identify quantum emitters as valley qubits, whose energy and composition of wave function are tunable, to a certain degree, by external fields.

Materials and Methods

Sample Preparation.

WSe2 flakes for Helium ion irradiation were prepared via mechanical exfoliation in air of commercially available bulk crystals (2D Semiconductors) onto Si/SiO2 (300 nm) substrates. The hBN/WSe2 heterostructures were assembled via the dry pickup method with PDMS/Polycarbonate (5% in Dichloromethane). SiO2 and hBN-supported ML-WSe2 were irradiated with a 30 kV He+ beam in a Zeiss ORION Nanofab Helium ion microscope. The beam current was I=1.1 pA, dwell time was t=10 μs and the dose was varied in the range 1015 to 1018 ions.cm‒2.
All crystals for the LED devices were exfoliated in the air onto PMMA membranes (hBN and Graphene onto 8% PMMA in anisole and WSe2 onto 3% PMMA in anisole) and after suitable flakes were identified, the substrates were placed in a nitrogen desiccator. The WSe2 monolayers for the LED devices were prepared onto either SiO2 substrates or onto spin-coated membranes of PMMA on polymethylglutarimide. A dry peel membrane technique was used for all transfers, using established pick-up and peel-off techniques. First, the WSe2 monolayer, exfoliated on a PMMA membrane, is brought into contact with a nanostructure (nanopillars or simple step edges), and the flake is allowed to conform to the structure for 30 min while the substrate was heated at 65 °C. Before proceeding to the next fabrication step, a check of the optical properties of the WSe2 flake revealed enhancement of the photoluminescence intensity at room temperature under laser excitation above the band gap, at 514.5 nm (2.4098 eV). Next, the WSe2 flake is lifted from the nanopillars and incorporated into a tunneling heterostructure composed of top and bottom graphene layers, separated from the flake via thin, few-layer hBN barriers. The graphene electrodes are electrically contacted with Cr/Au leads via electron beam lithography and subsequent electron beam evaporation.

Optical Spectroscopy.

The photoluminescence spectra were measured under the conditions of above band gap excitation with a 515 nm diode laser. The optical response of the sample was inspected in standard backscattering geometry. The laser was focused on the sample with a high numerical aperture and low-temperature objective with titanium housing that allows stable measurements in high magnetic fields. Piezo-positioners in x-y-z configuration were used to position the sample with respect to the focused laser beam. A set of filters, linear polarizers, and λ/4 waveplates was used to detect photoluminescence spectra in circular polarization. The sample was cooled down to 4.2 K via Helium exchange gas or via thermal contact with a cold finger. A 500 cm spectrometer with a charge-coupled device camera was used to measure spectrally resolved optical signals. The magnetic field generated by superconducting coils was applied to the sample in the Faraday configuration. The bias to the graphene electrodes was applied by Keithley 2614B source-meter. The photon correlation experiments were done in Hanbury Brown and Twiss set-up with avalanche photodiodes acting as single photon detectors. The statistics of the time delays between pulses generated by the avalanche photodiodes were measured by a specialized counting module. The correlation functions were described by a formula characteristic of a two-level system:
g2(τ)=1Ae|τ|τ0,
[4]
where τ is the time delay between the photon detection events, A is the amplitude of the antibunching, and t0 is the characteristic time constant.

Data, Materials, and Software Availability

The data that support the findings of this study (52) are openly available at the following DOI: https://doi.org/10.58132/BPYO1C. All other data are included in the manuscript and/or SI Appendix.

Acknowledgments

This project was supported by the Ministry of Education (Singapore) through the Research Centre of Excellence program (Grant EDUN C-33-18-279-V12, I-FIM), and AcRF Tier 3 (MOE2018T3-1-005). This research is supported by the Ministry of Education, Singapore, under its Academic Research Fund Tier 2 (T2EP50122-0012). This material was based upon work supported by the Air Force European Office of Aerospace Research and Development Office of Scientific Research and the Office of Naval Research Global under Award No. FA8655-21-1-7026. The work was supported by the National Science Centre, Poland (Grant No. 2022/46/E/ST3/00166). K.S.N. acknowledges support from the Royal Society (UK, Grant No. RSRP/R/190000).

Author contributions

K.S.N. and M.K. designed research; J.H., K.V., M.G., G.B., L.H., and A.K. performed research; all authors analyzed data; and J.H., K.V., A.K., and M.K. wrote the paper.

Competing interests

The authors declare no competing interest.

Supporting Information

Appendix 01 (PDF)

References

1
M. Reimer et al., Towards scalable gated quantum dots for quantum information applications. Physica E 40, 1790–1793 (2008).
2
J. Kobak et al., Designing quantum dots for solotronics. Nat. Commun. 5, 3191 (2014).
3
C. Chakraborty, L. Kinnischtzke, K. M. Goodfellow, R. Beams, A. N. Vamivakas, Voltage-controlled quantum light from an atomically thin semiconductor. Nat. Nanotechnol. 10, 507–511 (2015).
4
Y. M. He et al., Single quantum emitters in monolayer semiconductors. Nat. Nanotechnol. 10, 497–502 (2015).
5
M. Koperski et al., Single photon emitters in exfoliated WSe2 structures. Nat. Nanotechnol. 10, 503–506 (2015).
6
A. Srivastava et al., Optically active quantum dots in monolayer WSe2. Nat. Nanotechnol. 10, 491–496 (2015).
7
M. Koperski et al., Midgap radiative centers in carbon-enriched hexagonal boron nitride. Proc. Natl. Acad. Sci. U.S.A. 117, 13214–13219 (2020).
8
A. M. Jones et al., Optical generation of excitonic valley coherence in monolayer WSe2. Nat. Nanotechnol. 8, 634–638 (2013).
9
C. Palacios-Berraquero et al., Large-scale quantum-emitter arrays in atomically thin semiconductors. Nat. Commun. 8, 15093 (2017).
10
A. Ciarrocchi, F. Tagarelli, A. Avsar, A. Kis, Excitonic devices with van der Waals heterostructures: Valleytronics meets twistronics. Nat. Rev. Mater. 7, 449–464 (2022).
11
T. T. Tran, K. Bray, M. J. Ford, M. Toth, I. Aharonovich, Quantum emission from hexagonal boron nitride monolayers. Nat. Nanotechnol. 11, 37–41 (2016).
12
M. Koperski, K. Nogajewski, M. Potemski, Single photon emitters in boron nitride: More than a supplementary material. Opt. Commun. 411, 158–165 (2018).
13
A. B. D. Shaik, P. Palla, Optical quantum technologies with hexagonal boron nitride single photon sources. Sci. Rep. 11, 12285 (2021).
14
M. Koperski et al., Towards practical applications of quantum emitters in boron nitride. Sci. Rep. 11, 15506 (2021).
15
D. I. Badrtdinov et al., Dielectric environment sensitivity of carbon centers in hexagonal boron nitride. Small 19, 2300144 (2023).
16
P. Tonndorf et al., Single-photon emission from localized excitons in an atomically thin semiconductor. Optica 2, 347–352 (2015).
17
D. Yim, M. Yu, G. Noh, J. Lee, H. Seo, Polarization and localization of single-photon emitters in hexagonal boron nitride wrinkles. ACS Appl. Mater. Interfaces 12, 36362–36369 (2020).
18
T. P. Darlington et al., Imaging strain-localized excitons in nanoscale bubbles of monolayer WSe2 at room temperature. Nat. Nanotechnol. 15, 854–860 (2020).
19
A. Branny, S. Kumar, R. Proux, B. D. Gerardot, Deterministic strain-induced arrays of quantum emitters in a two-dimensional semiconductor. Nat. Commun. 8, 15053 (2017).
20
Z. Li, F. Chen, Ion beam modification of two-dimensional materials: Characterization, properties, and applications. Appl. Phys. Rev. 4, 011103 (2017).
21
J. Klein et al., Site-selectively generated photon emitters in monolayer MoS2 via local helium ion irradiation. Nat. Commun. 10, 2755 (2019).
22
V. Iberi et al., Nanoforging single layer MoSe2 through defect engineering with focused helium ion beams. Sci. Rep. 6, 30481 (2016).
23
F. I. Allen, A review of defect engineering, ion implantation, and nanofabrication using the helium ion microscope. Beilstein J. Nanotechnol. 12, 633–664 (2021).
24
S. Kretschmer et al., Supported two-dimensional materials under ion irradiation: The substrate governs defect production. ACS Appl. Mater. Interfaces 10, 30827–30836 (2018).
25
W. Yang et al., Interface effects on He ion irradiation in nanostructured materials. Materials 12, 2639 (2019).
26
E. Mitterreiter et al., Atomistic positioning of defects in helium ion treated single-layer MoS2. Nano Lett. 20, 4437–4444 (2020).
27
M. Ghorbani-Asl, S. Kretschmer, A. V. Krasheninnikov, “Chapter 9—Two-dimensional materials under ion irradiation: From defect production to structure and property engineering” in Defects in Two-Dimensional Materials, R. Addou, L. Colombo, Eds. (Elsevier, 2022), pp. 259–301.
28
A. Hötger et al., Spin-defect characteristics of single sulfur vacancies in monolayer MoS2. npj 2D Mater. Appl. 7, 30 (2023).
29
S. Bertolazzi, J. Brivio, A. Kis, Stretching and breaking of ultrathin MoS2. ACS Nano 5, 9703–9709 (2011).
30
H. J. Conley et al., Bandgap engineering of strained monolayer and bilayer MoS2. Nano Lett. 13, 3626–3630 (2013).
31
B. Amin, T. P. Kaloni, U. Schwingenschlögl, Strain engineering of WS2, WSe2, and WTe2. RSC Adv. 4, 34561–34565 (2014).
32
D. Lloyd et al., Band gap engineering with ultralarge biaxial strains in suspended monolayer MoS2. Nano Lett. 16, 5836–5841 (2016).
33
O. B. Aslan, M. Deng, T. F. Heinz, Strain tuning of excitons in monolayer WSe2. Phys. Rev. B 98, 115308 (2018).
34
A. V. Tyurnina et al., Strained bubbles in van der Waals heterostructures as local emitters of photoluminescence with adjustable wavelength. ACS Photonics 6, 516–524 (2019).
35
R. Cheng et al., Electroluminescence and photocurrent generation from atomically sharp WSe2/MoS2 heterojunction p-n diodes. Nano Lett. 14, 5590–5597 (2014).
36
F. Withers et al., WSe2 light-emitting tunneling transistors with enhanced brightness at room temperature. Nano Lett. 15, 8223–8228 (2015).
37
F. Withers et al., Light-emitting diodes by band-structure engineering in van der Waals heterostructures. Nat. Mater. 14, 301–306 (2015).
38
C. Palacios-Berraquero et al., Atomically thin quantum light-emitting diodes. Nat. Commun. 7, 12978 (2016).
39
J. Binder et al., Sub-bandgap voltage electroluminescence and magneto-oscillations in a WSe2 light-emitting van der Waals heterostructure. Nano Lett. 17, 1425–1430 (2017).
40
J. Zultak et al., Ultra-thin van der Waals crystals as semiconductor quantum wells. Nat. Commun. 11, 125 (2020).
41
J. Binder et al., Upconverted electroluminescence via Auger scattering of interlayer excitons in van der Waals heterostructures. Nat. Commun. 10, 2335 (2019).
42
G. Wang et al., Colloquium: Excitons in atomically thin transition metal dichalcogenides. Rev. Mod. Phys. 90, 021001 (2018).
43
M. Koperski et al., Orbital, spin and valley contributions to Zeeman splitting of excitonic resonances in MoSe2, WSe2 and WS2 monolayers. 2D Mater. 6, 015001 (2019).
44
Z. Y. Zhu, Y. C. Cheng, U. Schwingenschlögl, Giant spin-orbit-induced spin splitting in two-dimensional transition-metal dichalcogenide semiconductors. Phys. Rev. B Condens. Matter Mater. Phys. 84, 153402 (2011).
45
Z. Cui et al., Effect of vacancy defect on optoelectronic properties of monolayer tungsten diselenide. Opt. Quant. Electron. 50, 1 (2017).
46
A. Wojs, P. Hawrylak, Negatively charged magnetoexcitons in quantum dots. Phys. Rev. B 51, 10880–10885 (1995).
47
A. Srivastava et al., Valley Zeeman effect in elementary optical excitations of monolayer WSe2. Nat. Phys. 11, 141–147 (2015).
48
M. Koperski et al., Optical properties of atomically thin transition metal dichalcogenides: Observations and puzzles. Nanophotonics 6, 1289–1308 (2017).
49
Y. Guo, D. Liu, J. Robertson, Chalcogen vacancies in monolayer transition metal dichalcogenides and fermi level pinning at contacts. Appl. Phys. Lett. 106, 173106 (2015).
50
Y. J. Zheng et al., Point defects and localized excitons in 2d WSe2. ACS Nano 13, 6050–6059 (2019).
51
A. Altıntaş et al., Spin-valley qubits in gated quantum dots in a single layer of transition metal dichalcogenides. Phys. Rev. B 104, 195412 (2021).
52
J. Howarth et al., Data from “Electroluminescent vertical tunneling junctions based on WSe2 monolayer quantum emitter arrays: Exploring tunability with electric and magnetic fields.” University of Warsaw Research Data Repository. https://doi.org/10.58132/BPYO1C. Deposited 17 May 2024.

Information & Authors

Information

Published in

The cover image for PNAS Vol.121; No.23
Proceedings of the National Academy of Sciences
Vol. 121 | No. 23
June 4, 2024
PubMed: 38820004

Classifications

Data, Materials, and Software Availability

The data that support the findings of this study (52) are openly available at the following DOI: https://doi.org/10.58132/BPYO1C. All other data are included in the manuscript and/or SI Appendix.

Submission history

Received: January 26, 2024
Accepted: April 29, 2024
Published online: May 31, 2024
Published in issue: June 4, 2024

Keywords

  1. quantum emitters
  2. monolayer transition metal dichalcogenides
  3. electroluminescence
  4. tunneling junction
  5. light-emitting diodes

Acknowledgments

This project was supported by the Ministry of Education (Singapore) through the Research Centre of Excellence program (Grant EDUN C-33-18-279-V12, I-FIM), and AcRF Tier 3 (MOE2018T3-1-005). This research is supported by the Ministry of Education, Singapore, under its Academic Research Fund Tier 2 (T2EP50122-0012). This material was based upon work supported by the Air Force European Office of Aerospace Research and Development Office of Scientific Research and the Office of Naval Research Global under Award No. FA8655-21-1-7026. The work was supported by the National Science Centre, Poland (Grant No. 2022/46/E/ST3/00166). K.S.N. acknowledges support from the Royal Society (UK, Grant No. RSRP/R/190000).
Author contributions
K.S.N. and M.K. designed research; J.H., K.V., M.G., G.B., L.H., and A.K. performed research; all authors analyzed data; and J.H., K.V., A.K., and M.K. wrote the paper.
Competing interests
The authors declare no competing interest.

Notes

Reviewers: P.H., University of Ottawa; and L.V., Universidad Autónoma de Madrid.

Authors

Affiliations

James Howarth1
School of Physics and Astronomy, University of Manchester, Manchester M13 9PL, United Kingdom
Kristina Vaklinova2,1 [email protected]
Institute for Functional Intelligent Materials, National University of Singapore, Singapore 117544, Singapore
Magdalena Grzeszczyk
Institute for Functional Intelligent Materials, National University of Singapore, Singapore 117544, Singapore
Giulio Baldi
Department of Physics, National University of Singapore, Singapore 119077, Singapore
Lee Hague
School of Physics and Astronomy, University of Manchester, Manchester M13 9PL, United Kingdom
Laboratoire National des Champs Magnétiques Intenses, CNRS-Université Grenoble Alpes-Université Paul Sabatier-Institut National des Sciences Appliquées Toulouse, Grenoble 38042, France
Center for Terahertz Research and Applications Labs, Institute of High Pressure Physics, Polish Academy of Sciences, Warsaw 01-142, Poland
Institute for Functional Intelligent Materials, National University of Singapore, Singapore 117544, Singapore
Department of Materials Science and Engineering, National University of Singapore, Singapore 117575, Singapore
Aleksey Kozikov
Faculty of Science, Agriculture & Engineering, School of Mathematics, Statistics and Physics, Newcastle University, Newcastle Upon Tyne NE1 7RU, United Kingdom
Institute for Functional Intelligent Materials, National University of Singapore, Singapore 117544, Singapore
Department of Materials Science and Engineering, National University of Singapore, Singapore 117575, Singapore

Notes

2
To whom correspondence may be addressed. Email: [email protected], [email protected], or [email protected].
1
J.H. and K.V. contributed equally to this work.

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Altmetrics




Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Electroluminescent vertical tunneling junctions based on WSe2 monolayer quantum emitter arrays: Exploring tunability with electric and magnetic fields
    Proceedings of the National Academy of Sciences
    • Vol. 121
    • No. 23

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media