Skip to content
BY 4.0 license Open Access Published by De Gruyter April 21, 2020

Engineered telecom emission and controlled positioning of Er3+ enabled by SiC nanophotonic structures

  • Natasha Tabassum , Vasileios Nikas , Alex E. Kaloyeros , Vidya Kaushik , Edward Crawford , Mengbing Huang and Spyros Gallis ORCID logo EMAIL logo
From the journal Nanophotonics

Abstract

High-precision placement of rare-earth ions in scalable silicon-based nanostructured materials exhibiting high photoluminescence (PL) emission, photostable and polarized emission, and near-radiative-limited excited state lifetimes can serve as critical building blocks toward the practical implementation of devices in the emerging fields of nanophotonics and quantum photonics. Introduced herein are optical nanostructures composed of arrays of ultrathin silicon carbide (SiC) nanowires (NWs) that constitute scalable one-dimensional NW-based photonic crystal (NW-PC) structures. The latter are based on a novel, fab-friendly, nanofabrication process. The NW arrays are grown in a self-aligned manner through chemical vapor deposition. They exhibit a reduction in defect density as determined by low-temperature time-resolved PL measurements. Additionally, the NW-PC structures enable the positioning of erbium (Er3+) ions with an accuracy of 10 nm, an improvement on the current state-of-the-art ion implantation processes, and allow strong coupling of Er3+ ions in NW-PC. The NW-PC structure is pivotal in engineering the Er3+-induced 1540-nm emission, which is the telecommunication wavelength used in optical fibers. An approximately 60-fold increase in the room-temperature Er3+ PL emission is observed in NW-PC compared to its thin-film analog in the linear pumping regime. Furthermore, 22 times increase in the Er3+ PL intensity per number of exited Er ions in NW-PC was observed at saturation while using 20 times lower pumping power. The NW-PC structures demonstrate broadband and efficient excitation characteristics for Er3+, with an absorption cross-section (~2 × 10−18 cm2) two-order larger than typical benchmark values for direct absorption in rare-earth-doped quantum materials. Experimental and simulation results show that the Er3+ PL is photostable at high pumping power and polarized in NW-PC and is modulated with NW-PC lattice periodicity. The observed characteristics from these technologically friendly nanophotonic structures provide a promising route to the development of scalable nanophotonics and formation of single-photon emitters in the telecom optical wavelength band.

1 Introduction

In recent years, there has been a tremendous interest in the synthesis, properties, and applications of semiconductor nanowires (NWs). This interest has been complemented by the industry’s enthusiasm toward faster nanodevices exhibiting high functionality with reduced energy consumption. The synthesis and simultaneous on-demand positioning of NWs, coupled with the ability to control their orientation and spatial assembly, are critical factors toward the production of nanosystems and nanodevices [1]. The primary limiting challenge commonly faced, especially for feature sizes below 100 nm and bottom-up approaches, is the required deterministic and scalable integration, which involves control over the density, orientation, and spacing of the synthesized NWs. It has been a consensus that semiconductor NWs can help cultivate the next generation of electronics and photonics [2], [3], biological and medical technologies [4], [5], as well as energy technologies [6], [7]. Following this trend, significant effort has been focused on the development of nanostructured materials that can be employed ubiquitously in several exciting emerging quantum technologies, spanning quantum sensing, quantum photonics, and quantum communication and storage [8], [9].

Macroscopic rare-earth-doped materials are one of the most promising candidates for quantum storage of single photons and signal processing [10], [11], [12]. The realization of scalable on-chip optical quantum memory and quantum optical signal processing devices requires novel nanostructured materials that must be highly integrable and compatible with existing electronic circuits, waveguide architectures, and current chip-scale and silicon (Si) process technology [2], [13]. The controlled placement of rare-earth ions into Si-based wide bandgap nanostructured materials with high integration functionality, such as silicon carbide (SiC) NWs [9], [14], [15], [16], can serve as critical building blocks toward the implementation of such quantum devices. Silicon carbide is a silicon-based wide-band-gap material with excellent thermal, mechanical, and physical properties. SiC is chemically inert and exhibits almost negligible magnetic moment (a necessary requirement for hosting quantum emitters) [17], making it a promising host material for single-photon emitters and qubits with several applications in quantum technologies [9]. Rare-earth erbium (Er3+) ions have an emission at the technologically important wavelength of ~1540 nm, as it falls in the lowest loss wavelength range (telecom C-band) of fiber optics networks [18]. As such, Er-doped materials are the only materials that can potentially meet the current requirements to realize a quantum memory/repeater at telecom C-band wavelengths [11], [12].

For optical quantum storage and signal processing applications, optically active rare-earth ions in solid-state hosts must exhibit high optical pumping and photoluminescence (PL) emission to enable efficient storage and high-fidelity optical readout of the ion’s quantum state. Furthermore, near-radiative-limited excited state lifetimes of the ions are important to achieve long storage and coherence times [10], [11]. To this end, photonic crystal (PC) nanostructures and PC cavities have attracted attention for their ability to suppress or enhance spontaneous emission rate, respectively, and control light directionality of emitters [19], [20]. By properly engineering such nanophotonic structures, it is possible to suppress (photonic bandgap structures) or enhance (photonic cavities) the local density of photon states compared to the free-space value, thus respectively decreasing or increasing the emission rate. This has been recently demonstrated for rare-earth neodymium- and erbium-based quantum sources and memories [13], [21]. Such nanostructures are typically fabricated via top-down approaches, such as electron beam lithography (as is the case in [21]) or focused ion beam [9], or bottom-up approaches, such as the vapor-liquid-solid [22]. However, these approaches suffer from drawbacks; top-down fabrication of features below 30 nm features becomes difficult and costly, while bottom-up approaches must facilitate the required control over density, geometry, orientation, spacing, and alignment of individual structures [23]. Realization of PC nanostructures requires the development of advanced synthesis methods and novel nanofabrication schemes [9].

In this paper, we report on a new class of complementary metal oxide semiconductor-compatible nanophotonic structures based on scalable SiC NW-based PCs (NW-PC) doped with and without oxygen, and Er ions. The ultrathin NWs were fabricated in a self-aligned manner through a novel catalyst-free chemical vapor deposition (CVD) synthesis route, without the use of a lithographic pattern transfer technique. This nanofabrication scheme reduces defect density and enables precise control over the geometry of the NWs. The surface roughness and crystallinity of the NWs enable the growth of ultrathin ≤20-nm structures and enable the controlled placement of Er3+ ions down to a nanoscale resolution. Another major advantage is the engineering of the Er3+-induced 1540-nm emission through efficient coupling with the NW-PC. We demonstrate that these one-dimensional (1D) PC structures provide an extremely efficient excitation route for Er3+, exhibiting a photostable and polarized Er3+ emission for potential applications in nanophotonics, and long-distance telecom optical quantum networks.

2 Nanofabrication

2.1 Fabrication of self-aligned ultrathin NW arrays

Self-aligned 1D NW arrays grown at predetermined positions were fabricated on silicon substrates via thermal CVD. An abridged fabrication process of the NW-PC is schematically depicted in Figure 1. First, a hydrogen silsesquioxane (HSQ) negative-tone resist layer was spin-coated onto Si. We used electron beam lithography to expose line patterns, and the resulting wafer piece was developed in a chemical solution bath yielding an HSQ ribbon array (Figure 1A-i) [24]. After the development process, we conformally deposited ≤20-nm-thick SiC (with an index of refraction, n=2.7) or SiC doped with oxygen, SiC:O (SiC0.5O0.8, n=2.2), onto the HSQ template using CVD processing (Figure 1A-ii), described elsewhere [25], [26].

Figure 1: SiC nanostructured materials through on-demand placement of NW and Er3+.(A) Fabrication process steps for realizing self-aligned 20-nm-width NW-PC structures and representative cross-section SEM images. i. Spin-coating and electron beam lithography of HSQ (blue) to create a ribbon template on a substrate (gray). ii. Self-aligned deposition of conformal SiC or SiC:O ultrathin layer (green). Geometry control using a conformal oxide (purple) and SiC layer. iii. Anisotropic RIE. iv. Wet-etch removal of HSQ. (B) On-demand integration of Er ions in the NW-PC. i. Deposition of sacrificial oxide layer (orange) and CMP. ii. Combination of RIE and wet-etch to expose the top of NWs. iii. Deposition of thin oxide layer (yellow) based on targeted ion implantation depth. The corresponding top-down SEM images of step ii and iii are shown. iv. Er ion (red circles) implantation and wet-etch to remove oxide layers. (C) Schematic cross-section diagram of the resulting Er-doped 1D NW-PC structures with lattice periodicity P1. P2: sub-lattice periodicity; H: height of the NWs; W: width of the NWs. (D) Low magnification top-down and (E) high magnification cross-section SEM images of the resulting Er-doped NW-PC structures (W=20 nm, P1=400 nm, P2=100 nm, H=130 nm).
Figure 1:

SiC nanostructured materials through on-demand placement of NW and Er3+.

(A) Fabrication process steps for realizing self-aligned 20-nm-width NW-PC structures and representative cross-section SEM images. i. Spin-coating and electron beam lithography of HSQ (blue) to create a ribbon template on a substrate (gray). ii. Self-aligned deposition of conformal SiC or SiC:O ultrathin layer (green). Geometry control using a conformal oxide (purple) and SiC layer. iii. Anisotropic RIE. iv. Wet-etch removal of HSQ. (B) On-demand integration of Er ions in the NW-PC. i. Deposition of sacrificial oxide layer (orange) and CMP. ii. Combination of RIE and wet-etch to expose the top of NWs. iii. Deposition of thin oxide layer (yellow) based on targeted ion implantation depth. The corresponding top-down SEM images of step ii and iii are shown. iv. Er ion (red circles) implantation and wet-etch to remove oxide layers. (C) Schematic cross-section diagram of the resulting Er-doped 1D NW-PC structures with lattice periodicity P1. P2: sub-lattice periodicity; H: height of the NWs; W: width of the NWs. (D) Low magnification top-down and (E) high magnification cross-section SEM images of the resulting Er-doped NW-PC structures (W=20 nm, P1=400 nm, P2=100 nm, H=130 nm).

Post-deposition thermal treatment was carried out in forming gas ambient (95% argon [Ar] and 5% hydrogen [H2]) at 1100°C for 1 h. Following the annealing of the ultrathin chemically grown layer, we performed anisotropic reactive ion etching (RIE) to remove the top SiC or SiC:O layer, leaving the sidewall self-aligned layer (NWs) intact and exposing the HSQ template (Figure 1A-iii). For control over the geometry of the NWs, the SiC sidewalls are encapsulated with a conformal oxide and SiC layer. This helps protect the self-aligned sidewalls during the ensuing RIE and enables control over the shape and height of the resulting NWs. A wet etch, buffered hydrofluoric (BHF) acid, was used to remove the HSQ, yielding ultrathin NW arrays synthesized in a self-aligned manner (Figure 1A-iv).

This synthesis route utilizes new precursor chemistry, TSCH (1,3,5-trisilacyclohexane) from Gelest Inc. (see Methods), resulting in highly crystalline matrix with good surface roughness, enabling the fabrication of NWs with width (W) of 10 nm (examples of more structures are shown in Figure S1). The critical dimension, width, of the NWs solely depends on the thickness of the conformal sidewall layer. The height (H), lattice periodicity (P1), and sub-lattice periodicity (P2) of the NW array are also shown in Figure 1C.

2.2 On-demand placement and integration of Er3+ into NW-PC

Panel b of Figure 1 presents the major process steps employed to engineer the position of Er3+ ions in the NW-PC. First, we deposited a sacrificial thick oxide layer onto the NW array using CVD [24], followed by chemical mechanical planarization (CMP) to flatten and thin the oxide layer. We then performed controlled RIE and wet-etch steps to recess the oxide layer and to expose the top of the NWs. A thin (15 nm) conformal CVD-grown oxide layer was then deposited to encapsulate the NWs. We can modulate the thickness of the oxide layer based on the targeted ion-implantation depth in the NWs. Accordingly, the remaining thicker oxide layer between the NWs prevents ion implantation into the substrate.

Samples were then subjected to a 150 keV beam of erbium ions targeting ions to be embedded into 30-nm depth from the top of the NW arrays, as determined by Transport of Ions in Matter simulations. Various doses of Er ions – 1×1013 cm−2 (low), 5×1013 cm−2 (medium), and 1×1014 cm−2 (high) – were implanted into the NW-PC structures and (thin film) samples by ion implantation. We then carried out a wet-etch to remove the oxide layers yielding the controlled-Er-doped NW-PC structures (Figure 1D–E). This process was followed by a postimplantation Ar annealing (1 h, 900°C) and a forming gas (95% nitrogen [N2] and 5% H2) thermal treatment (1 h, 850°C) to activate the Er ions optically and for surface passivation [27]. The integration scheme dictates the final positions of the implanted ions to be within the 10 nm width of the NWs, which is an improvement to the accuracy of a state-of-the-art deterministic ion implantation process for single atom devices in [28] and the single ion implantation process in [29].

3 Results and discussion

3.1 Modeling of a dipole emission in NW-PC structures

The extraction efficiency of the spontaneous emission from a dipole in 1D NW-PC and a uniform slab without a PC structure was explored using finite-difference time-domain (FDTD) calculations (Figure 2). We used a monitor parallel to the xy-plane to calculate the extraction efficiency of the emitted radiation power from the dipole (see Methods and Figure 2A). The extraction efficiency, next, is defined as the fraction of the radiation power transmitted through the monitor to the total emitted power from the dipole [19], [30].

Figure 2: FDTD modeling of the spontaneous emission extraction efficiency for the NW-PC and reference (Ref.) thin-film slabs.(A) i. Schematic diagram of the geometry used in the FDTD computation with the relative position of the monitor used for extraction efficiency calculation. ii. Top-down schematic of the geometry showing the orientation of y-polarized dipole. (B) The emission extraction efficiencies, ηext, from the NW-PC structures and reference as a function of wavelength. Gray data points refer to SiC:O structure with n=2.2. (C) Calculated spatial distribution of |E|2 at 1540 nm in the zx plane (y-polarized dipole, n=2.7) for i. the reference and ii. the 600-nm-P1 NW-PC. The dashed lines in each map represent the outline of the structure. iii. Dispersion relation of the 600-nm-P1 NW-PC structure (periodicity, a=P1=600 nm), showing an incomplete photonic bandgap extending from 190 THz (1577 nm) to 229.2 THz (1308 nm).
Figure 2:

FDTD modeling of the spontaneous emission extraction efficiency for the NW-PC and reference (Ref.) thin-film slabs.

(A) i. Schematic diagram of the geometry used in the FDTD computation with the relative position of the monitor used for extraction efficiency calculation. ii. Top-down schematic of the geometry showing the orientation of y-polarized dipole. (B) The emission extraction efficiencies, ηext, from the NW-PC structures and reference as a function of wavelength. Gray data points refer to SiC:O structure with n=2.2. (C) Calculated spatial distribution of |E|2 at 1540 nm in the zx plane (y-polarized dipole, n=2.7) for i. the reference and ii. the 600-nm-P1 NW-PC. The dashed lines in each map represent the outline of the structure. iii. Dispersion relation of the 600-nm-P1 NW-PC structure (periodicity, a=P1=600 nm), showing an incomplete photonic bandgap extending from 190 THz (1577 nm) to 229.2 THz (1308 nm).

FDTD calculations revealed that the emission extraction efficiency from the SiC slab was ~3.6% at most wavelengths across the simulated range (Figure 2B). The observed behavior is consistent with findings from other studies indicating a high density of guided optical modes within the high-refractive-index SiC over a broad wavelength range [19], [30]. Similarly, we performed FDTD calculations in SiC NW-PC structures with three different P1 values that revealed two major differences. First, the extraction efficiency was enhanced throughout the whole investigated wavelength range with values up to ~40% (Figure 2B). Specifically, in the 600-nm-P1 NW-PC sample, the Er3+-induced 1540-nm emission extraction efficiency was 39%, an order of magnitude higher than that in the slab with ~3.6%. Second, a modulation of the extraction efficiency was observed with NW periodicity P1 as shown in Figure 2B. For example, at 1000 nm, next. from the 200-nm-P1 and 600-nm-P1 SiC NW-PC structures are 24% and 37%, respectively. Similar trends were also observed in SiC:O systems (Figure 2B, gray data points).

The extraction efficiency enhancement of the dipole’s spontaneous emission in NW-PC structures is expected because of the suppression of in-plane emission modes (guided modes). By eliminating guided modes at transition frequencies, spontaneous emission can be coupled efficiently to free space vertical modes, resulting in substantially enhanced extraction efficiency [19], [30]. Therefore, radiation originating from NW-PC structures results in enhanced light emission in the vertical direction. This behavior is supported by our spatial distribution calculation of the extracted power radiation (~|E|2) at 1540 nm (Figure 2C). In the case of the thin-film slab, a significant portion of the radiation is confined within the slab (x-direction in this simulation) (Figure 2C-i). Conversely in NW-PC, guided modes are eliminated as it is presented in Figure 2C-ii and Figure S2 (Supplementary Material). For the 600-nm-P1 structures, 1540-nm emission is inhibited along the x-direction more prominently than other NW-PC structures, as the 1540-nm wavelength falls within the computed incomplete photonic bandgap of the structure (Figure 2C-iii). Only the 600-nm-P1 structures among our fabricated arrays exhibited a photonic bandgap containing 1540 nm. Nevertheless, due to the ultrathin geometry of nanostructures, the extraction efficiency is enhanced due to lack of total internal reflection even if there is no bandgap present containing the frequency of interest [19].

3.2 Morphological and structural properties of materials

Fourier transform infrared (FTIR) spectroscopy measurements showed complete crystallization of the as-deposited amorphous SiC materials at temperatures ≥1000°C, in accordance with our previously reported results (Figure 3A-i) [31]. In this respect, the line shape of the Si-C stretching mode (~800 cm−1) of the as-deposited material transformed from Gaussian to Lorentzian after annealing at 1100°C, which was accompanied by a substantial narrowing of the full width at half maximum (FWHM) (~30 cm−1), comparable to values for high-quality crystal SiC [32]. The surface roughness of as-deposited and annealed materials was analyzed by atomic force microscopy. Both materials have extremely flat surfaces with a mean surface roughness of 1–2 Å (measured from several 1 μm2 areas). X-ray diffraction and high-resolution transmission electron microscope (HR-TEM) analyses agreed with the FTIR results, revealing the formation of cubic 3C-SiC. The NWs were polycrystalline, with an average nanocrystal grain size of ~5 nm as presented in Figure 3A-ii.

Figure 3: Structural and PL properties.(A) i. FTIR absorption spectra of the Si-C stretching mode for the as-deposited (AD) and 1100°C-annealed 20-nm thin films showing the transformation from amorphous to polycrystalline phase. ii. Representative HR-TEM image of annealed NW-PC showing the average grain size of ~5 nm with 3C-SiC crystalline phase. (B) Visible PL spectra from NW-PC structures under 476 nm laser excitation. PL of the same wavelength range from outside of the NWs is shown in orange. (C) Characteristic room-temperature Er3+ emissions at ~1540 nm and ~1555 nm in NW-PC structures under 488-nm (solid lines) and 476-nm (dotted line) excitations. (D) Photoluminescence excitation (PLE) spectra at 1540-nm emission from NW-PC (square) and SiO2 (circle symbols) under a range of laser excitation wavelengths from 458 nm to 514 nm. Error bars are not depicted as the errors are smaller than the symbol size.
Figure 3:

Structural and PL properties.

(A) i. FTIR absorption spectra of the Si-C stretching mode for the as-deposited (AD) and 1100°C-annealed 20-nm thin films showing the transformation from amorphous to polycrystalline phase. ii. Representative HR-TEM image of annealed NW-PC showing the average grain size of ~5 nm with 3C-SiC crystalline phase. (B) Visible PL spectra from NW-PC structures under 476 nm laser excitation. PL of the same wavelength range from outside of the NWs is shown in orange. (C) Characteristic room-temperature Er3+ emissions at ~1540 nm and ~1555 nm in NW-PC structures under 488-nm (solid lines) and 476-nm (dotted line) excitations. (D) Photoluminescence excitation (PLE) spectra at 1540-nm emission from NW-PC (square) and SiO2 (circle symbols) under a range of laser excitation wavelengths from 458 nm to 514 nm. Error bars are not depicted as the errors are smaller than the symbol size.

3.3 Er3+ PL properties in SiC NW-PC structures

We investigated the room-temperature (RT) visible and Er3+ emission properties of undoped and Er-doped NW-PC structures, respectively, by PL spectroscopy using a micro-PL (μPL) system (Figure S3 in Supplementary Material) at different laser visible excitation wavelengths (e.g. 458 nm, 476 nm, 488 nm, and 514.5 nm). Figure 3B shows the matrix-related PL emission spectra of undoped NW-PC structures, with a peak PL emission at ~550 nm in concurrence with our previous studies [24], [25]. Pertaining to Er3+ emission, the 476-nm excitation is non-resonant excitation, while the 488-nm excitation corresponds to resonant transitions from the ground state 4I15/2 to the 4F7/2 excited state in Er3+. The characteristic Er3+ emission, due to intra-4f transitions 4I13/24I15/2, [18] at ~1540 nm and ~1555 nm were observed in both SiC and SiC:O NW-PC as shown in Figure 3C (Er dose: 1014 cm−2). Er3+-1540-nm PL emission was observed using various visible pumping wavelengths in NW-PCs, demonstrating the structures’ broadband excitation characteristics (Figure 3D). Conversely, for the SiO2:Er system, only resonant transitions [33] to the excited states of Er3+ occurred around the excitation wavelengths of 458 nm, 488 nm, and 514.5 nm in agreement with other reports [34], [35].

NW-PC structures with larger P1 periodicity, resulting in a lower density of NWs, would be expected to have lower visible PL due to their volumetric difference and also lower Er3+-1540-nm PL due to the corresponding decreased number of total Er3+ ions in the NW-PC. This volumetric effect was observed in the visible PL emission at 550 nm, where PL decreased for increasing P1 periodicity (Figure 4A). However, the opposite trend was observed for the Er3+ PL. The structures with the smallest periodicity, P1=200 nm, exhibited the lowest Er3+ PL among the studied NW-PC structures (Figure 4B). After accounting for the decrease in the number of Er3+ ions in the air region of the different NW-PC structures (Section S4 and S5 for more details), the Er3+ PL increases in a monotonic fashion by a factor of about five as P1 increases from 200 to 600 nm (Figure 4C). Furthermore, the modulation of the Er3+ PL with P1 periodicity presented a direct correlation with the FDTD-computed extraction efficiency in NW-PC (Figure 4B). This correlation was observed for equivalent SiC and SiC:O NW-PC structures, and we observed the highest Er3+ PL intensity in the 600-nm-P1 SiC:O structures.

Figure 4: Volumetric effects on PL properties of NWs.(A) PL intensity (counts per second [cps]) at 550 nm as a function of P1 periodicity. (B) Measured PL intensity at 1540 nm (green circle) and FDTD-computed extraction efficiency (pink triangle) for different NW-PC structures. Error bars are not depicted as the errors are smaller than the symbol size. Inset: PL intensity mapping of NW-PC with P2×P1=150×600. The scale bar is 500 nm. (C) Er3+ PL intensity in SiC NW-PC and SiC:O NW-PC as a function of P1; the difference of implanted Er ions was accounted for among the NW-PC. Three different NW-PC structures were measured for each P1. The gray data points represent the Er3+PL in SiC:O NW-PC without the volumetric normalization.
Figure 4:

Volumetric effects on PL properties of NWs.

(A) PL intensity (counts per second [cps]) at 550 nm as a function of P1 periodicity. (B) Measured PL intensity at 1540 nm (green circle) and FDTD-computed extraction efficiency (pink triangle) for different NW-PC structures. Error bars are not depicted as the errors are smaller than the symbol size. Inset: PL intensity mapping of NW-PC with P2×P1=150×600. The scale bar is 500 nm. (C) Er3+ PL intensity in SiC NW-PC and SiC:O NW-PC as a function of P1; the difference of implanted Er ions was accounted for among the NW-PC. Three different NW-PC structures were measured for each P1. The gray data points represent the Er3+PL in SiC:O NW-PC without the volumetric normalization.

To further our understanding of the NW-PC effect on Er3+ emission, we studied a representative 600-nm-P1 SiC:O NW-PC device (Er dose: 1013 cm−2). It is worth highlighting that with the experimental μPL setup used in this study, the Er3+ PL emission was not detected from the thin film without a PC structure (Ref.) at pumping power densities below 40 kW/cm2. This power value can be defined as the threshold power for detecting Er3+ PL from the thin film. In contrast, Er3+ PL from the 600-nm-P1 NW-PC was measurable with a pumping power as low as 0.9 kW/cm2. This behavior suggests that, collectively, the pumping and emission-extraction efficiencies of Er3+ in NW-PC are higher compared to their thin-film analog. At the threshold power, and accounting for the same number of Er ions (Supplementary Material), the Er3+ PL collected from the 600-nm-P1 NW-PC was found to be approximately 60 times higher than thin film (Figure 5A). Moreover, the PL intensity ratio between the NW-PC and thin-film samples increases monotonically with decreasing Er dose. For example, the PL ratio was approximately 17 and 35 for the 1014 cm−2 and 5×1013 cm−2 doses, respectively (inset in Figure 5A). The time-resolved PL (TRPL) of the 1540-nm emission revealed a luminescence lifetime, τ, of ~2.7 ms in 600-nm-P1 NW-PC structures.

Figure 5: Comparison of 1540-nm emission characteristics between Er3+ in 600-nm-P1 NW-PC and thin film.Photostability of Er3+ emission in 600-nm-P1 NW-PC. (A) Room-temperature steady-state Er3+ PL spectra in 600-nm-P1 NW-PC and thin film (Ref.) under 488-nm excitation (pumping power density: 40 kW/cm2; Er dose: 1×1013 cm−2; the difference of implanted Er ions was accounted for between the NW-PC and thin film). The Er3+ PL intensity in thin film was multiplied by 20 for presentation purposes. Inset: Ratio of Er3+ PL intensity between NW-PC and thin film as a function of Er dose measured at their respective threshold pumping powers. (B) Er3+ PL emission peak intensity as a function of pumping power density from i. 600-nm-P1 NW-PC and ii. thin film (Er dose: 1013 cm−2). The dashed lines indicate fits to the data using Eq. (1). (C) i. Peak position and ii. FWHM of Er3+ PL spectra as a function of pumping power. iii. Er3+ PL time traces with sampling bins of Δt=10 ms and Δt=100 ms (excitation power density: 250 kW/cm2, Er: 1×1014 cm−2). (D) Polar plots of the integrated Er3+ PL peak area as a function of the emission polarization from i. 600-nm-P1 NW-PC and ii. thin film. The polarization angle of the excitation source was kept parallel along to the NWs’ length.
Figure 5:

Comparison of 1540-nm emission characteristics between Er3+ in 600-nm-P1 NW-PC and thin film.

Photostability of Er3+ emission in 600-nm-P1 NW-PC. (A) Room-temperature steady-state Er3+ PL spectra in 600-nm-P1 NW-PC and thin film (Ref.) under 488-nm excitation (pumping power density: 40 kW/cm2; Er dose: 1×1013 cm−2; the difference of implanted Er ions was accounted for between the NW-PC and thin film). The Er3+ PL intensity in thin film was multiplied by 20 for presentation purposes. Inset: Ratio of Er3+ PL intensity between NW-PC and thin film as a function of Er dose measured at their respective threshold pumping powers. (B) Er3+ PL emission peak intensity as a function of pumping power density from i. 600-nm-P1 NW-PC and ii. thin film (Er dose: 1013 cm−2). The dashed lines indicate fits to the data using Eq. (1). (C) i. Peak position and ii. FWHM of Er3+ PL spectra as a function of pumping power. iii. Er3+ PL time traces with sampling bins of Δt=10 ms and Δt=100 ms (excitation power density: 250 kW/cm2, Er: 1×1014 cm−2). (D) Polar plots of the integrated Er3+ PL peak area as a function of the emission polarization from i. 600-nm-P1 NW-PC and ii. thin film. The polarization angle of the excitation source was kept parallel along to the NWs’ length.

To obtain more insight into these compelling Er3+ PL behaviors in the NW-PC, we measured the Er3+ PL emission as a function of excitation power (Figures 5B and S6). After subtracting the linear background and the detector dark counts, we fitted the 1540-nm emission data to the following equation [15], [36],

(1)I(P)=Isat1+PsatP

where I(P) is the Er3+ PL intensity at a given excitation power P, Isat is the saturation intensity, and Psat is the excitation power required to yield half of Isat. Figure 5B-i presents the results from a representative 600-nm-P1 NW-PC nanostructure, revealing Isat=4244±114 cps and Psat=~8.3 kW/cm2. Isat and Psat were also extracted for a representative thin-film sample, giving Isat=192±10 cps and Psat=~150 kW/cm2 (Figure 5B-ii). These values indicate an approximate 20-fold increase in both pumping efficiency and Er3+ emission output flux (Isat) in NW-PC. Furthermore, Eq. (1) can be expressed as a function of the photon flux of incoming photons, φ, by [18], [35],

(2)I=Isat1+1σφτ

where σ is the effective excitation cross-section and τ is the luminescence lifetime that was experimentally determined in TRPL. Isat is given by Isat=N*/τr, where τr and N* are, respectively, the radiative lifetime and the concentration of the optically active erbium ions. We then fitted the experimental power dependence data to Eq. (2) using the observed Er PL lifetime, which yielded an effective cross-section value of 2.2×10−18 cm−2 for the NW-PC. This value is one order larger than that of the thin-film sample (2.8×10−19 cm−2), and two orders of magnitude higher than the typical range of ~10−20–10−21 cm2 for Er-doped materials [35], [37], [38].

The matrix-related PL spectrum of the NW-PC devices is characterized by a broadband luminescence feature in the spectral range of 500–750 nm, with the dominant PL occurring around 550 nm wavelength. To this end, some of the Er3+ transitions, such as 4I15/22H11/2 (514–520 nm), 4I15/24S3/2 (550 nm), and 4I15/24F9/2 (660 nm), are in resonance with the broadband PL of the NW-PC devices [33]. The PL and PLE data thus demonstrate an efficient energy transfer mechanism between the nanostructured matrix and Er3+ ions, as the Er emission can be sensitized by the nanostructured matrix. Effective light trapping by multiple scattering events can also increase absorption probability of the incident light in the NW-PC devices [7].

Photostability under high pumping is important for photonic applications. The peak position and lineshape (FWHM) of the observed Er3+ PL emission in NW-PC remained unchanged with pumping power (Figure 5C-i,ii), showing good photostability of the Er-doped nanophotonic structures. Furthermore, we examined the photostability of Er3+ PL under high pumping power as a function of time. The Er3+ PL time-traces on two different timescales are shown in Figure 5C-iii. The Er3+ PL count rate remained constant over several minutes with no indication of photobleaching. Furthermore, the large geometric anisotropy of the ultrathin NW also results into anisotropy of the emission and absorption properties of the emitters [39], [40]. As presented in Figure 5D-i, the Er3+ emission from NW-PC was found to be predominantly polarized along the length of the NWs, with an emission polarization degree, ρem, of ~0.4. Even if the dipoles in a NW are randomly oriented, the emitted light from the NW will be preferentially oriented along the NW axis and ρem is expressed as [41]:

(3)ρem=|III+I|

where I or I is the intensity of the most (polarizer parallel to NWs) or least (polarizer perpendicular to NWs) intense PL, respectively. In contrast, Er3+ PL emission from the thin film appears completely isotropic (independent from emission polarization direction) (Figure 5D-ii).

3.4 Optically active Er3+ concentration calculation and defect density

To estimate the concentration of the optically active Er3+ ions in the NW-PC and thin film, we used the following approach described in more detail in Section S7. In the linear pumping regime (low pumping power), where the excitation rate constant, σφ, is much smaller than the de-excitation rate constant, 1/τ (σφτ<<1), Eq. (2) can be approximated by

(4)I=nextσφτN*τr

At saturation (high pumping power), where σφτ>>1, the PL intensity is equal to

(5)Isat=nextN*τr

As discussed, at saturation the Isat ratio between the NW-PC and thin film was ~22. We calculated the radiative lifetime (τr,NW=Γr,NW1), reciprocal of the radiative PL decay rate Γr, of the NW-PC structure relative to the lifetime in the reference thin film (τr,ref=Γr,ref1) using FDTD simulation [30], [42]. In this respect, the total power radiated by the dipole within the NW-PC (PNW) is proportional to the radiative spontaneous emission rate, Γr,NW, and equivalently for the reference thin film. Thus, in the case of the 600-nm-P1 NW-PC and thin film, we calculated the emission rates ratio as

Γr,NWΓr,ref=(τr,NWτr,ref)1=PNWPref0.39

Using Eq. (5) and the extraction efficiency ratio between the NW-PC and thin film, as determined by FDTD computations, we estimated [43] the optically active Er3+ concentration in NW-PC to be ~12.4 times higher than that in the thin film. Considering the decrease in the number of Er3+ ions in the air region for the NW-PC structures (~0.083× decrease), the total number of Er3+ was found to be approximately ~3% more in NW-PC. Nevertheless, we observed that the Er3+ PL was saturated at much lower power in the NW-PC structures, indicating a more efficient mechanism of pumping Er3+ compared to the thin film. As mentioned in the Introduction, an enhanced pumping efficiency is important for solid-state photonics and quantum communications applications [9], [11].

The integration of rare-earth ions in wide-bandgap ultrathin NWs can offer supplemental benefits. It can be argued that Auger and energy back-transfer effects are appreciably reduced in wide-bandgap SiC and SiC:O due to a low concentration of free carriers and the large mismatch between their bandgap and the first excited state of Er3+, respectively [24], [35]. Confinement effects on electron-phonon interaction may be expected primarily due to the absence of low-energy acoustic phonon modes in ultrathin NWs, resulting in different luminescence dynamics of Er3+ at low temperatures [44]. Moreover, in ultrathin NWs, a decreased bulk defect density [45] and a strain-relaxed environment in the material is expected [46], [47] and, hence, a smaller probability of nonradiative sites within the ion’s recombination volume. Thus, in ultrathin NW-PC having well-passivated surfaces, excited Er3+ is expected to be exposed to a smaller number of bulk nonradiative recombination sites compared to its bulk analog, which can lead to an increased PL lifetime [45]. From deconvolution of the PL spectra at RT, we observed that spectral line width is narrower in NW-PC structures (~10 nm) compared to the thin-film sample (~18 nm), which can be due to reduction in defect density and/or reduction in the number of optically active ions.

To access the influence of nonradiative quenching processes in the NW-PC structures, we performed TRPL of the 1540-nm emission at 77 K. As can be seen in Figure 6A, the observed Er3+ luminescence lifetime did not change (2.7±0.1 ms) with respect to its 300 K value. The Er3+ PL dynamics of the NW-PC implies a good-quality material, with significantly low defect density and a well-passivated surface. It may be thus inferred that the higher concentration of optically active Er3+ in NW-PC can be the result of reduction of bulk defect-density, as compared to the Er3+ PL dynamics of the thin film (Figure S8), and a more effective thermal annealing process in forming optically active Er ions due to the NW geometry. Additionally, with decreasing Er dose, the number of Er3+ ions in NW-PC also decreases, which results in an even lower probability of a defect center being within the ions recombination volume. This may explain the trend of increased PL intensity ratio with decreasing Er dose between the NW-PC and thin film mentioned in Section 3.3 (inset of Figure 5A).

Figure 6: Effect of NW-PC structure on Er3+ PL emission dynamics.(A) Er3+ PL (~1540 nm) decay in the 600-nm-P1 NW-PC structure under 488-nm pulsed excitation at 77 K and 300 K (red points: 300 K [RT]; green points: 77 K). The Er3+ PL lifetime in NW-PC, as obtained from single exponential fitting, was calculated to be ~2.7 ms for both 77 K and 300 K. The solid lines indicate fits to the data. (B) PL decay of 550-nm emission from the NW-PC structures (legend: P1) with instrument response function (IRF) in gray. (C) Room-temperature Er3+ PL decay at 1540 nm from different NW-PC structures (legend: P1). Solid lines represent single exponential fits. (D) Er3+ PL lifetimes for the different NW-PC structures and for a single NW (orange). (E) Computed relative radiative lifetime from NW-PC with different P1 and a single NW. Data points corresponding to the NW-PC with periodicity values beyond our fabricated structures are shown under the shaded region.
Figure 6:

Effect of NW-PC structure on Er3+ PL emission dynamics.

(A) Er3+ PL (~1540 nm) decay in the 600-nm-P1 NW-PC structure under 488-nm pulsed excitation at 77 K and 300 K (red points: 300 K [RT]; green points: 77 K). The Er3+ PL lifetime in NW-PC, as obtained from single exponential fitting, was calculated to be ~2.7 ms for both 77 K and 300 K. The solid lines indicate fits to the data. (B) PL decay of 550-nm emission from the NW-PC structures (legend: P1) with instrument response function (IRF) in gray. (C) Room-temperature Er3+ PL decay at 1540 nm from different NW-PC structures (legend: P1). Solid lines represent single exponential fits. (D) Er3+ PL lifetimes for the different NW-PC structures and for a single NW (orange). (E) Computed relative radiative lifetime from NW-PC with different P1 and a single NW. Data points corresponding to the NW-PC with periodicity values beyond our fabricated structures are shown under the shaded region.

The much shorter Er3+ PL lifetime in thin films (Figure S8) can be attributed to an increase in nonradiative de-excitation paths for Er ions in the thin-film matrix in agreement with other reports [48], [49]. Furthermore, a reduction in the effective refractive index in NW-PC and confinement of the Er ions within the NW may reduce the photon density of states for emission and thereby reduce the transition rate and correspondingly increase the Er luminescence lifetime [35]. The effective refractive index in NW-PC is expected to be between that of air (n=1) and thin film (n=2.2).

As discussed, we observed that the Er3+ PL from NW-PC was modulated with P1 periodicity and was correlated with the computed extraction efficiency trend at 1540 nm (Figure 4B). In conjunction to the different behavior between the steady-state PL at 550-nm and 1540-nm emissions from NW-PC structures, we also observed different trends in the luminescence dynamics at these wavelengths. Average lifetime at 550 nm was found to be the same, ~680±40 ps, for all the NW-PC structures (Figure 6B). However, luminescence lifetime of Er-PL at 1540 nm was found to be approximately the same, 2.3±0.1 ms, only for NW-PCs with 300-, 400-, and 500-nm-P1 periodicity (Figure 6C,D). For the 600-nm-P1 NW-PC structures, in agreement with the computed radiative lifetimes (Figure 6E), we observed a longer lifetime (2.7±0.1 ms), which suggests that an additional physical mechanism is contributing to the modulation besides nanostructuring and reduced defect density effects. The spontaneous emission rate of PC structures is predicted to decrease when the emission spectrum lies within the photonic bandgap [20], [30], as in the case of 600-nm-P1 NW-PC structures, yielding longer luminescence lifetime. We observed a ~15% and ~30% relative difference in lifetimes from experimental (Figure 6D) and computational (Figure 6E) results, respectively, between 600-nm-P1 and the three other NW-PC that do not have a photonic bandgap containing 1540 nm. As expected for our 1D 600-nm-P1 NW-PC structures, the increase in PL lifetime was less pronounced compared to the increase observed in two-dimensional photonic bandgap materials (~80% in Ref. [30]).

The increase of the Er3+ PL lifetime in NW-PC, decrease of its PL emission rate as ΓNW=τNW1=Γr,NW+Γnr,NW, is coupled to the above mechanisms (Γnr is the nonradiative decay rate). ΓNW is influenced by the complex interplay of the radiating Er3+(Γr,NW=τr,NW1) and its interactions with the nanophotonic structure (e.g. NW geometry, nanostructuring effect, photonic bandgap), as well as the surrounding defectivity (e.g. ion’s energy transfer to competing nonradiative decay channels, Γr,NW=τr,NW1).

4 Conclusions

In summary, we present a new nanofabrication scheme for the integration of rare-earth Er ions into solid-state scalable SiC nanostructures. These nanostructures encompass a new class of tailorable ultrathin PC structures, thus enabling the engineering of the spontaneous emission characteristics of Er3+. Both FDTD calculations and PL measurements revealed that the Er-induced emission efficiency can be enhanced by tailoring the geometry of the NW-PC structure. We experimentally observed substantial enhancements for both the RT 1540-nm emission and absorption cross-section in NW-PC. This holistic approach can cultivate an alternative pathway toward the development of scalable nanophotonics (e.g. integration into photonic circuity and cavity) at telecom wavelengths that could be benefited by the controlled placement of rare-earth Er3+ ions and the modification of the ion’s emission in scalable SiC nanophotonic structures.

5 Methods

5.1 Synthesis of NW array structures

Silicon (100) wafers were spin-coated with HSQ negative-tone resist at 1000 rpm followed by a soft-bake, yielding an approximately 130-nm-thick HSQ layer. We exposed the HSQ layer using a Vistec VB300 (Vistec Electron Beam GmbH, Jena, Germany) and a Voyager Raith (Raith GmbH, Dortmund, Germany) electron-beam lithography tool using line patterns created in the Layout Editor. Following the exposure, development was performed in tetramethylammonium hydroxide, yielding HSQ ribbon arrays. We conformally deposited 20 nm of SiC/SiC:O onto the HSQ ribbon arrays using a home-built thermal CVD system at 800°C. CVD-742 (1,1,3,3-tetramethyl-1,3-disilacyclobutane) from Starfire Systems or TSCH (1,3,5-trisilacyclohexane) from Gelest Inc. (Morrisville, PA, USA) was used as the silicon and carbon source. After the SiC or SiC:O growth, anisotropic fluorine-based (combination of CHF3 and CF4 gases) reactive ion etch was performed using a Plasma-Therm Versalock 700 (Plasma-Therm, St. Petersburg, FL, USA) to remove the top SiC or SiC:O layer, leaving the sidewall layers intact and exposing the HSQ ribbon template. A wet-etch using BHF acid was then carried out to remove the HSQ, yielding ultrathin NWs synthesized in a self-aligned manner.

5.2 Device modeling

The calculations of the extraction efficiency and the spatial distribution of the dipole radiation were carried out using 3D FDTD simulations (Lumerical Inc.). A y-polarized single dipole source was positioned at the center of a uniform slab (thin film) and of a NW-PC structure (Figure 2A). The spatial dimension of the simulation window was varied to keep the total number of NWs (21 pairs) the same for all structures. We used a fixed grid size of 1 nm and a perfectly matched layer as the boundary condition in all directions [30]. We used a monitor (parallel to the xy-plane) to calculate the transmitted flux through the top surface of the slab. It consisted of one 2 μm×2 μm xy-plane at z=500 nm from the surface of the material (Figure 2A).

The position and dimensions of the monitor were kept the same for all simulations to ensure equivalent solid angle detection. We then calculated the total power injected by the dipole (Pin) by integrating the time averaged Poynting flux over the six surfaces of a cube enclosing the radiating dipole. The extracted power (Pout) was calculated from the power transmitted through the monitor plane. The extraction efficiency, ηext, is defined as the ratio of Pout/Pin [19], [30]. The index of refraction, n, thickness, and geometry values for the thin film and NW-PC structures, used in the simulations, were experimentally determined by scanning electron microscopy (SEM) and spectroscopic ultraviolet-visible ellipsometry measurements [J.A. Woollam (Orlando, FL, USA) RC2 dual rotating compensator ellipsometer].

5.3 PL characterization

A home-built μPL system – composed of an argon laser (Beamlock 2065-7S, Spectra-Physics, Santa Clara, CA, USA) coupled to an FLSP920 spectrometer from Edinburgh Instruments (Livingston, UK), a Triax-550 spectrometer from Horiba Jobin Yvon (Kyoto, Japan), and liquid nitrogen cooled Ge/InGaAs detectors – was utilized for PL and power-dependence PL (PDPL) measurements. We performed the PL and PDPL measurements at room temperature using an argon laser (Beamlok 2065-7S, Spectra-Physics, Santa Clara, CA, USA), to optically excite the Er3+ ions, through a Zeiss 50× [numerical aperture (NA)=0.85] objective lens. TRPL studies were conducted in the FLSP920 spectrometer utilizing a multichannel scaling technique. An acousto-optic modulator (Gooch & Housego, Fremont, CA, USA) was used to create laser pulses with a ~500 μs FWHM and 50 Hz repetition rate for the NW-PC structures and 100 Hz for the thin film. For PL measurements at 77 K, a close-system cold-finger cryostat with Mitutoyo NIR 50× objective lens (NA=0.55) was used in the same spectrometer. For the emission-polarization anisotropy measurements, PL spectra were collected at different angles of a polarizer placed in front of the tube lens.

Acknowledgments

This work was supported by the National Science Foundation through grant no. ECCS-1842350. This work was also supported by the College of Nanoscale Science and Engineering of SUNY Polytechnic Institute and the Research Foundation for the State University of New York. Their support is gratefully acknowledged. The authors gratefully acknowledge engineer C. Johnson for the TEM sample preparation and B. Harrington for helping in conducting experiments during his undergraduate research (REU) project.

  1. Author contributions: S.G. perceived the experiments; V.N. and N.T. carried out the nanofabrication and N.T., S.G, and M.H. organized the ion implantation of the NW-PC nanostructures; N.T. and V.N. performed the FDTD calculations; N.T., V.K, and E.C. conducted the structural characterization of the NW-PC and thin-film samples; N.T. and A.K. performed the PL experiments and analyzed the experimental data; S.G and N.T wrote the main manuscript text.

References

[1] Kwiat M, Elnathan R, Pevzner A, et al. Highly ordered large-scale neuronal networks of individual cells – toward single cell to 3D nanowire intracellular interfaces. ACS Appl Mater Interfaces 2012;4:3542–9.10.1021/am300602eSearch in Google Scholar PubMed

[2] Koenderink AF, Alu A, Polman A. Nanophotonics: shrinking light-based technology. Science 2015;348:516–21.10.1126/science.1261243Search in Google Scholar PubMed

[3] Wu R, Zhou K, Yue CY, Wei J, Pan Y. Recent progress in synthesis, properties and potential applications of SiC nanomaterials. Prog Mater Sci 2015;72:1–60.10.1016/j.pmatsci.2015.01.003Search in Google Scholar

[4] Peretz-Soroka H, Pevzner A, Davidi G, et al. Optically-gated self-calibrating nanosensors: monitoring pH and metabolic activity of living cells. Nano Lett 2013;13:3157–68.10.1021/nl401169kSearch in Google Scholar PubMed

[5] He Y, Fan C, Lee S-T. Silicon nanostructures for bioapplications. Nano Today 2010;5:282–95.10.1016/j.nantod.2010.06.008Search in Google Scholar

[6] Peng K-Q, Wang X, Li L, Hu Y, Lee S-T. Silicon nanowires for advanced energy conversion and storage. Nano Today 2013;8:75–97.10.1016/j.nantod.2012.12.009Search in Google Scholar

[7] Priolo F, Gregorkiewicz T, Galli M, Krauss TF. Silicon nanostructures for photonics and photovoltaics. Nat Nanotechnol 2014;9:19–32.10.1038/nnano.2013.271Search in Google Scholar PubMed

[8] Babinec TM, Hausmann BJM, Khan M, et al. A diamond nanowire single-photon source. Nat Nanotechnol 2010;5:195–9.10.1038/nnano.2010.6Search in Google Scholar PubMed

[9] Lohrmann A, Johnson BC, McCallum JC, Castelletto S. A review on single photon sources in silicon carbide. Rep Prog Phys 2017;80:34502.10.1088/1361-6633/aa5171Search in Google Scholar PubMed

[10] Thiel CW, Böttger T, Cone RL. Rare-earth-doped materials for applications in quantum information storage and signal processing. J Lumin 2011;131:353–61.10.1016/j.jlumin.2010.12.015Search in Google Scholar

[11] Simon C, Afzelius M, Appel J, et al. Quantum Memories. Eur Phys J D 2010;58:1–22.10.1140/epjd/e2010-00103-ySearch in Google Scholar

[12] Saglamyurek E, Jin J, Verma VB, et al. Quantum storage of entangled telecom-wavelength photons in an erbium-doped optical fibre. Nat Photonics 2015;9:83–7.10.1038/nphoton.2014.311Search in Google Scholar

[13] Zhong T, Kindem JM, Bartholomew JG, et al. Nanophotonic rare-earth quantum memory with optically controlled retrieval. Science 2017;357:1392–5.10.1126/science.aan5959Search in Google Scholar PubMed

[14] Aharonovich I, Toth M. Silicon carbide goes quantum. Nat Phys 2014;10:93–4.10.1038/nphys2858Search in Google Scholar

[15] Fuchs F, Stender B, Trupke M, et al. Engineering near-Infrared Single-Photon Emitters with Optically Active Spins in Ultrapure Silicon Carbide. Nat Commun 2015;6:1–7.10.1038/ncomms8578Search in Google Scholar PubMed

[16] Castelletto S, Johnson BC, Zachreson C, et al. Room temperature quantum emission from cubic silicon carbide nanoparticles. ACS Nano 2014;8:7938–47.10.1021/nn502719ySearch in Google Scholar PubMed

[17] Böttger T, Thiel CW, Sun Y, Cone RL. Optical decoherence and spectral diffusion at 1.5 m in Er3+:Y2SiO5 versus magnetic field, temperature, and Er3+ concentration. Phys Rev B 2006;73:075101.10.1103/PhysRevB.73.075101Search in Google Scholar

[18] Gallis S, Huang M, Kaloyeros AE. Efficient energy transfer from silicon oxycarbide matrix to Er ions via indirect excitation mechanisms. Appl Phys Lett 2007;90:161914.10.1063/1.2730583Search in Google Scholar

[19] Fan S, Villeneuve PR, Joannopoulos JD, Schubert EF. High extraction efficiency of spontaneous emission from slabs of photonic crystals. Phys Rev Lett 1997;78:3294–7.10.1103/PhysRevLett.78.3294Search in Google Scholar

[20] Noda S, Fujita M, Asano T. Spontaneous-emission control by photonic crystals and nanocavities. Nat Photonics 2007;1:449–58.10.1038/nphoton.2007.141Search in Google Scholar

[21] Dibos AM, Raha M, Phenicie CM, Thompson JD. Atomic source of single photons in the telecom band. Phys Rev Lett 2018;120:243601.10.1103/PhysRevLett.120.243601Search in Google Scholar PubMed

[22] Schmidt V, Gosele U. Materials science: how nanowires grow. Science 2007;316:698–9.10.1126/science.1142951Search in Google Scholar PubMed

[23] Kwiat M, Cohen S, Pevzner A, Patolsky F. Large-scale ordered 1D-nanomaterials arrays: assembly or not? Nano Today 2013;8:677–94.10.1016/j.nantod.2013.12.001Search in Google Scholar

[24] Nikas V, Tabassum N, Ford B, Smith L, Kaloyeros AE, Gallis S. Strong visible light emission from silicon-oxycarbide nanowire arrays prepared by electron beam lithography and reactive ion etching. J Mater Res 2015;30:3692–9.10.1557/jmr.2015.346Search in Google Scholar

[25] Tabassum N, Kotha M, Kaushik V, et al. On-demand CMOS-compatible fabrication of ultrathin self-aligned SiC nanowire arrays. Nanomaterials 2018;8:906.10.3390/nano8110906Search in Google Scholar PubMed PubMed Central

[26] Tabassum N, Nikas V, Ford B, Huang M, Kaloyeros AE, Gallis S. Time-resolved analysis of the white photoluminescence from chemically synthesized SiCxOy thin films and nanowires. Appl Phys Lett 2016;109:43104.10.1063/1.4959834Search in Google Scholar

[27] Ford B, Tabassum N, Nikas V, Gallis S. Strong photoluminescence enhancement of silicon oxycarbide through defect engineering. Materials 2017;10:446.10.3390/ma10040446Search in Google Scholar PubMed PubMed Central

[28] Pacheco JL, Singh M, Perry DL, et al. Ion implantation for deterministic single atom devices. Rev Sci Instrum 2017;88:123301.10.1063/1.5001520Search in Google Scholar PubMed

[29] Shinada T, Okamoto S, Kobayashi T, Ohdomari I. Enhancing semiconductor device performance using ordered dopant arrays. Nature 2005;437:1128–31.10.1038/nature04086Search in Google Scholar PubMed

[30] Fujita M, Takahashi S, Asano T, et al. Controlled spontaneous-emission phenomena in semiconductor slabs with a two-dimensional photonic bandgap. J Opt A Pure Appl Opt 2006;8:S131–8.10.1088/1464-4258/8/4/S12Search in Google Scholar

[31] Gallis S, Nikas V, Huang M, Eisenbraun E, Kaloyeros AE.Comparative study of the effects of thermal treatment on the optical properties of hydrogenated amorphous silicon-oxycarbide. J Appl Phys 2007:102.10.1063/1.2753572Search in Google Scholar

[32] Madapura S. Heteroepitaxial growth of SiC on Si (100) and (111) by chemical vapor deposition using trimethylsilane. J Electrochem Soc 1999;146:1197–202.10.1149/1.1391745Search in Google Scholar

[33] Miniscalo WJ. Erbium-doped glasses for fiber amplifiers at 1500 nm. J Lightwave Technol 1991;9:234–50.10.1109/50.65882Search in Google Scholar

[34] Kenyon AJ. Recent developments in rare-earth doped materials for optoelectronics. Prog Quant Electron 2002;26:225–84.10.1016/S0079-6727(02)00014-9Search in Google Scholar

[35] Elliman RG, Wilkinson AR, Kim T-H, Sekhar PK, Bhansali S. Optical emission from erbium-doped silica nanowires. J Appl Phys 2008;103:104304.10.1063/1.2924420Search in Google Scholar

[36] Kurtsiefer C, Mayer S, Zarda P, Weinfurter H. Stable solid-state source of single photons. Phys Rev Lett 2000;85:290–3.10.1103/PhysRevLett.85.290Search in Google Scholar PubMed

[37] Sardar DK, Russell III, CC, Gruber JB, Allik TH. Absorption intensities and emission cross sections of principal intermanifold and inter-Stark transitions of Er3+(4f11) in polycrystalline ceramic garnet Y3Al5O12. J Appl Phys 2005;97:123501.10.1063/1.1928327Search in Google Scholar

[38] Suh K, Lee M, Chang JS, et al. Cooperative upconversion and optical gain in ion-beam sputter-deposited ErxY2-xSiO5 waveguides. Opt Exp 2010;18:7724–31.10.1364/OE.18.007724Search in Google Scholar PubMed

[39] Ruda HE, Shik A. Polarization-sensitive optical phenomena in semiconducting and metallic nanowires. Phy Rev B 2005;72:115308.10.1103/PhysRevB.72.115308Search in Google Scholar

[40] Rivas JG, Muskens OL, Borgström MT, Diedenhofen SL, Bakkers EPAM. Optical anisotropy of semiconductor nanowires. In One-dimensional nanostructures. Wang ZM, Springer, New York, 2008:127–45.10.1007/978-0-387-74132-1_6Search in Google Scholar

[41] Wang , Zhiming M. ed. One-dimensional nanostructures. Vol. 3. Springer Science & Business Media, 2008.10.1007/978-0-387-74132-1Search in Google Scholar

[42] Inam FA, Grogan MDW, Rollings M, et al. Emission and nonradiative decay of nanodiamond NV centers in a low refractive index environment. ACS Nano 2013;7:3833–43.10.1021/nn304202gSearch in Google Scholar PubMed

[43] Vinh NQ, Minissale S, Vrielinck H, Gregorkiewicz T. Concentration of Er3+ ions contributing to 1.5-μm emission in Si/Si:Er nanolayers. Phys Rev B 2007;76:85339.10.1103/PhysRevB.76.085339Search in Google Scholar

[44] Liu GK, Zhuang HZ, Chen XY. Restricted phonon relaxation and anomalous thermalization of rare earth ions in nanocrystals. Nano Lett 2002;2:535–9.10.1021/nl0255303Search in Google Scholar

[45] Estes MJ, Moddel G. Luminescence from amorphous silicon nanostructures. Phys Rev B 1996;54:14633–42.10.1103/PhysRevB.54.14633Search in Google Scholar PubMed

[46] Thillosen N, Sebald K, Hardtdegen H, et al. The state of strain in single GaN nanocolumns as derived from micro-photoluminescence measurements. Nano Lett 2006;6:704–8.10.1021/nl052456qSearch in Google Scholar PubMed

[47] Thelander C, Agarwal P, Brongersma S, et al. Nanowire- based one-dimensional electronics. Mater Today 2006;9: 28–35.10.1016/S1369-7021(06)71651-0Search in Google Scholar

[48] Li R, Yerci S, Dal Negro L. Temperature dependence of the energy transfer from amorphous silicon nitride to Er ions. Appl Phys Lett 2009;95:041111.10.1063/1.3186062Search in Google Scholar

[49] Yiyang G, Maria M, Selçuk Y, et al. Linewidth narrowing and Purcell enhancement in photonic crystal cavities on an Er-doped silicon nitride platform. Opt Express 2010;18:2601–12.10.1364/OE.18.002601Search in Google Scholar PubMed


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2019-0535).


Received: 2019-12-19
Revised: 2020-02-28
Accepted: 2020-03-05
Published Online: 2020-04-21

©2020 Spyros Gallis et al., published by De Gruyter, Berlin/Boston

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

Downloaded on 1.6.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2019-0535/html
Scroll to top button