Rapid #: -16480465

CROSS REF ID: 1104701 LENDER: ZCU :: Electronic BORROWER: PUL :: Interlibrary Services, Firestone

TYPE: Article CC:CCG JOURNAL TITLE: Advanced optical materials USER JOURNAL TITLE: Advanced Optical Materials ARTICLE TITLE: Light Emission from Self-Assembled and Laser-Crystallized Chalcogenide Metasurface ARTICLE AUTHOR: Feifan Wang, Zi Wang, Dun Mao, Mingkun Chen, Qiu L VOLUME: 8 ISSUE: 8 MONTH: April YEAR: 2020 PAGES: NA ISSN: 2195-1071 OCLC #:

Processed by RapidX: 8/17/2020 9:02:16 AM

This material may be protected by copyright law (Title 17 U.S. Code) Full Paper

www.advopticalmat.de Light Emission from Self-Assembled and Laser-Crystallized Chalcogenide Metasurface

Feifan Wang, Zi Wang, Dun Mao, Mingkun Chen, Qiu Li, Thomas Kananen, Dustin Fang, Anishkumar Soman, Xiaoyong Hu, Craig B. Arnold,* and Tingyi Gu*

demonstrated to improve the light emis- Subwavelength periodic confinement can collectively and selectively sion efficiency through enhancing light enhance local light intensity and enable control over the photoinduced phase outcoupling efficiency[1,2] and sponta- [3–5] transformations at the nanometer scale. Standard nanofabrication process neous emission rate. Nanophotonic engineering can lead to narrowband can result in geometrical and compositional inhomogeneities in optical phase and directive light emission.[6–8] Chal- change materials, especially chalcogenides, as those materials exhibit poor cogenide materials with unique phase chemical and thermal stability. Here the self-assembled planar chalcogenide change properties have been explored nanostructured array is demonstrated with resonance-enhanced light for tunable thermal emission[9] or reflec- [10] emission to create an all-dielectric optical metasurface, by taking advantage tion in infrared wavelength ranges. of the fluid properties associated with solution-processed films. A patterned PhC waveguide has been demonstrated on thin-film chalcogenide by e-beam silicon membrane serves as a template for shaping the chalcogenide lithography, showing slow light and res- metasurface structure. Solution-processed arsenic sulfide metasurface onance-enhanced parametric nonlinear structures are self-assembled in the suspended 250 nm silicon membrane process in material.[11,12] To reduce cost templates. The periodic nanostructure dramatically manifests the local light– and improve scalability, solution-processed matter interaction such as absorption of incident photons, Raman emission, chalcogenide nanostructure is recently reported through solution process and and photoluminescence. Also, the thermal distribution is modified by the self-assembling.[13] In this work, we dem- boundaries and thus the photothermal crystallization process, leading to the onstrate the first active chalcogenide formation of anisotropic nanoemitters within the enhancement area. This metasurface fabricated by self-assembling hybrid structure shows wavelength-selective anisotropic photoluminescence, with tunable dimensions. The nanopho- which is a characteristic behavior of the collective response of the resonant- tonic structure is tailored to enhance the guided modes in a periodic nanostructure. The resonance-enhanced Purcell light emission efficiency through reso- nance-enhanced absorption of excitation, effect can manifest the quantum efficiency of localized light emission. suppressing guided mode at emission wavelength and Purcell enhancement.[14–17] The 2D hexagonal chalcogenide nanorod 1. Introduction arrays are self-assembled[13] on a silicon template through solu- tion processing.[18–24] Silicon nanophotonic structure with a few Periodic perturbation of refractive index in subwavelength nanometer surface roughness enables delamination of top and dimensions can boost efficiency of light emission device. Many bottom bulk chalcogenides during solvent evaporation leaving nanophotonic structures, such as photonic crystals (PhCs), only a filled nanorod array. Nanophotonic confinement can plasmonic structures, gratings, and resonators, have been simultaneously manifest local photon density for enhanced

F. Wang, Z. Wang, D. Mao, M. Chen, Dr. Q. Li, T. Kananen, D. Fang, Dr. Q. Li A. Soman, Dr. T. Gu Tianjin Key Laboratory of High Speed Cutting and Precision Machining Electrical and Computer Engineering School of Mechanical Engineering University of Delaware Tianjin University of Technology and Education Newark, DE 19716, USA Tianjin 300222, P. R. China E-mail: [email protected] Dr. C. B. Arnold F. Wang, Dr. X. Hu Princeton Institute for the Science and Technology of Materials State Key Laboratory for Mesoscopic & Department Princeton University of Physics Collaborative Innovation Center of Quantum Matter Princeton, NJ 08544, USA Peking University E-mail: [email protected] Beijing 100871, P. R. China The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/adom.201901236. DOI: 10.1002/adom.201901236

Adv. Optical Mater. 2020, 8, 1901236 1901236 (1 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

Figure 1. Self-assembled chalcogenide metasurface with the solution process. a) Fabrication process: i) Suspended planar silicon nanostructure; ii) drop-casting As2S3 solution; iii) reflow and delamination; and iv) top illuminated optical excitation. b) Scanning electron microscope of delaminated As2S3–silicon interface after step (iii). The silicon nanostructured template has hole radii from 80 to 150 nm, with 10 nm per step. Scale bars: 1 µm. c) Atomic microscope image of the self-assembled surface topology. resonance absorption[25] and nonlinear process.[26] The guided on a 250 nm thick silicon-on-insulator device layer via ultravi- mode with a concentration in the light-emitting materials can olet lithography and etching for reduced disorder scattering[30] significantly enhance the quantum efficiency of both stimulated (Figure S1, Supporting Information). The lattice constant of Raman emission and spontaneous photoluminescence (PL) the periodic structure is 415 nm, and we vary the hole radii process, which plays a major role in the device compared to the from 80 to 150 nm with 10 nm per step (Figure 1b). The sup- leaky mode enhancement of light outcoupling efficiency.[27–29] porting silicon oxide buffer layer is removed via buffered oxide wet etching. As2S3 solution is prepared by uniformly dissolving the powder into n-propylamine solvent according to recipes [14,18] 2. Localized Photon and Heat Distribution in the prior literature. The solution is then drop-casted in the Hybrid Metasurface onto the suspended silicon membrane, forming a thick As2S3 film on top of the Si layer. The sample is kept in the oxygen-free Fabrication of the light-emitting device is composed of two glove box environment for several days at room temperature to steps: 1) assembling the As2S3 nanostructures in the silicon evaporate out excess solvent. As2S3 dissolved in alkyl amines template; and 2) local laser annealing for enhanced PL. The can be described as a nanocolloidal solution consisting of flat first step is for shaping the chalcogenide nanostructures and clusters that internally retain the structure of the layer-like the second step removes the solvent-related atomic defects and starting material but capped by ionic pairs of sulfide dangling induces phase transitions in chalcogenide for effective light bonds and alkyl ammonium molecules.[31] Without annealing, emission. those solvent-related atomic defect states suppress light emis- [14] sion from As2S3. Due to the fluid nature of the drop-casted materials, the 2.1. Self-Assembled Chalcogenide Metasurface solution will accumulate in the hollow region[13] with excess chalcogenide material above and below the silicon membrane. Here we demonstrate a simple and scalable way of producing The sample is left in the glove box at room temperature over- chalcogenide nanostructures, combining advanced silicon night. Upon evaporating the excess solvent, the dried bulk nanofabrication and solution processing. Figure 1 shows the chalcogenide cleanly delaminates from the smooth, flat silicon fabrication process and scanning electron microscope (SEM) surface (Figure S2, Supporting Information). Our previous images of a self-assembled chalcogenide–silicon metasurface work shows that the room-temperature dried chalcogenide structures. The silicon patterns (triangular lattice) are fabricated nanorods still contain solvent-related residue (CH/NH

Adv. Optical Mater. 2020, 8, 1901236 1901236 (2 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

Figure 2. Device geometry. a) Side and b) top views of self-assembled chalcogenide metasurface in a silicon matrix. The yellow region indicates the intensity distribution of the top incident Gaussian beam profile, which is a continuous-wave light centered at 532 nm. c) Top view of the temperature distribution, with its z-axis marked in panel (d). d) Temperature distribution on the cross section, with its y-axis marked in panel (e). e) Temperature distribution near the bottom plane of the hybrid metasurface, with its z-axis marked in panel (d). f) Photoluminescence of As2S3 and laser-crystallized As4S4.

bond),[32] which can be removed by hot plate annealing (170 °C) mode profile leads to localized heating in plane (Figure S4, or laser annealing.[14] Energy-dispersive X-ray spectroscopy Supporting Information) and out of the plane (Figure S5, (EDX) mapping is used to verify the chalcogenide arrays in the Supporting Information). The cross-sectional view shows that silicon structure (Figure S3, Supporting Information), and little the highest temperature is found on the interface between the residual on silicon top surface. The results show the forma- metasurface and bulk chalcogenide substrate (Figure 2d). The tion of uniform As2S3 triangular arrays without solvent-specific constructive interference between incident and reflected light composition. With limited residual from the solvent, the sur- from metasurface–bulk substrate chalcogenide interface can face roughness of the delaminated bare silicon region is limited result in the highest optical field on the interface. The area with to be a few nanometers, as verified by atomic force microscope highest photon density has the highest heat generation rate. imaging (Figure 1c). Beyond phase transition temperature, extra influence from the crystallization might slightly modify the temperature distribu- tion. The top view of the hot spot distribution on the interface 2.2. Localized Laser Annealing between the metasurface and bulk substrate is shown in the Figure 2e. The local temperature can be up to 285 °C, which As shown in the cross-sectional image in Figure 2a, isolated is beyond the glass transition temperature of 191.7 °C.[14] The As2S3 nanorod arrays are self-assembled in the plane of the local photothermal heating removes the amine content and suspended silicon template. Laser annealing is then achieved reconstructs the atomic structure from As2S3 to As4S4. The by focusing a 532 nm continuous-wave laser onto the planar atomic state is stabilized after 10 s laser exposure (Figure S6, chalcogenide nanostructure. The laser spot size is about Supporting Information). As4S4 exhibits higher PL intensity in ≈1 µm. The intensity profile is superimposed onto the SEM the visible band, which likely comes from the loosely bound [14] image in Figure 2b. An example of the photothermal heat Wannier–Mott exciton in the atomic network of As4S4. distribution in the middle of nanostructured plane (marked as Figure 2f compares the PL spectra of bulk chalcogenide before a dashed blue line in Figure 2d) is shown in Figure 2c at a hole and after laser crystallization. Enhancement factors up to 35 radius of 140 nm, simulated by the combination of the discon- (gray dotted curve) are observed near 760 nm. The PL spec- tinuous Galerkin time-domain algorithm (for Maxwell equa- trum gives a major PL peak at 703 nm (1.76 eV) with 98 nm tion) and finite-difference method (for heat transport equation), bandwidth and a minor peak at 745 nm (1.66 eV) with 30 nm assuming an environmental temperature of 300 K.[33] As the bandwidth. PL spectra under different wavelength excitation excitation laser power is set at 40 µW µm−2, the photothermal are distinguished from each other, showing the multicom- effect heats the local area to 170 °C. The laser Gaussian profile ponent nature of the laser-crystallized As4S4 (Figure S7, Sup- is dramatically modified by the nanostructure. The modified porting Information).

Adv. Optical Mater. 2020, 8, 1901236 1901236 (3 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

Figure 3. Enhancement of photon density of states at the excitation wavelength. a) Raman spectrum of chalcogenide metasurface with hole radii of 70, 120, and 130 nm (accumulative vertical offset of 0.5 has been included in the spectra for clarity). b) The intensity of Raman peak of chalcogenide (near 345 cm−1) versus the hole radius. Blue dots with error bar are experimental results, and the gray empty circles linked by the dashed line show theoretical prediction by FDTD calculations, incorporating the mode coupling effect in the chalcogenide region. Inset: Top view (left) and side view (right) of the for an excitation wavelength of 532 nm coupling into the metasurface plane, for the hole radius of 140 nm. The chalcogenide region is highlighted by the dashed curves. c) The electric field on the top surface of metasurface for hole radii of 90, 140, and 170 nm.

3. Metasurface-Enhanced Light Emission cross-sectional (right) views of the photon distribution in the hybrid nanostructure at R = 140 nm, simulated by the finite-difference The nanophotonic structure tailors both Raman emission time-domain method (FDTD). The total enhancement factor of and PL. The Raman emission has wavelengths close to exci- an exciting photon is defined as the integral of the local elec- tation, narrowbandwidth, and is weakly influenced by defect tric field enhancement (FEx(λ0, r)) over the chalcogenide region: states accelerated nonradiative recombination, and thus used 2 ∫ |(FrEx λ,)|dr , where λ0 is the wavelength of the excitation for characterizing the metasurface enhancement of excita- As23S tion light. PL is broadband response, and its intensity is laser and r is a vector of position. The field enhancement factor strongly influenced by those nonradiative recombination. 2 2 is calculated by the integral ∫ |(FrEx λλ0 ,)||× FrEm-Raman (,)| dr, The PL spectra depend on the superposition of resonance- As23S enhanced mode at excitation and emission wavelengths. By where FEm-Raman is the Raman emission enhancement factor. The comparing the metasurface enhancement of both emissions, enhancement of the Raman emission wavelength is similar to we can obtain insights about the light emission enhancement the excitation, as their wavelength is closely spaced. The trend mechanisms. of simulated Raman emission intensity versus hole dimension (gray circles linked by the dashed line in Figure 3b) aligns well with the experimental data (blue dots with error bars in the 3.1. Raman Emission and Excitation Enhancement same figure).

The frequency offset of the Raman emission to the excita- tion laser wavelength fingerprints the unique phonon mode 3.2. Photoluminescence from the Nanostructured of the target material. Characterization of the Raman emis- Chalcogenide Emitter sion illustrates the mechanism of the enhancement of excita- tion light through coupling into the guided mode with a mode 3.2.1. Modeling Metasurface-Modified PL Spectrum profile that overlaps well with the As2S3 region. Local excita- tion photon density enhancement in As2S3 is examined by Different from narrowband ultrafast Raman emission micro-Raman spectrometry. Figure 3a shows the normalized (femtosecond scale lifetime), PL is a nonparametric and broad- Raman spectra of chalcogenide metasurface with increasing band quantum process (nanosecond scale lifetime).[37,38] The hole radii of 70, 120, and 130 nm (with 0.5 offset for clarity). emission bandwidth usually is across a few photonic guided/ In the hybrid nanostructure, the Raman peak near 305 cm−1 leaky modes. Selective enhancement of emission only occurs is the transverse acoustic mode in silicon. The Raman peaks at at resonance-enhanced wavelengths. The resonant mode can 345 cm−1 represent asymmetric stretching vibrational modes of effectively enhance the local density of states (LDOS), ρ, and [34–36] the AsS3/2 pyramids. The relative Raman peak intensity thus, the spontaneous emission rate for radiative recombi- ratio of the AsS bond and silicon increases with hole radius nation. The local spontaneous emission rate is defined as (R). The enhancement of the 532 nm photon excitation initially 2π   2 Γ=r 〈〈dE⋅〉〉 ρ, where d is the electric dipole moment, E is increases with R, and then decreases as the R becomes larger than 140 nm (R/a = 0.337). At R = 140 nm, the incident photon the local electric field, and ρ is the density of the electromag- has the highest overlap factor/enhancement factor with the netic modes or LDOS.[21] LDOS ρ(λ,r) can be obtained through chalcogenide area (Figure 3b,c), and thus, the highest Raman 3D FDTD simulation. The internal quantum efficiency of the peak intensity. The insets of Figure 2b show the top (left) and PL is η = Γr/(Γr + Γnr), where Γr and Γnr are the radiative and

Adv. Optical Mater. 2020, 8, 1901236 1901236 (4 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

nonradiative recombination rates. The carrier recombination 3.2.2. Contributions from Field Enhancement Effect and Light time (1/Γnr) of the chalcogenide semiconductor nanomaterial is Outcoupling Efficiency usually at the picosecond scale, which is three orders of magni- tude smaller than the radiation recombination lifetime (1/Γr). In the hybrid metasurface, it is important to differentiate Because the quantum efficiency of the amorphous material is and understand individual contributions from the local field usually small (Γr ≪ Γnr), η (≈Γr/Γnr) linearly increases with Γr. enhancement effect (FEexFEem) and extraction efficiency (EE). The local enhancement factor (FEm(λ,r)) of the electric field for Here we compare the experimentally measured PL from the the emitted photon is ρ/ρ0. ρ0 is the LDOS in bulk. The overall metasurface with increasing hole radius and FDTD calcula- emission intensity from the arsenic sulfide emitter is given by tions of outcoupling efficiency and field enhancement factors Equation (1), assuming the intensity of light is well below the (Figure 4). The measured PL spectra of seven devices with con- saturation level secutively increasing hole radius are plotted in Figure 4a (solid red line), compared to the line shape of PL from bulk (same Iphc 2 2 as the one in Figure 2f). The PL intensity is normalized to the =×CF(,λλrF)(,)rr× EE(,λ )dr (1) ∫ Ex 0 Em Raman signal of the As S As at 325 cm−1, to ensure that the Ibulk As44S   normalized PL is proportional to the quantum efficiency and where C is a constant describing the dipole alignment to the excludes effect from sample uniformity and changing filling guided modes (0 < C < 1); FEm and FEx are the enhancement factors, etc. Comparing the PL from devices with increasing factor of the local electric field for emitted (from 600 to 800 nm) hole radii, we observe that the PL efficiency increases with and pump/excitation (532 nm) light, respectively; λ and λ0 are hole radius, and the resonance-enhanced modes (highlighted the wavelengths of emission and pump light. EE(λ,r) is the by dashed gray curves) are blueshifted. The experimentally optical power extraction efficiency of the local dipole emission, measured values align well with the trend of field enhancement depending on the quality factor matching of intrinsic and out- simulated by FDTD simulation. Figure 4b plots the radius- coupling quality factors of the mode and wavelength detunings dependent outcoupling efficiency within 15° to the Gamma to its resonant wavelength.[2] point, which is correspondent to the band diagrams (Figure S8,

Figure 4. Collective PL near ⇩Γ point. a) PL spectra of the metasurface emitters with the hole radius from 80 to 140 nm in the step of 10 nm. The gray dashed curve marks the shift of modes. The PL spectra from nanoemitter embedded in metasurface (red solid curves) are compared to the line shape from the thin film (gray solid curves). b) FDTD simulation of the optical power extraction efficiency of local dipole emission embedded in planar metasurface with the hole radius as in panel (a). The trend of mode resonance shift with radius is highlighted in black dashed lines for guiding eyes. 2 2 c) The product of enhancement factors of the local electric field in nanoemitter: ∫ |(FrEx λλ0,)||× FrEm(,)| dr . The trend of geometry-dependent As44S modes is highlighted by dashed white lines.

Adv. Optical Mater. 2020, 8, 1901236 1901236 (5 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

Figure 5. Photoluminescence from the laser-crystallized nanoemitter. a) Photoluminescence spectrum of bulk chalcogenide (blue dashed curve), and metasurface before (blue solid curve) and after (black solid curve) laser irritation. Excitation laser of 532 nm and 4 mW is applied for probing the photoluminescence spectrums. b) Angle-dependent Raman and photoluminescence emission intensities. The relative intensity to normal incidence for the Raman peak of chalcogenide (blue cross), silicon (red star), and photoluminescence at three peaks (at wavelengths of 610, 650, and 725 nm). The solid line is the fitting of Gaussian beam profile with the angular spreading of 14.9°.

Supporting Information). As the dipole is confined in the chal- thin chalcogenide film on silicon is examined with exact laser cogenide region, the increased hole radius shifts those reso- crystallization and measurement procedures as for metasurface nance-enhanced modes to longer wavelengths. The trend of the samples. Only a 5% variation of the PL intensity is observed field enhancement factors is opposite, as plotted in Figure 4c, within 8° of sample rotation in tilting/polar angle. which incorporates the mode overlap among excitation, emis- sion, and As4S4 region (Figure S9, Supporting Information). Thus, we conclude that the PL enhancement is dominated by 4. Conclusion the contribution from local field enhancement. Among all the PL measurements, the maximum enhance- In conclusion, we present analytical and experimental investi- ment is achieved at a hole radius of 140 nm at the wavelength gations of self-assembled and laser-crystallized chalcogenide of 670 nm, where both excitation and emission resonance nanoemitters embedded within silicon template. Near-bandgap modes are excited, along with good mode overlap with the illumination is used to initiate localized laser crystallization As4S4 region. We compare the FDTD-calculated spectra and (high power) and to probe the Raman and PL (low power). We experimental measurement in Figure 5a. The blue solid curve first quantify the enhancement of photon density at excitation shows the experimental measurement of the chalcogenide wavelength through micro-Raman emission. A model incor- metasurface, and the black solid curve shows the FDTD-sim- porating excitation enhancement, Purcell factor, and extraction ulated spectrum. Major spectral features are resembled in the efficiency is compared to experimentally measured PL spectra simulation, including major peaks near 670, 700, 800 nm, and from the nanostructured As4S4 emitter. The relation between ripples around 600 nm. The bulk contribution from the chal- peak wavelength shift and the rod dimension indicates that the cogenide under the metasurface plane is marked via the blue PL enhancement is dominant by the Purcell effect. The overall dashed curve. Comparing the metasurface PL (solid blue curve) PL enhancement is attributed to the spatial overlap between the and bulk PL (blue dashed curve), a 4× enhancement is observed photonic-guided modes at excitation and emission wavelengths. with field enhancement in the nanostructure.

3.2.3. Directional PL 5. Experimental Section Device Fabrication: The silicon device was manufactured in a All the measurement and FDTD simulations are made near the complementary metal–oxide–semiconductor (CMOS) foundry. Deep- Gamma point (Figures 4 and 5a). Then we explore the polar ultraviolet lithography defines the template features on intrinsic silicon-on- [29] angle dependence of the PL. We measure the polar angle– insulator substrates, followed by reactive ion etching. The dimension of dependent Raman and PL spectra. The PL is also observed to the nanostructured area was 20 µm by 200 µm, which was much larger than the laser spot size. The finely grounded As2S3 dissolved in n-propylamine be highly directional. The Raman scattering of AsSAs varies for highly concentrated solutions (0.4 g mL−1). The high concentration little at an increasing tilting angle, with little influence from improves the viscosity and thus changes the flow properties of the solution. the nanophotonic structure. The red stars and blue crosses in The concentrated solution was then drop-casted onto a silicon template and Figure 5b represent Raman scatterings from As4S4 and silicon, dried, forming the self-assembled of As2S3 in the periodic voids. respectively. In contrast, the PL intensity rapidly decreases SEM images were taken with a Quanta 200 FEG environmental SEM with the incident angle, which drops to less than 10% at 8° at 5 keV in high vacuum mode. Electron diffraction X-ray equipped with SEM was then used for material composition analysis (Figure S3, tilting/polar angle. The empty gray squares, circles, and dots Supporting Information). in Figure 5b mark the relative intensity of the three peaks on Optical Measurements: The chalcogenide metasurface was the As4S4 PL spectra. The directional PL can be attributed to characterized by Raman spectra and PL. Raman spectra were collected metasurface nanostructure. As a control sample, a uniform by coupling the light scattered from the sample to an inVia Raman

Adv. Optical Mater. 2020, 8, 1901236 1901236 (6 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

spectrometer through a ×20 objective, with an ≈1 µm2 laser spot size [3] M. Fujta, S. Takahashi, Y. Tanaka, T. Asano, S. Noda, Science 2005, and 0.2 numerical aperture (Renishaw). The excitation wavelength was 308, 1296. set at 532 nm. [4] S. Dona, M. Fujita, T. Asano, Nat. Photonics 2007, 1, 449. Numerical Simulation: The simulation of excitation mode [5] A. Faraon, D. Englund, D. Bulla, B. Luther-Davies, B. J. Eggleton, enhancement was carried out by forwarding design. A plane wave of N. Stoltz, P. Petroff, J. Vucˇkovic´, Appl. Phys. Lett. 2008, 92, 532 nm light was normally incident onto the metasurface plane. The 043123. broadband emission mode profile was performed by inverse design. A [6] T. Asano, M. Suemitsu, K. Hashimoto, M. De Zoysa, T. Shibahara, dipole source covering the light emission bandwidth of chalcogenide T. Tsutsumi, S. Noda, Sci. Adv. 2016, 2, e1600499. was placed above the metasurface plane. The spectral response in [7] J. J. Greffet, R. Carminati, K. Joulain, J.-P. Mulet, S. Mainguy, chalcogenide region was collected and summarized at four random Y. Chen, Nature 2002, 416, 61. locations the source placed. The spatial resolution was set at 5 nm. The [8] H. Chalabi, A. Alù, M. L. Brongersma, Phys. Rev. B 2016, 94, material absorption and dispersion of the silicon and chalcogenide were separately measured and taken into consideration in simulation. 094307. The extraction efficiency of the local dipole emission and the field [9] Y. Qu, Q. Li, L. Cai, M. Pan, P. Ghosh, K. Du, M. Qiu, Light: Sci. enhancement factors in Equation (1) were simulated using the commercial Appl. 2018, 7, 26. software Lumerical FDTD solutions. To calculate the power extraction [10] M. Rudé, V. Mkhitaryan, A. E. Cetin, T. A. Miller, A. Carrilero, S. Wall, efficiency near the Gamma point, a dipole was set in the central hole of F. J. G. de Abajo, H. Altug, V. Pruneri, Adv. Opt. Mater. 2016, 4, the metasurface in the model. The power was collected by a point monitor 1060. right above the central hole. Then the power spectrum was normalized [11] B. J. Eggleton, B. Luther-Davies, K. Richardson, Nat. Photonics to the spectrum of the dipole source in the model. To calculate the field 2011, 5, 141. enhancement factors, first a Gaussian source of 532 nm was set above [12] B. Corcoran, C. Monat, C. Grillet, D. J. Moss, B. J. Eggleton, the central hole of metasurface as the excitation light, and a monitor was T. P. White, L. O’Faolain, T. F. Krauss, Nat. Photonics 2009, set to get the field profile inside the 3D photonic nanostructure. Then in 3, 206. another model, several Gaussian sources were set in the holes as emitters [13] T. D. Gupta, L. Martin-Monier, W. Yan, A. Le Bris, T. Nguyen-Dang, to get the field distribution profile of emission light. A. G. Page, K. T. Ho, F. Yesilköy, H. Altug, Y. Qu, F. Sorin, Nat. Nanotechnol. 2019, 14, 320. [14] T. Gu, H. Jeong, K. Yang, F. Wu, N. Yao, R. D. Priestley, C. E. White, Supporting Information C. B. Arnold, Appl. Phys. Lett. 2017, 110, 041904. [15] J. D. Musgraves, K. Richardson, H. Jain, Opt. Mater. Express 2011, Supporting Information is available from the Wiley Online Library or 1, 921. from the author. [16] A. Canciamilla, F. Morichetti, S. Grillanda, P. Velha, M. Sorel, V. Singh, A. Agarwal, L. C. Kimerling, A. Melloni, Opt. Express 2012, 20, 15807. [17] M. Wuttig, N. Yamada, Nat. Mater. 2007, 6, 824. Acknowledgements [18] Y. Zha, P. T. Lin, L. Kimerling, A. Agarwal, C. B. Arnold, ACS F.W., Z.W., and D.M. contributed equally to this work. The authors Photonics 2014, 1, 153. acknowledge experimental support from Y. Li and T. Heinz, and [19] L. Li, H. Lin, S. Qiao, Y. Zou, S. Danto, K. Richardson, discussion with A. Rodriguez. This was supported in part by the J. D. Musgraves, N. Lu, J. Hu, Nat. Photonics 2014, 8, 643. National Science Foundation through the MRSEC Center (Grant No. [20] N. Carlie, J. D. Musgraves, B. Zdyrko, I. Luzinov, J. Hu, V. Singh, DMR-1420541) and the AFOSR Young Investigator Program (FA9550-18- A. Agarwal, L. C. Kimerling, A. Canciamilla, F. Morichetti, 1-0300). Z.W. are partially supported by the NASA Early Career Faculty A. Melloni, Opt. Express 2010, 18, 26728. Award (80NSSC17K0526). F.W. and X.H. acknowledge support from the [21] A. Zakery, S. R. Elliott, J. Non-Cryst. Solids 2003, 330, 1. National Natural Science Foundation of China (Grant Nos. 61775003 [22] J. M. Laniel, N. Ho, R. Vallee, A. Villeneuve, J. Opt. Soc. Am. B 2005, and 11734001). Q.L. thanks the support from the National Natural 22, 437. Science Foundation of China (Grant No. 11772227). [23] Q. M. Ngo, K. Q. Le, V. D. Lam, J. Opt. Soc. Am. B 2012, 29, 1291. [24] Q. M. Ngo, K. Q. Le, V. D. Lam, J. Opt. Soc. Am. B 2014, 31, 1054. Conflict of Interest [25] T. Gu, A. Andryieuski, Y. Hao, Y. Li, J. Hone, C. W. Wong, The authors declare no conflict of interest. A. Lavrinenko, T. Low, T. F. Heinz, ACS Photonics 2015, 2, 1552. [26] A. Laucht, T. Günthner, S. Pütz, R. Saive, S. Frédérick, N. Hauke, M. Bichler, M.-C. Amann, A. W. Holleitner, M. Kaniber, J. J. Finley, Keywords J. Appl. Phys. 2012, 112, 093520. laser processing, metasurfaces, optical antennas, optical nanostructures, [27] S. Wu, S. Buckley, A. M. Jones, J. S. Ross, N. J. Ghimire, J. Yan, Raman emission D. G. Mandrus, W. Yao, F. Hatami, J. Vucˇkovic´, A. Majumdar, 2D Mater. 2014, 1, 011001. Received: July 22, 2019 [28] M. Boroditsky, R. Vrijen, T. F. Krauss, R. Coccioli, R. Bhat, Revised: December 13, 2019 E. Yablonovitch, J. Lightwave Technol. 1999, 17, 2096. Published online: February 17, 2020 [29] B. Zhen, S.-L. Chua, J. Lee, A. W. Rodriguez, X. Liang, S. G. Johnson, J. D. Joannopoulos, M. Soljacˇic´, O. Shapira, Proc. Natl. Acad. Sci. USA 2013, 110, 13711. [30] S. Kocaman, M. S. Aras, P. Hsieh, J. F. McMillan, C. G. Biris, [1] J. J. Wierer, A. David, M. M. Megens, Nat. Photonics 2009, 3, 163. N. C. Panoiu, M. B. Yu, D. L. Kwong, A. Stein, C. W. Wong, Nat. [2] S. Fan, P. R. Villeneuve, J. D. Joannopoulos, E. F. Schubert, Phys. Photonics 2011, 5, 499. Rev. Lett. 1997, 78, 3294. [31] N. S. Dutta, C. B. Arnold, RSC Adv. 2018, 8, 35819.

Adv. Optical Mater. 2020, 8, 1901236 1901236 (7 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advancedsciencenews.com www.advopticalmat.de

[32] T. Gu, J. Gao, E. E. Ostroumov, H. Jeong, F. Wu, R. Fardel, N. Yao, [35] D. Zhao, Ph.D. Theses, Lehigh University 2011. R. D. Priestley, G. D. Scholes, Y. L. Loo, C. B. Arnold, ACS Appl. [36] K. Richardson, T. Cardinal, M. Richardson, A. Schulte, S. Seal, Mater. Interfaces 2017, 9, 18911. in Photo-induced Metastability in Amorphous Semiconductors [33] https://apps.lumerical.com/sp_photothermal-heating-in-plasmo.html (Ed: A. V. Kolobov), Wiley-VCH,Weinheim, Germany 2003, Ch. 23. (accessed: July 2019). [37] A. Mahdavi, G. Sarau, J. Xavier, T. K. Paraïso, S. Christiansen, [34] A. Zakery, S. R. Elliott, Optical Nonlinearities in Chalcogenide Glasses F. Vollmer, Sci. Rep. 2016, 6, 25135. and Their Applications, Springer Series in Optical Sciences, Vol. 135, [38] L. Ondicˇ, M. Varga, I. Pelant, J. Valenta, A. Kromka, R. G. Elliman, Springer, Berlin, Germany 2007. Sci. Rep. 2017, 7, 5763.

Adv. Optical Mater. 2020, 8, 1901236 1901236 (8 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim