Second harmonic generation and broad-band photoluminescence in mesoporous Si/SiO2 nanoparticles
-
Viktoriia Mastalieva
, Vladimir Neplokh
, Andrey Machnev
, Dmitry Kurdyukov
Abstract
Efficient second harmonic generation and broad-band photoluminescence from deeply subwavelength and nontoxic nanoparticles is essential for nanophotonic applications. Here, we explore nonlinear optical response from mesoporous Si/SiO2, SiO2, and Si nanoparticles, considering various fabrication and treatment procedures. We show that thermal annealing (including femtosecond laser treatment) of mesoporous Si/SiO2 nanoparticles provides the transformation of Si phase from amorphous to crystalline, enhancing the second harmonic and nonlinear photoluminescent response. Notably, the SiO2 mesoporous frame of the considered Si/SiO2 nanoparticles plays a dual positive role for the nonlinear process: it stabilizes the Si material, and SiO2:OH− material has a second-order nonlinearity itself and impacts to the observed second harmonic signal.
1 Introduction
Second-order optical nonlinearity is a well-known mechanism for modulating and generating coherent light in photonic devices and bio-imaging applications. Due to strong photon confinement and long photon lifetime, integrated microresonators have emerged as a promising platform for the investigation of nonlinear optical effects [1], [2], [3], [4]. Optical microresonators provide an implementation of efficient nonlinear effects at the micro- and nanoscale. However, existing silicon-based structures (including SiN x and SiO x ) still require the development of fabrication methods and nanoscale design since bulk silicon has a near-zero second-order response due to its centrosymmetric structure; thus, the second harmonics generation (SHG) efficiency is governed by the properties of Si surface. A variety of novel material platforms possessing high χ (2) nonlinearity have been developed over the past two decades [5], [6], [7], [8]. These studies carry a high value for fundamental science [8], [9], [10], [11] as well classical and quantum applications [12], [13], quantum communication and computation [14], [15], [16], [17], [18], [19], [20], [21], long-haul communication systems [18], infrared light visualization, etc. [19], [20], [21]. Moreover, optical micro- and nanoresonators help to overcome the requirement of strong phase matching for efficient SHG generation in bulk materials by replacing them by a modest condition of quasi-modal overlap [22].
AlGaAs is considered a versatile platform for efficient SHG generation [23], [24]. Among semiconductor materials, AlGaAs has one of the highest values of χ (2) nonlinearity and refractive index. However, it is not a trivial task to fabricate AlGaAs-based nanoparticles or nanocylinders on optically transparent wafers. Expensive epitaxial techniques are normally employed, which also require an additional layer having a higher content of Al with further oxidation to AlO x to achieve optical separation (decoupling) of active (Al)GaAs layer (with lower Al content) and GaAs wafer.
At the same time, SiO x and Si-based nanoparticles can be synthesized using cost-efficient colloidal approaches with a control of mesoporosity and ability to fill the pores with the required materials. In the early 1990s, porous Si emerged as a suitable platform for photonics [25]. Since the porous Si structure can provide a very high specific surface area above 100 m2/g [26], [27], surface-related SHG in mesoporous Si has been extensively studied in the last decades [28], [29], [30]. On the other hand, Si microresonators such as nanowires and nanoparticles are also marked by prominent second-order nonlinear properties [18], [19], [30], [31]. The electromagnetic field matching of pump irradiation and the generated second harmonic (SH) in nanoscale resonators [19], [32]–[35] leads to the elevated conversion efficiency compared to thin films and allows tuning of the nonlinear system properties, including the directivity of SHG emission [36], [37], [38], [39]. In other words, the effective volume of nanoresonators can be multiplied by a quality factor of optical mode. Thus, deeply subwavelength Si nanostructures with a nanocrystalline internal structure and high surface area properly passivated should further improve the efficiency of SHG and other nonlinear processes.
Concerning Si-based SHG structures, the mesoporous spherical particles [19] appear to be the ideal material system. According to our previous report [24], for these nanoparticles: (i) literally all Si material is a surface, no Si material is thick enough to be considered as bulk; (ii) synthesis method can be adjusted to any desired particle size in the range of few tens to several thousands of nanometers, while preserving the quality of mesoporous structure; (iii) diameter of pores is tunable in the range of 2–20 nm; thus, (iv) particles can be filled with beneficial material that further enhance SHG efficiency or related properties (refractive index, light absorption, control of surface electron states etc.); (v) synthesis method based on organic micelles provides a construction set for homogeneous arrays of Si particles with selected morphology, so they can be arranged in a 2D or 3D periodic grid (photonic crystal or synthetic opals [40], [41]) having linear optical properties differ from single nanoparticles or their chains and metasurfaces [42], [43], which can be useful for nonlinear response tailoring.
In this work, we for the first time propose mesoporous nanoparticles with a SiO2 frame filled with a Si phase, previously developed for drug delivery [26], for efficient SHG owing to high surface to volume ratio. By means of mapping in a IR-femtosecond (fs) laser scanning system, the SHG-to-pump characteristics were compared among mesoporous Si/SiO2, SiO2, and Si nanoparticles, considering various fabrication and treatment procedures. Thermal or laser-induced Si annealing leads to transformation of the Si phase from amorphous to nanocrystalline one, which improves the nonlinear performance of the studied nanoparticles. Moreover, the thermally treated Si/SiO2 nanoparticles also exhibit a broad-band photoluminescence under linear and fs-laser excitation, which can be employed for visualization in optics at the nanoscale.
2 Materials and methods
2.1 Synthesis of nanoparticles
Monodisperse spherical SiO2 nanoparticles (reference SiO 2 NPs) with low porosity were produced through the hydrolysis of tetraethoxysilane (TEOS) in an alcohol-based solution containing ammonia and water [44], [45]. The primary substance concentration in the aqueous ammonia solution was 24 wt%. The initial alcohol concentration was 95.7 wt%. Both the alcohol and ammonia solutions were used as received. The synthesis process utilized deionized water with a specific resistance of 10 MΩ cm.
TEOS underwent fractional distillation with a boiling temperature of 166–168 °С [45], [46], [47]. This fraction was then treated with a 0.5 wt% aqueous ammonia solution for 20 min. The ratio of TEOS to the aqueous ammonia solution was 5:1 by mass (the procedure is detailed in Ref. [46]). The synthesis process had a duration of 4 h and resulted in a formation of 550 ± 20 nm monodisperse low-porosity (less than 15 %) nanoparticles having a high concentration of OH− groups [26], [48]–[50]. The synthesized particles were centrifuged and annealed at 900 °C for 5 h.
Spherical mesoporous Si/SiO2 nanoparticles (meso Si/SiO 2 NPs) were synthesized in several steps (see the workflow scheme in Figure 1). Monodisperse spherical mesoporous silica nanoparticles (meso SiO 2 NPs) were synthesized according to our previously developed method [51] by hydrolysis of TEOS in a water–ethanol–ammonia medium containing cylindrical micelles of cetyltrimethylammonium bromide as a pore-forming agent. The silica particles obtained were centrifuged, dried in air at 70 °C for 24 h, and annealed at 550 °C for 5 h to remove micelles and form the pores. The specific surface area and pore volume of the particles synthesized were 810 m2/g and 0.55 cm3/g (about 50 % of particle volume), respectively, and the sectional diameter of pores was about 3 nm [51].

Fabrication scheme for mesoporous spherical SiO2 nanoparticles, hybrid Si/SiO2 NPs, and derived mesoporous Si nanoparticles.
Silicon phase was synthesized within the pores of silica particles by the thermal decomposition of monosilane [26]. The synthesis was carried out in the reactor consisting of a fused silica tube equipped with an external resistive heater. A gas mixture of SiH4 (5 %) and Ar was passed through the tube at a flow rate of 1 sccm over a 1 g portion of the silica particles powder placed in a fused silica crucible. The thickness of the particle layer did not exceed 1 mm. The reactor pressure was 1,000 Torr. A constant temperature of 440 °C was maintained within the reactor, while the synthesis duration was 40 h. As a result of the thermal decomposition of silane, silicon filled the pores of meso SiO2 particles producing meso Si/SiO2 NPs. Note, the employed procedure provided the synthesis amorphous silicon phase filling the nanopores [26]. Additionally, to obtain meso Si/SiO2 NPs with nanocrystalline silicon (nc-Si) phase, the produced NPs were thermally annealed in an ampoule at a temperature of 850 °C (annealed meso Si/SiO 2 NPs) [26].
The meso SiO2 NPs for the Si/SiO2 NP formation were found to have a diameter of 450 ± 30 nm [52]. The meso Si/SiO2 NP morphology was studied in [52], inherited from the micelle structure, the pores are clusters of 7 cylinders, hexagonally packed, with a diameter of ∼3 nm and a length of ∼10 nm (see Figure 1 schematics). The meso SiO2 NPs according to the porosimetry measurements had a porosity of 50 vol% [51]. This porosity allowed almost complete filling, so the synthesized meso Si/SiO2 NPs had about equal volume proportion of Si and SiO2 phases, and the remaining unfilled space inside the NPs was determined by comparing the sedimentation rates of unfilled and filled particles [50].
To obtain mesoporous Si nanoparticles (meso Si NPs), the SiO2 material was selectively etched out from the meso Si/SiO2 composite. For that purpose, the meso Si/SiO2 NPs were placed in a 0.015 M aqueous solution of hydrofluoric acid (HF). Etching was carried out for 10 h. The resulting Si particles were repeatedly washed with deionized water (10 MΩ cm) and sedimented by centrifugation.
2.2 Material characterization
Transmission electron microscopy (TEM) study was carried out to analyze the structural properties of the synthesized nanoparticles at the nanometer scale. The study was performed using JEM 2100F (Tokyo, Japan) high-resolution transmission electron microscopy (HRTEM) setup operating at 200 kV accelerating voltage.
Raman spectroscopy was employed to identify Si crystalline properties in meso Si/SiO2 NPs [53]. Raman spectra were acquired at ambient conditions using a Horiba Jobin Yvon T64000 (Japan) spectrometer, utilizing the second harmonic of an Nd:YAG laser (532 nm wavelength) as a light source. The spectrometer was equipped with a confocal microscope, allowing laser beam focusing into a spot of approximately 1.5 μm in diameter. The excitation density on the sample surface during Raman measurements did not exceed 1 kW/cm2. The spectrometer was calibrated using a reference Si (111) wafer.
Infrared (IR) transmission spectra were obtained at ambient conditions using a vacuum Bruker Vertex 80v (Germany) Fourier spectrometer operating in rapid-scan mode and equipped with a globar as a source of light, KBr beam splitter and pyroelectric photodetector with a CsI window. Spectral resolution was about 4 cm−1. Samples were mounted on a copper holder with an aperture of about 4 mm. To obtain the transmission spectra of the meso Si/SiO2 NPs, the intensity spectra of light passing through the Si/SiO2 NPs on the substrate were normalized to reference spectra of Si substrate.
2.3 Optical response and second harmonic measurements
For nonlinear optical measurements, the samples from an isopropanol colloid were applied onto precleaned tempered quartz holders to form several NP layers.
Nonlinear optical investigations were conducted using a confocal laser-scanning microscope (LSM) setup (LSM-980, Zeiss, Germany). The LSM external acousto-optic modulator (AOM) port was utilized to deliver femtosecond (fs) laser pulses (Discovery-NX, Coherent, USA) with the following parameters: (i) a repetition rate of 80 MHz, (ii) a duration of approximately 150 fs, and (iii) linear polarization. The tunable wavelength range for the pulses’ central wavelength of 800–1,020 nm (spectral full width at half maximum, FWHM < 10 nm) was limited by LSM optics. Si detector (S120C, PM100USB, Thorlabs, USA) was used to set the power value of laser irradiation applied to the sample, and it was calibrated for AOM levels of LSM. The maximum power of radiation can reach up to ∼25 mW (0.3 nJ per single pulse).
For the optical study, air microscopy objectives with numerical apertures of 0.3 and 0.8 (corresponding to magnifications of 10× and 20×, respectively) were employed. LSM measurements produced 16-bit images using a sensitive GaAsP photomultiplier tube detector, enabled by the galvo mirror system of LSM at an average speed of tens of mm/s (e.g., giving about 20 µs for a map pixel with a size of approximately 500 nm for the 20× objective). The nonlinear optical measurements were performed in a reflection mode, i.e., the focusing objective was also used to collect the optical response from the studied samples. The integrated SH signal was collected within a range of ±10 nm relative to the central wavelength of the harmonic for each pixel of the LSM image.
The spectral characteristics of the nonlinear response were obtained in the so-called “scanning Lambda LSM” mode with an optimal for the studied samples pump wavelength in the range of 900–980 nm, which allowed hyperspectral mapping in the visible range with a spectral resolution of 3 nm. For meso Si/SiO2 NPs, the spectral characteristics were obtained in the 425–700 nm range, demonstrating not only the SHG signal but also a broadband photoluminescence (PL) response.
3 Results and discussion
The performed scanning electron microscopy (SEM) imaging demonstrated no distinguishable difference between as-synthesized and annealed meso Si/SiO2 NPs (see Figure S1 in the Supporting Information), while TEM imaging for Si/SiO2 NPs before and after annealing revealed an enlarging of Si clusters that we attribute to improved crystallinity due to annealing (Figure 2a and b). Indeed, the obtained TEM images allow distinguishing two material phases; the nc-Si material is manifested by a bright contrast that is more pronounced after annealing.

TEM images of meso Si/SiO2 NPs before (a) and after (b) thermal annealing at 850 °C. (c) Raman spectra of meso Si/SiO2 NPs before and after thermal annealing at 850 °C. For the convenience, the spectrum of as-synthesized particles is amplified by ×4 factor. (d) Transmittance spectrum of meso Si/SiO2 NPs sample after annealing (green line). Blue line corresponds to the spectrum of reference c-Si (100) wafer.
Importantly, the SiO2 material in meso Si/SiO2 NPs had also a high OH− group concentration (similar to reference SiO2 NPs). The Si material, however, appeared to be amorphous for the as-synthesized NPs [26], which was confirmed by Raman measurements (Figure 2c). Indeed, the representative Raman scattering spectrum for the particles before annealing did not demonstrate any clear line typical for the crystal phase [26], [54], while the observed weak Raman signal with a broad spectral band is typical for amorphous material [55]. After annealing of Si/SiO2 NPs, the acquired Raman spectra confirmed the presence of crystalline Si phase manifested by a prominent narrow line at 518.5 cm−1 corresponding to the transverse optical phonon mode of crystalline Si (Figure 2c) [56]. The 1.5 cm−1 shift relative to the bulk c-Si 520 cm−1 corresponds to the 5–10 nm size of nc-Si inclusions in accordance with our previously reported results [26] that also correlates to the data reported in the literature [57]. In Ref. [26], the volume fraction of the crystalline Si component relative to the amorphous component reaches 70 %.
IR transmission spectroscopy of meso Si/SiO2 NP samples provided essential information about the quality of the Si/SiO2 interface [58]. The absorption line at 1,080 cm−1 is characteristic for Si–O bonds (Figure 2d), and its presence confirms a high-quality interface in meso Si/SiO2 NPs strongly affecting SHG properties [53], [59], [60], [61]. Also peaks at 810 and 456 cm−1 could assign the symmetric stretching and rocking modes of the Si–O–Si vibrations in SiO2 phase [61]. For a comparison, Figure 2d shows the corresponding IR transmittance spectrum for a reference c-Si (100) substrate, confirming the absence of Si–O bonds.
According to the described synthesis procedure, the meso Si NPs were obtained from the meso Si/SiO2 NPs by SiO2 etching; thus, they keep the same diameter (450 ± 30 nm in our case). As shown in Ref. [26], after SiO2 removal, the Si material kept an amorphous structure and configuration of initial pore clusters. After annealing in the inert atmosphere of the sealed ampoules, the Si material transformed into nanocrystalline phase; however, the meso Si NP colloid oxidizes in the atmosphere in a matter of days. The dried meso Si NPs (e.g., transferred to quartz holders for the nonlinear measurements) acquire natural oxide surface layer in half an hour, that in case of mesoporous structure take a significant portion of the material volume (i.e., 0.5 nm thick shell for the 3.1 nm Si veins counts for about 54 % of their volume). Thus, these samples were characterized during 24 h after preparation, transported as a colloid, and placed to quartz holders just before SHG measurements. It should be noted that meso Si NPs turn to pure SiO2 in several weeks while being stored in an isopropanol colloid.
To summarize the material characterization results, we expected the strongest SHG signal from the annealed meso Si NP samples due to their structure being a practically pure nc-Si surface that is known to demonstrate the highest SHG efficiency among silicon and silica materials (without doping) [62]. A comparable SHG signal was expected for the annealed meso Si/SiO2 NP samples, because they also contain mesoporous nc-Si phase, but the SHG efficiency was expected to be second to the meso Si NP samples due to the questionable quality of the Si/SiO2 interface and parasitic light absorption in SiO2 material both reducing the SHG efficiency. The nonannealed as-synthesized meso Si/SiO2 NP and meso SiO2 NP samples were expected to be the least efficient due to the absence of nc-Si phase and modest SHG efficiency of SiO2:OH− material, which was reported in literature [63], [64], [65]. However, the SHG experiments revealed a picture that was divergent from these expectations.
To study nonlinear optical phenomena, 4 types of samples were considered: (i) meso Si NPs, (ii) SiO2 NPs, (iii) meso Si/SiO2 NPs with amorphous silicon in the pores, and (iv) annealed meso Si/SiO2 NPs with nc-Si in the pores. The pumping laser wavelengths were in the range of (910–960) nm, which corresponds to efficient light localization in the considered NPs (see Section S2 for details).
The systematic study of meso Si NPs, both nonannealed and annealed, recently synthesized and aged, revealed a very weak SHG signal. We consider it might have been due to the structure damages during the chemical removal of SiO2 material from the blank meso Si/SiO2 NPs or the deteriorating effect of the inevitable formation of natural oxide layer under the SHG measurement conditions. In fact, pure meso Si NPs demonstrated negligible SHG signal in comparison with the reference SiO2:OH− NPs. Apparently, the later demonstrated fairly distinguishable SH with an intensity-to-pump power ratio in a log-log scale (commonly referred in the literature as slope) close to the characteristic value of 2 (Figure 3a); the spectral characteristic at 960 nm IR fs excitation pulses is dominated by a narrow line at 480 nm (Figure 4b). Apart from the line attributed to the SH wavelength, the optical response for reference SiO2:OH− NPs had no other distinguishable features in the studied region of 400–550 nm that indicates an absence of PL or other parasitic luminescence. A typical measured SHG map is presented in Figure 4a; the resolution allows distinguishing the individual SiO2:OH− NPs. The inhomogeneous SHG intensity from different NPs is due to the different focusing since the NPs were not lying in the same plane.

SHG-to-pump curves for (a) reference SiO2 NPs, (b) thermally annealed, and (c) as-synthesized meso Si/SiO2 NPs. For (b) and (c), the blue curves correspond to initial measurements, and green curves correspond to the repeated measurements. The colored areas in (c) present as follows: blue area – normal SHG response at relatively low pump power, green area – an increased slope on SHG response, pink area – irreversible sample damage at high pump power.

Optical response from (a, b) single reference SiO2 NPs under fs-laser excitation of 960 nm wavelength and (c, d) single meso Si/SiO2 NPs annealed under 910 nm wavelength fs laser with high power. (a) LSM map of reference SiO2 NP layers, integrated in (470–485) nm spectral range, corresponding to SHG signal, (c) LSM map of meso Si/SiO2 NP layers, integrated in (500–680) nm spectral range, corresponding to PL signal. (b, d) Corresponding spectral characteristics. Circles on LSM maps denote the positions of spectra acquisition.
The slope of 2.3 for reference SiO2:OH− NPs (Figure 3a) is in full agreement with the data reported in the literature [65], [66], while the strong signal indicates a remarkably efficient SHG process that will be thoroughly studied elsewhere. It should be noted that the SiO2:OH− NPs did not demonstrate significant changes of SH intensity due to the fs-laser annealing, i.e., the OH− group-related SHG in SiO2 NPs is thermally stable.
The most efficient SHG was observed for the meso Si/SiO2 NPs. The thermally annealed meso Si/SiO2 NPs demonstrated a standard SH slope value of 2, and SH intensity-pump curves are reproducible at consequent measurements (Figure 3b). Interestingly, the as-synthesized (nonannealed) meso Si/SiO2 NPs demonstrated hysteresis in initial measurements of SH intensity-pump curves (Figure 3c), i.e., the firstly acquired curve was different from the consequent. We consider it is due to the in situ annealing of the meso Si/SiO2 NPs (with amorphous Si phase) under the probing IR laser pulses leading to the formation of nc-Si phase and possible changes of the Si/SiO2 interface. Indeed, the as-synthesized meso Si/SiO2 NPs demonstrated at the first measurement a significantly lower SHG signal compared to the thermally annealed meso NPs, while the slope was close to 2 before the threshold pump power value of about 3 mW (the corresponding optical power range is marked by a blue color in Figure 3c). After the threshold, the SH intensity started to improve rapidly that gave a superquadratic slope value of about 4.1, which can be associated with local annealing under fs laser irradiation (this power range is marked by green color in Figure 3c). It should be noted, that this effect is similar to the superquadratic hysteresis behavior previously reported for the laser pump-induced modifications of Si nanowire array [18]. SHG reached the maximum efficiency in the range of 10–12 mW pump power that is limited by the material degradation due to the overheating (the overheat power range is marked in Figure 3c by pink color). The measurements, repeated after the first acquisition, demonstrated a reproducible pump power dependency with a near-quadratic slope. Remarkably, the SH intensity value corresponding to the in situ annealed meso Si/SiO2 NPs is close to the performance of the meso Si/SiO2 NPs annealed in a sealed ampule.
Following the experimental observation of the SHG signal dependency on the pump power, the mesoporous Si/SiO2 nanoparticles undergo an ablation and the crystalline fraction of Si in the system starts depending on the input power. For moderately low intensities, the SH signal grows quadratically on the pump as the ablation process remains negligible. For higher intensities, however, the fraction of Si grows owing to ablation and, as a result, the effective polarizability of the nonlinear particle begins to depend on the input power, leading to 4th power dependence. Similar effects of power-dependent polarizability were observed in Ref. [67], where linear scattering spectra were investigated.
We proposed a model for ablation-dependent SHG, based on a single dipole. The dipole moment is given by: p L = α dip E 0, while we suggest that the dipole polarizability (due to the recrystallization process) depends on the pumping intensity α dip = α 0 + α 1 I 0. Therefore, it gives subquadratic dependence at high excitation intensity (see Section S2 in the Supporting Information for details).
Since the energy of excitation photons (corresponding to 910–960 nm or 1.29–1.36 eV) is low compared to the energy bandgap of amorphous Si (around 1.7–1.8 eV [68], depending on methods of deposition), we suggest that the laser annealing can be caused by a multiphoton absorption process. Thus, at high fs laser intensity, multiphoton processes can give a significant contribution, which also explains the subquadratic behavior of SHG on laser intensity. At the same time, when the crystalline Si phase appears (energy band gap of 1.1 eV [69] and higher, depending on the size of crystallites), linear absorption can also take place.
Finally, we acquired spectral characteristics from individual meso Si/SiO2 NPs measured after annealing under the fs laser irradiation of 910 nm wavelength (Figure 4c and d). One can see that optical response demonstrates not only SHG signal (at 455 nm) but also a broad PL response in the range of (500–700) nm. This response becomes more prominent with an increasing laser irradiation power (Figure 4d), demonstrating the annealing of NPs and formation of nc-Si phase. The emergence of visible range PL indicates the presence of nanocrystalline silicon in Si/SiO2 structure within the pores of the mesoporous particle [70]. Moreover, the spectral positions of PL maxima indicate the average size of Si nanocrystals that is 3.1 nm in our case [53], [71], [72], [73]. This value is less compared to data acquired from Raman measurements for the thermally annealed meso Si/SiO2 NPs (Figure 2c), which indicates the difference in annealing regimes. Additional dark-field and linear PL measurements presented in Sections S3 and 4 in the Supporting Information also confirm the recrystallization phenomenon of silicon material in meso Si/SiO2 NPs under fs-laser illumination.
Theoretical calculations for light-induced heating in Si/SiO2 nanospheres, reported in Ref. [74], show that higher temperature can be achieved under laser irradiation in nanospheres consisting of amorphous Si (α-Si) compared to ones made of crystalline Si (c-Si). Due to the higher thermal conductivity of c-Si (compared to α-Si), better heat dissipation is observed in c-Si-based nanostructures in comparison with α-Si ones. Moreover, the enhanced optical absorption governed by optical resonances of nanostructures resulted in the increase of the local temperature compared to films or nonresonant nanostructures [75]. Thus, the local temperature in α-Si/SiO2 nanospheres under laser irradiation can reach values comparable with a melting point even at moderate laser intensity [74].
The reported theoretical results support our observation that the as-synthesized meso Si/SiO2 NPs, consisting of Si in the amorphous phase, can be efficiently heated by a laser. This heating promotes the Si phase transition from amorphous to crystalline. Moreover, this transition can occur locally within Si-based nanostructures, without its significant reshaping [75].
Thus, we consider that thermal annealing in the ampoule affects the crystallization of silicon in the pores of meso Si/SiO2 NPs almost in the same manner as the in situ laser-induced annealing, but the crystallization process using an LSM setup is controlled by knowing the reference value of the power, at which destruction of the material occurs. In the case of annealing under laser pulses, we control not only the power used to change the crystallization, but we can also understand how much silicon in the pore has undergone a change in its phase, in contrast to the annealing in an ampoule, when it is impossible to determine a priori the exact temperature and annealing time required for crystallization of the material in the pores.
4 Conclusions
In this work, SHG in mesoporous Si-based nanoparticles has been studied, where the enhancement of SHG in mesoporous Si/SiO2 nanoparticles has been demonstrated. The SHG efficiency has been found to be increased as the silicon in nanopores is annealed under femtosecond near-infrared pulses and transforms into the nanocrystalline phase. Remarkably, the SiO2 mesoporous frame of Si/SiO2 nanoparticles plays a dual positive role for the SHG process: first, it stabilizes the Si material; second, SiO2:OH− material has a second-order nonlinearity itself and impacts to the observed SH signal. Moreover, Si/SiO2 nanoparticles emit light in a broad spectral range under fs-laser excitation, which can be employed for visualization purposes.
The proposed mesoporous Si/SiO2 nanoparticles can be considered as a very promising structure for SHG studies and applications at the nanoscale, since they combine most of the known approaches to facilitate SHG in silicon-based material (most notably: microcavity, nanocrystal phase, high specific surface area), while the fabrication method is scalable and cost-effective.
Funding source: FRCCP RAS
Award Identifier / Grant number: 506694
Funding source: Russian Science Foundation
Award Identifier / Grant number: 23-79-00018
Funding source: the Ministry of Science and Higher Education of the Russian Federation
Award Identifier / Grant number: FSEG-2024-0017
-
Research funding: This work was supported by the Russian Science Foundation (project No. 23-79-00018, Investigation of nonlinear optical properties of Si and hybrid SiO2/Si nanostructures for IR visualizers). Femtosecond laser measurements were performed using core research facilities of the N.N. Semenov Federal Research Center for Chemical Physics (FRCCP RAS, No. 506694). AM and IM acknowledge the financial support by the Ministry of Science and Higher Education of the Russian Federation FSEG-2024-0017 for theoretical support and numerical modeling.
-
Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission. Mastalieva Viktoriia – experimental data, writing original draft, data curation, visualization; Vladimir Neplokh – writing – writing original draft, writing – review & editing, methodology; Arseniy Aybush – resources, data curation; Ekaterina Stovpiaga – resources; Daniil Eurov – resources; Maksim Vinnichenko – resources, writing original draft; Danila Karaulov – formal analysis; Demid Kirillenko – data curation, formal analysis; Alexey Mozharov – validation, data analysis, formation of a theoretical model; Vladislav Sharov – experimental data; Denis Kolchanov – experimental data; Andrey Machnev – experimental data; Valery Golubev – writing original draft, writing – review & editing; Alexander Smirnov – visualization; Pavel Ginzburg – writing original draft, writing – review & editing, formation of a theoretical model; Sergey Makarov – writing original draft, writing – review & editing, project administration; Dmitry Kurdyukov – writing original draft, writing – review & editing; Ivan Mukhin – writing original draft, writing – review & editing, project administration.
-
Conflict of interest: Authors state no conflicts of interest.
-
Informed consent: Informed consent was obtained from all individuals included in this study.
-
Ethical approval: The conducted research is not related to either human or animal use.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.
References
[1] D. Smirnova and Y. S. Kivshar, “Multipolar nonlinear nanophotonics,” Optica, vol. 3, no. 11, pp. 1241–1255, 2016. https://doi.org/10.1364/optica.3.001241.Search in Google Scholar
[2] S. Kruk and Y. Kivshar, “Functional meta-optics and nanophotonics governed by Mie resonances,” ACS Photonics, vol. 4, no. 11, pp. 2638–2649, 2017. https://doi.org/10.1021/acsphotonics.7b01038.Search in Google Scholar
[3] B. Sain, C. Meier, and T. Zentgraf, “Nonlinear optics in all-dielectric nanoantennas and metasurfaces: a review,” Adv. Photonics, vol. 1, no. 2, p. 024002, 2019. https://doi.org/10.1117/1.ap.1.2.024002.Search in Google Scholar
[4] K. Koshelev and Y. Kivshar, “Dielectric resonant metaphotonics,” ACS Photonics, vol. 8, no. 1, pp. 102–112, 2020. https://doi.org/10.1021/acsphotonics.0c01315.Search in Google Scholar
[5] Y. R. Shen, “Surface properties probed by second-harmonic and sum-frequency generation,” Nature, vol. 337, no. 6207, pp. 519–525, 1989. https://doi.org/10.1038/337519a0.Search in Google Scholar
[6] J. He, et al.., “Nonlinear nanophotonic devices in the ultraviolet to visible wavelength range,” Nanophotonics, vol. 9, no. 12, pp. 3781–3804, 2020. https://doi.org/10.1515/nanoph-2020-0231.Search in Google Scholar
[7] T. P. McKenna, et al.., “Ultra-low-power second-order nonlinear optics on a chip,” Nat. Commun., vol. 13, no. 1, p. 4532, 2022. https://doi.org/10.1038/s41467-022-31134-5.Search in Google Scholar PubMed PubMed Central
[8] N. Bloembergen, R. K. Chang, S. S. Jha, and C. H. Lee, “Optical second-harmonic generation in reflection from media with inversion symmetry,” Phys. Rev., vol. 174, no. 3, p. 813, 1968. https://doi.org/10.1103/physrev.174.813.Search in Google Scholar
[9] H. W. K. Tom, T. F. Heinz, and Y. R. Shen, “Second-harmonic reflection from silicon surfaces and its relation to structural symmetry,” Phys. Rev. Lett., vol. 51, no. 21, 1983. https://doi.org/10.1103/physrevlett.51.1983.Search in Google Scholar
[10] J. E. Sipe, D. J. Moss, and H. M. Van Driel, “Phenomenological theory of optical second-and third-harmonic generation from cubic centrosymmetric crystals,” Phys. Rev. B, vol. 35, no. 3, p. 1129, 1987. https://doi.org/10.1103/physrevb.35.1129.Search in Google Scholar PubMed
[11] W. L. Mochán, J. A. Maytorena, B. S. Mendoza, and V. L. Brudny, “Second-harmonic generation in arrays of spherical particles,” Phys. Rev. B, vol. 68, no. 8, p. 085318, 2003. https://doi.org/10.1103/physrevb.68.085318.Search in Google Scholar
[12] D. Bouwmeester, J. W. Pan, K. Mattle, M. Eibl, H. Weinfurter, and A. Zeilinger, “Experimental quantum teleportation,” Nature, vol. 390, no. 6660, pp. 575–579, 1997. https://doi.org/10.1038/37539.Search in Google Scholar
[13] K. L. Vodopyanov, et al.., “Terahertz-wave generation in quasi-phase-matched GaAs,” Appl. Phys. Lett., vol. 89, no. 14, 2006. https://doi.org/10.1063/1.2357551.Search in Google Scholar
[14] F. Bussières, et al.., “Quantum teleportation from a telecom-wavelength photon to a solid-state quantum memory,” Nat. Photonics, vol. 8, no. 10, pp. 775–778, 2014. https://doi.org/10.1038/nphoton.2014.215.Search in Google Scholar
[15] M. Bock, et al.., “High-fidelity entanglement between a trapped ion and a telecom photon via quantum frequency conversion,” Nat. Commun., vol. 9, no. 1, p. 1998, 2018. https://doi.org/10.1038/s41467-018-04341-2.Search in Google Scholar PubMed PubMed Central
[16] L. Tang, J. Tang, M. Chen, F. Nori, M. Xiao, and K. Xia, “Quantum squeezing induced optical nonreciprocity,” Phys. Rev. Lett., vol. 128, no. 8, p. 083604, 2022. https://doi.org/10.1103/physrevlett.128.083604.Search in Google Scholar
[17] C. Wang, et al.., “Integrated lithium niobate electro-optic modulators operating at CMOS-compatible voltages,” Nature, vol. 562, no. 7725, pp. 101–104, 2018. https://doi.org/10.1038/s41586-018-0551-y.Search in Google Scholar PubMed
[18] V. Mastalieva, et al.., “Laser-activated second harmonic generation in flexible membrane with Si nanowires,” Nanomaterials, vol. 13, no. 9, p. 1563, 2023. https://doi.org/10.3390/nano13091563.Search in Google Scholar PubMed PubMed Central
[19] V. V. Fedorov, et al.., “Gallium phosphide nanowires in a free-standing, flexible, and semitransparent membrane for large-scale infrared-to-visible light conversion,” ACS Nano, vol. 14, no. 8, pp. 10624–10632, 2020. https://doi.org/10.1021/acsnano.0c04872.Search in Google Scholar PubMed
[20] X. Hu, et al.., “Infrared-light visualization by organic upconversion devices,” ACS Appl. Electron. Mater., vol. 5, no. 10, pp. 5378–5385, 2023. https://doi.org/10.1021/acsaelm.3c00943.Search in Google Scholar
[21] O. Bisi, S. Ossicini, and L. Pavesi, “Porous silicon: a quantum sponge structure for silicon based optoelectronics,” Surf. Sci. Rep., vol. 38, nos. 1–3, pp. 1–126, 2000. https://doi.org/10.1016/s0167-5729(99)00012-6.Search in Google Scholar
[22] P. Ginzburg, et al.., “Nonlinearly coupled localized plasmon resonances: resonant second-harmonic generation,” Phys. Rev. B Condens. Matter Mater. Phys., vol. 86, no. 8, p. 085422, 2012. https://doi.org/10.1103/physrevb.86.085422.Search in Google Scholar
[23] L. Carletti, A. Locatelli, O. Stepanenko, G. Leo, and C. De Angelis, “Enhanced second-harmonic generation from magnetic resonance in AlGaAs nanoantennas,” Opt. Express, vol. 23, no. 20, pp. 26544–26550, 2015. https://doi.org/10.1364/oe.23.026544.Search in Google Scholar PubMed
[24] T. Liu, X. Fang, and S. Xiao, “Tuning nonlinear second-harmonic generation in AlGaAs nanoantennas via chalcogenide phase-change material,” Phys. Rev. B, vol. 104, no. 19, p. 195428, 2021. https://doi.org/10.1103/physrevb.104.195428.Search in Google Scholar
[25] J. Wang, et al.., “Surface-engineered mesoporous silicon microparticles as high-Coulombic-efficiency anodes for lithium-ion batteries,” Nano Energy, vol. 61, pp. 404–410, 2019, https://doi.org/10.1016/j.nanoen.2019.04.070.Search in Google Scholar
[26] D. A. Kurdyukov, et al.., “Fabrication of doxorubicin-loaded monodisperse spherical micro-mesoporous silicon particles for enhanced inhibition of cancer cell proliferation,” Microporous Mesoporous Mater., vol. 281, pp. 1–8, 2019, https://doi.org/10.1016/j.micromeso.2019.02.029.Search in Google Scholar
[27] M. Cazzanelli and J. Schilling, “Second order optical nonlinearity in silicon by symmetry breaking,” Appl. Phys. Rev., vol. 3, no. 1, 2016. https://doi.org/10.1063/1.4941558.Search in Google Scholar
[28] F. Priolo, T. Gregorkiewicz, M. Galli, and T. F. Krauss, “Silicon nanostructures for photonics and photovoltaics,” Nat. Nanotechnol., vol. 9, no. 1, pp. 19–32, 2014. https://doi.org/10.1038/nnano.2013.271.Search in Google Scholar PubMed
[29] M. Ha, et al.., “Multicomponent plasmonic nanoparticles: from heterostructured nanoparticles to colloidal composite nanostructures,” Chem. Rev., vol. 119, no. 24, pp. 12208–12278, 2019. https://doi.org/10.1021/acs.chemrev.9b00234.Search in Google Scholar PubMed
[30] J. F. Rhoads, S. W. Shaw, and K. L. Turner, “Nonlinear dynamics and its applications in micro-and nanoresonators,” J. Dyn. Syst. Meas. Control, vol. 132, p. 034001, 2010. https://doi.org/10.1115/1.4001333.Search in Google Scholar
[31] G. Kozyreff, J. L. Dominguez-Juarez, and J. Martorell, “Nonlinear optics in spheres: from second harmonic scattering to quasi-phase matched generation in whispering gallery modes,” Laser Photonics Rev., vol. 5, no. 6, pp. 737–749, 2011. https://doi.org/10.1002/lpor.201000036.Search in Google Scholar
[32] S. V. Makarov, et al.., “Efficient second-harmonic generation in nanocrystalline silicon nanoparticles,” Nano Lett., vol. 17, no. 5, pp. 3047–3053, 2017. https://doi.org/10.1021/acs.nanolett.7b00392.Search in Google Scholar PubMed
[33] A. Nikolaeva, et al.., “Large-scale flexible membrane with resonant silicon nanowires for infrared visualization via efficient third harmonic generation,” Appl. Phys. Lett., vol. 120, no. 15, 2022, https://doi.org/10.1063/5.0088217.Search in Google Scholar
[34] D. G. Baranov, S. V. Makarov, A. E. Krasnok, P. A. Belov, and A. Alù, “Tuning of near- and far-field properties of all-dielectric dimer nanoantennas via ultrafast electron-hole plasma photoexcitation,” Laser Photonics Rev., vol. 10, no. 6, pp. 1009–1015, 2016. https://doi.org/10.1002/lpor.201600164.Search in Google Scholar
[35] D. Ryabov, O. Pashina, G. Zograf, S. Makarov, and M. Petrov, “Nonlinear optical heating of all-dielectric super-cavity: efficient light-to-heat conversion through giant thermorefractive bistability,” Nanophotonics, vol. 11, no. 17, pp. 3981–3991, 2022. https://doi.org/10.1515/nanoph-2022-0074.Search in Google Scholar
[36] A. Rudenko, A. Han, and J. V. Moloney, “Trade-off between second- and third-order nonlinearities, ultrafast free carrier absorption and material damage in silicon nanoparticles,” Adv. Opt. Mater., vol. 11, no. 2, p. 2201654, 2023. https://doi.org/10.1002/adom.202201654.Search in Google Scholar
[37] S. S. Kruk, et al.., “Nonlinear optical magnetism revealed by second-harmonic generation in nanoantennas,” Nano Lett., vol. 17, no. 6, pp. 3914–3918, 2017. https://doi.org/10.1021/acs.nanolett.7b01488.Search in Google Scholar PubMed
[38] L. Carletti, A. Locatelli, D. Neshev, and C. De Angelis, “Shaping the radiation pattern of second-harmonic generation from AlGaAs dielectric nanoantennas,” ACS Photonics, vol. 3, no. 8, pp. 1500–1507, 2016. https://doi.org/10.1021/acsphotonics.6b00050.Search in Google Scholar
[39] M. Scalora, et al.., “Pulsed second-harmonic generation in nonlinear, one-dimensional, periodic structures,” Phys. Rev. A, vol. 56, no. 4, p. 3166, 1997. https://doi.org/10.1103/physreva.56.3166.Search in Google Scholar
[40] A. B. Pevtsov, A. N. Poddubny, S. A. Yakovlev, D. A. Kurdyukov, and V. G. Golubev, “Light control in Ge2Sb2Te5-coated opaline photonic crystals mediated by interplay of Wood anomalies and 3D Bragg diffraction,” J. Appl. Phys., vol. 113, no. 14, 2013, https://doi.org/10.1063/1.4799384.Search in Google Scholar
[41] L. Pallavidino, et al.., “Synthesis of amorphous silicon/magnesia based direct opals with tunable optical properties,” Opt. Mater., vol. 33, no. 3, pp. 563–569, 2011. https://doi.org/10.1016/j.optmat.2010.11.003.Search in Google Scholar
[42] S. Kruk, et al.., “Nonlinear light generation in topological nanostructures,” Nat. Nanotechnol., vol. 14, no. 2, pp. 126–130, 2019. https://doi.org/10.1038/s41565-018-0324-7.Search in Google Scholar PubMed
[43] S. Kruk, et al.., “Edge states and topological phase transitions in chains of dielectric nanoparticles,” Small, vol. 13, no. 11, p. 1603190, 2017. https://doi.org/10.1002/smll.201603190.Search in Google Scholar PubMed
[44] W. Stober, A. Fink, and E. Bohn, “Controlled growth of monodisperse silica spheres in the micron size range,” J. Colloid Interface Sci., vol. 26, no. 1, p. 62, 1968. https://doi.org/10.1016/0021-9797(68)90272-5.Search in Google Scholar
[45] N. D. Deniskina, D. V. Kalinin, L. K. Kazantseva, and N. V. Sobolev, Blagorodnye opaly: prirodnye i sinteticheskie, USA, Nauka, Sibirskoe otd-nie, 1987.Search in Google Scholar
[46] E. A. Ryabenko, A. I. Kuznetsov, and B. Z. Shalumov, “O raspredelenii primesej mezhdu fazami pri glubokoj och-istke tetraetoksisilana rastvorom ammiaka (On the distri-bution of impurities between phases during deep purification of tetraethoxysilane with an ammonia solution),” Zh. Prikl. Khim., vol. 50, no. 7, pp. 1625–1627, 1977.Search in Google Scholar
[47] V. V. Potapov and D. V. Kamashev, “Synthesis of precious opal in a hydrothermal solution,” Glass Phys. Chem., vol. 32, pp. 89–98, 2006.10.1134/S1087659606010123Search in Google Scholar
[48] G. M. Gajiev, et al.., “Bragg reflection spectroscopy of opal-like photonic crystals,” Phys. Rev. B, vol. 72, no. 20, p. 205115, 2005. https://doi.org/10.1103/physrevb.72.205115.Search in Google Scholar
[49] D. A. Eurov, et al.., “Core–shell monodisperse spherical mSiO2/Gd2O3: Eu3+@ mSiO2 particles as potential multifunctional theranostic agents,” J. Nanopart. Res., vol. 17, pp. 1–10, 2015, https://doi.org/10.1007/s11051-015-2891-y.Search in Google Scholar
[50] E. Y. Trofimova, A. E. Aleksenskii, S. A. Grudinkin, I. V. Korkin, D. A. Kurdyukov, and V. G. Golubev, “Effect of tetraethoxysilane pretreatment on synthesis of colloidal particles of amorphous silicon dioxide,” Colloid J., vol. 73, pp. 546–550, 2011, https://doi.org/10.1134/s1061933x11040156.Search in Google Scholar
[51] E. Y. Trofimova, et al.., “Monodisperse spherical mesoporous silica particles: fast synthesis procedure and fabrication of photonic-crystal films,” Nanotechnology, vol. 24, no. 15, p. 155601, 2013. https://doi.org/10.1088/0957-4484/24/15/155601.Search in Google Scholar PubMed
[52] D. A. Kurdyukov, N. A. Feoktistov, D. A. Kirilenko, A. N. Smirnov, V. Y. Davydov, and V. G. E. Golubev, “Template synthesis of monodisperse submicrometer spherical nanoporous silicon particles,” Semiconductors, vol. 53, pp. 1048–1053, 2019. https://doi.org/10.1134/s106378261908013x.Search in Google Scholar
[53] G. Ledoux, et al.., “Photoluminescence properties of silicon nanocrystals as a function of their size,” Phys. Rev. B, vol. 62, no. 23, p. 15942, 2000. https://doi.org/10.1103/physrevb.62.15942.Search in Google Scholar
[54] İ. Doğan and M. C. M. van de Sanden, “Direct characterization of nanocrystal size distribution using Raman spectroscopy,” J. Appl. Phys., vol. 114, no. 13, p. 134310, 2013. https://doi.org/10.1063/1.4824178.Search in Google Scholar
[55] X. Zhang, Z. Liu, M. Tanaka, and T. Watanabe, “Formation mechanism of amorphous silicon nanoparticles with additional counter-flow quenching gas by induction thermal plasma,” Chem. Eng. Sci., vol. 230, p. 116217, 2021, https://doi.org/10.1016/j.ces.2020.116217.Search in Google Scholar
[56] Y. He, C. Yin, G. Cheng, L. Wang, X. Liu, and G. Y. Hu, “The structure and properties of nanosize crystalline silicon films,” J. Appl. Phys., vol. 75, no. 2, pp. 797–803, 1994. https://doi.org/10.1063/1.356432.Search in Google Scholar
[57] S. V. Makarov, et al.., “Nanoscale generation of white light for ultrabroadband nanospectroscopy,” Nano Lett., vol. 18, no. 1, pp. 535–539, 2018. https://doi.org/10.1021/acs.nanolett.7b04542.Search in Google Scholar PubMed
[58] A. G. Cullis, L. T. Canham, and P. D. J. Calcott, “The structural and luminescence properties of porous silicon,” J. Appl. Phys., vol. 82, no. 3, pp. 909–965, 1997. https://doi.org/10.1063/1.366536.Search in Google Scholar
[59] F. Huisken, et al.., “Silicon carbide nanoparticles produced by CO2 laser pyrolysis of SiH4/C2H2 gas mixtures in a flow reactor,” J. Nanopart. Res., vol. 1, pp. 293–303, 1999.Search in Google Scholar
[60] R. M. Almeida and C. G. Pantano, “Structural investigation of silica gel films by infrared spectroscopy,” J. Appl. Phys., vol. 68, no. 8, pp. 4225–4232, 1990. https://doi.org/10.1063/1.346213.Search in Google Scholar
[61] J. Wang, B. Zou, and M. A. El-Sayed, “Comparison between the polarized Fourier-transform infrared spectra of aged porous silicon and amorphous silicon dioxide films on Si (100) surface,” J. Mol. Struct., vol. 508, nos. 1–3, pp. 87–96, 1999. https://doi.org/10.1016/s0022-2860(99)00003-4.Search in Google Scholar
[62] J. D. Barnes, Investigations into Size and Surface Control of Silicon Nanocrystals for Improved Optical Properties, Doctoral dissertation, Portland State University, 2021.Search in Google Scholar
[63] M. Taheri and N. Mansour, “Silicon nanoparticles produced by two-step nanosecond pulsed laser ablation in ethanol for enhanced blue emission properties,” Silicon, vol. 12, no. 4, pp. 789–797, 2020. https://doi.org/10.1007/s12633-019-00168-8.Search in Google Scholar
[64] H. Aguiar, J. Serra, P. González, and B. León, “Structural study of sol–gel silicate glasses by IR and Raman spectroscopies,” J. Non-Cryst. Solids, vol. 355, no. 8, pp. 475–480, 2009. https://doi.org/10.1016/j.jnoncrysol.2009.01.010.Search in Google Scholar
[65] A. S. Kanaeva, A. Y. Limanskaya, A. A. Karmanov, I. A. Pronin, and I. A. Averin, “Issledovanie osobennostej adsorbcii OH-grupp na poverxnosti i v ob``eme poristy`x nanokompozicionny`x materialov na osnove SiO2-SnO2,” Nadezhnost i kachestvo, vol. 2, pp. 108–109, 2016.Search in Google Scholar
[66] V. Vandalon, “Second-harmonic generation: spectroscopic phase and intensity study of the Si(100) interface with SiO2 and Al2O3 thin films,” MSc. thesis, Eindhoven University of Technology, The Netherlands, 2011. Available at: https://pure.tue.nl/ws/files/46958768/772814-1.pdf.Search in Google Scholar
[67] U. Zywietz, A. B. Evlyukhin, C. Reinhardt, and B. N. Chichkov, “Laser printing of silicon nanoparticles with resonant optical electric and magnetic responses,” Nat. Commun., vol. 5, no. 1, pp. 1–7, 2014. https://doi.org/10.1038/ncomms4402.Search in Google Scholar PubMed
[68] D. T. Pierce and W. E. Spicer, “Electronic structure of amorphous Si from photoemission and optical studies,” Phys. Rev. B, vol. 5, no. 8, p. 3017, 1972. https://doi.org/10.1103/physrevb.5.3017.Search in Google Scholar
[69] M. A. Green, “Self-consistent optical parameters of intrinsic silicon at 300 K including temperature coefficients,” Sol. Energy Mater. Sol. Cells, vol. 92, no. 11, pp. 1305–1310, 2008. https://doi.org/10.1016/j.solmat.2008.06.009.Search in Google Scholar
[70] S. Tong, X. N. Liu, and X. M. Bao, “Study of photoluminescence in nanocrystalline silicon/amorphous silicon multilayers,” Appl. Phys. Lett., vol. 66, no. 4, pp. 469–471, 1995. https://doi.org/10.1063/1.114059.Search in Google Scholar
[71] C. M. Hessel, et al.., “Synthesis of ligand-stabilized silicon nanocrystals with size-dependent photoluminescence spanning visible to near-infrared wavelengths,” Chem. Mater., vol. 24, no. 2, pp. 393–401, 2012. https://doi.org/10.1021/cm2032866.Search in Google Scholar
[72] M. Zhong-Yuan, et al.., “The origin of blue photoluminescence from nc-Si/SiO2 multilayers,” Chin. Phys. B, vol. 17, no. 1, p. 303, 2008. https://doi.org/10.1088/1674-1056/17/1/053.Search in Google Scholar
[73] Z. Ma, et al.., “The evolution investigation of photoluminescence from a-Si: H/SiO2 to nc-Si/SiO2 multilayers,” J. Appl. Phys., vol. 95, no. 5, pp. 2448–2451, 2004. https://doi.org/10.1063/1.1646443.Search in Google Scholar
[74] S. Danesi, et al.., “Photo-induced heat generation in non-plasmonic nanoantennas,” Phys. Chem. Chem. Phys., vol. 20, no. 22, pp. 15307–15315, 2018. https://doi.org/10.1039/c8cp01919c.Search in Google Scholar PubMed
[75] S. Danesi and I. Alessandri, “Using optical resonances to control heat generation and propagation in silicon nanostructures,” Phys. Chem. Chem. Phys., vol. 21, no. 22, pp. 11724–11730, 2019. https://doi.org/10.1039/c8cp07573e.Search in Google Scholar PubMed
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2024-0218).
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Editorial
- New frontiers in nonlinear nanophotonics
- Reviews
- Tailoring of the polarization-resolved second harmonic generation in two-dimensional semiconductors
- A review of gallium phosphide nanophotonics towards omnipotent nonlinear devices
- Nonlinear photonics on integrated platforms
- Nonlinear optical physics at terahertz frequency
- Research Articles
- Second harmonic generation and broad-band photoluminescence in mesoporous Si/SiO2 nanoparticles
- Second harmonic generation in monolithic gallium phosphide metasurfaces
- Intrinsic nonlinear geometric phase in SHG from zincblende crystal symmetry media
- CMOS-compatible, AlScN-based integrated electro-optic phase shifter
- Symmetry-breaking-induced off-resonance second-harmonic generation enhancement in asymmetric plasmonic nanoparticle dimers
- Nonreciprocal scattering and unidirectional cloaking in nonlinear nanoantennas
- Metallic photoluminescence of plasmonic nanoparticles in both weak and strong excitation regimes
- Inverse design of nonlinear metasurfaces for sum frequency generation
- Tunable third harmonic generation based on high-Q polarization-controlled hybrid phase-change metasurface
- Phase-matched third-harmonic generation in silicon nitride waveguides
- Nonlinear mid-infrared meta-membranes
- Phase-matched five-wave mixing in zinc oxide microwire
- Tunable high-order harmonic generation in GeSbTe nano-films
- Si metasurface supporting multiple quasi-BICs for degenerate four-wave mixing
- Cryogenic nonlinear microscopy of high-Q metasurfaces coupled with transition metal dichalcogenide monolayers
- Giant second-harmonic generation in monolayer MoS2 boosted by dual bound states in the continuum
- Quasi-BICs enhanced second harmonic generation from WSe2 monolayer
- Intense second-harmonic generation in two-dimensional PtSe2
- Efficient generation of octave-separating orbital angular momentum beams via forked grating array in lithium niobite crystal
- High-efficiency nonlinear frequency conversion enabled by optimizing the ferroelectric domain structure in x-cut LNOI ridge waveguide
- Shape unrestricted topological corner state based on Kekulé modulation and enhanced nonlinear harmonic generation
- Vortex solitons in topological disclination lattices
- Dirac exciton–polariton condensates in photonic crystal gratings
- Enhancing cooperativity of molecular J-aggregates by resonantly coupled dielectric metasurfaces
- Symmetry-protected bound states in the continuum on an integrated photonic platform
- Ultrashort pulse biphoton source in lithium niobate nanophotonics at 2 μm
- Entangled photon-pair generation in nonlinear thin-films
- Directionally tunable co- and counterpropagating photon pairs from a nonlinear metasurface
- All-optical modulator with photonic topological insulator made of metallic quantum wells
- Photo-thermo-optical modulation of Raman scattering from Mie-resonant silicon nanostructures
- Plasmonic electro-optic modulators on lead zirconate titanate platform
- Miniature spectrometer based on graded bandgap perovskite filter
- Far-field mapping and efficient beaming of second harmonic by a plasmonic metagrating
Articles in the same Issue
- Frontmatter
- Editorial
- New frontiers in nonlinear nanophotonics
- Reviews
- Tailoring of the polarization-resolved second harmonic generation in two-dimensional semiconductors
- A review of gallium phosphide nanophotonics towards omnipotent nonlinear devices
- Nonlinear photonics on integrated platforms
- Nonlinear optical physics at terahertz frequency
- Research Articles
- Second harmonic generation and broad-band photoluminescence in mesoporous Si/SiO2 nanoparticles
- Second harmonic generation in monolithic gallium phosphide metasurfaces
- Intrinsic nonlinear geometric phase in SHG from zincblende crystal symmetry media
- CMOS-compatible, AlScN-based integrated electro-optic phase shifter
- Symmetry-breaking-induced off-resonance second-harmonic generation enhancement in asymmetric plasmonic nanoparticle dimers
- Nonreciprocal scattering and unidirectional cloaking in nonlinear nanoantennas
- Metallic photoluminescence of plasmonic nanoparticles in both weak and strong excitation regimes
- Inverse design of nonlinear metasurfaces for sum frequency generation
- Tunable third harmonic generation based on high-Q polarization-controlled hybrid phase-change metasurface
- Phase-matched third-harmonic generation in silicon nitride waveguides
- Nonlinear mid-infrared meta-membranes
- Phase-matched five-wave mixing in zinc oxide microwire
- Tunable high-order harmonic generation in GeSbTe nano-films
- Si metasurface supporting multiple quasi-BICs for degenerate four-wave mixing
- Cryogenic nonlinear microscopy of high-Q metasurfaces coupled with transition metal dichalcogenide monolayers
- Giant second-harmonic generation in monolayer MoS2 boosted by dual bound states in the continuum
- Quasi-BICs enhanced second harmonic generation from WSe2 monolayer
- Intense second-harmonic generation in two-dimensional PtSe2
- Efficient generation of octave-separating orbital angular momentum beams via forked grating array in lithium niobite crystal
- High-efficiency nonlinear frequency conversion enabled by optimizing the ferroelectric domain structure in x-cut LNOI ridge waveguide
- Shape unrestricted topological corner state based on Kekulé modulation and enhanced nonlinear harmonic generation
- Vortex solitons in topological disclination lattices
- Dirac exciton–polariton condensates in photonic crystal gratings
- Enhancing cooperativity of molecular J-aggregates by resonantly coupled dielectric metasurfaces
- Symmetry-protected bound states in the continuum on an integrated photonic platform
- Ultrashort pulse biphoton source in lithium niobate nanophotonics at 2 μm
- Entangled photon-pair generation in nonlinear thin-films
- Directionally tunable co- and counterpropagating photon pairs from a nonlinear metasurface
- All-optical modulator with photonic topological insulator made of metallic quantum wells
- Photo-thermo-optical modulation of Raman scattering from Mie-resonant silicon nanostructures
- Plasmonic electro-optic modulators on lead zirconate titanate platform
- Miniature spectrometer based on graded bandgap perovskite filter
- Far-field mapping and efficient beaming of second harmonic by a plasmonic metagrating