Next Article in Journal
Shredded-Coconut-Derived Sulfur-Doped Hard Carbon via Hydrothermal Processing for High-Performance Sodium Ion Anodes
Previous Article in Journal
Electrical Behavior of Combinatorial Thin-Film ZrxTa1−xOy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Processing and Characterization of High-Density Fe-Silicide/Si Core–Shell Quantum Dots for Light Emission

by
Katsunori Makihara
1,2,*,
Yuji Yamamoto
1,2,
Markus Andreas Schubert
2,
Andreas Mai
2 and
Seiichi Miyazaki
3
1
Graduate School of Engineering, Nagoya University, Nagoya 464-8601, Japan
2
IHP-Leibniz-Institut für Innovative Mikroelektronik, 15236 Frankfurt, Germany
3
Hiroshima University, Higashihiroshima 739-0046, Japan
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(10), 733; https://doi.org/10.3390/nano15100733
Submission received: 14 April 2025 / Revised: 9 May 2025 / Accepted: 13 May 2025 / Published: 14 May 2025
(This article belongs to the Section Nanophotonics Materials and Devices)

Abstract

:
Si-based photonics has garnered considerable attention as a future device for complementary metal–oxide–semiconductor (CMOS) computing. However, few studies have investigated Si-based light sources highly compatible with Si ultra large-scale integration processing. In this study, we observed stable light emission at room temperature from superatom-like β–FeSi2–core/Si–shell quantum dots (QDs). The β–FeSi2–core/Si–shell QDs, with an areal density as high as ~1011 cm−2 were fabricated by self-aligned silicide process of Fe–silicide capped Si–QDs on ~3.0 nm SiO2/n–Si (100) substrates, followed by SiH4 exposure at 400 °C. From the room temperature photoluminescence characteristics, β–FeSi2 core/Si–shell QDs can be regarded as active elements in optical applications because they offer the advantages of photonic signal processing capabilities and can be combined with electronic logic control and data storage.

Graphical Abstract

1. Introduction

The monolithic integration of Si-based photonics with electronic processes on a single chip is considered a key advancement for beyond–complementary metal–oxide–semiconductor (CMOS) computing [1,2,3,4,5,6,7,8,9]. Photonic components utilize photons allowing for the highest level of precision and highly efficient data processing and transmission, resulting in faster and more energy-efficient devices. However, from an application point of view, Si-based photonics necessitate Si-based light sources that are highly compatible with Si–ultra-large-scale integration (Si–ULSI) processing [10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34]. Most importantly, despite the excellent optical properties of Si-based light emitters due to their high refractive index and the high-quality Si/SiO2 interface, which means low defect density, they suffer from the inherent indirect band gap of Si, resulting in long carrier lifetimes and inefficient light emission. Therefore, integration of Si-based photonic processing with electrical processing on a single chip remains a significant challenge. In this study, we report on light emission from Si quantum dots (QDs) with a β–FeSi2 core referred to as superatom-like QDs. β–FeSi2 is a semiconductor phase with an indirect bandgap of ~0.7 eV, close to the direct bandgap value (~0.85 eV) at room temperature and shows light emission in the near-infrared region [35,36,37,38,39,40,41,42,43]. Therefore, to achieve good thermal stability and light emission characteristics, β–FeSi2 is one of the promising candidates.
Si-based QDs have garnered significant attention due to their potential application as active elements in group IV photonics, driven by their light emission properties attributed to the carrier confinement effect. Previously, we reported the self-assembly of Si-QDs with an areal density as high as ~1011 cm−2 on ultrathin SiO2 by controlling the early stages of thermal decomposition of SiH4, showcasing their light emission properties [44,45,46]. More recently, we also detailed the formation of the β–FeSi2 nanodots (NDs) through the exposure of an ultrathin Fe layer to a remote H2 plasma (H2–RP) followed by SiH4 exposure, and characterized their photoluminescence (PL) properties [47,48,49]. In this work, we successfully formed high-density Si–shell/β–FeSi2–core QDs, akin to superatoms, with a uniform size distribution by controlling self-aligned silicidation of the Si–QDs and subsequent selective Si growth at low temperature to minimize metal diffusion. Furthermore, we discovered that light emission with a narrow wavelength range at room temperature from the Si–QD with a β–FeSi2 core was due to radiative recombination between the electron-quantized state in the core and the hole-quantized state in the shell, indicating that the Si–QD with a β–FeSi2–core is a promising light-emitting material for Si-based photonics.

2. Materials and Methods

To systematically investigate the silicide process and selective Si growth, Si on insulator (SOI) substrates was used in this study, with a SiO2 thickness of ~145 nm. Figure 1a–d schematically illustrates the concept of the experiment. Initially, the Si layer on the SiO2 was thinned to ~10 nm by alternately controlling the thermal oxidation and wet etching using a diluted HF solution. Subsequently, ~1.0 nm thick Fe films were deposited via electron-beam evaporation. Film thickness in these experiments was evaluated by spectroscopic ellipsometry. The Fe film was then immediately removed using a HCl solution. Following this, the sample was exposed to SiH4 at 400 °C, with total gas pressure and exposure duration of 100 Pa and 1800 s, respectively.

3. Formation of Ultrathin Si/Fe-Silicide/Si Structures

Figure 1a′–d′ shows atomic force microscopy (AFM) images taken after each process step. The root mean square (RMS) roughness of the SOI substrate surface evaluated from the AFM image was as low as ~0.15 nm, which is almost the same as that before the thinning of the Si layer. AFM images of the surface taken after Fe film deposition and subsequent removal of the Fe film confirmed that the RMS roughness remained the same as before Fe deposition, as shown in Figure 1b′,c′. Furthermore, no significant change in the RMS roughness was observed after SiH4 exposure, indicating an extremely flat surface at each process step. To evaluate the chemical bonding features of the surface, we analyzed Si 2p3/2 and Fe 2p3/2 spectra of the Fe/SOI structure before and after HF etching and subsequent SiH4 exposure, as shown in Figure 1e and Figure 1f, respectively, measured by X-ray photoelectron spectroscopy (XPS) using monochromatized Al–Kα radiation ( = 1486.6 eV), with the binding energy of each spectrum normalized by the C 1s signal. However, an Fe oxide component in the Fe 2p spectrum after Fe deposition was observed due to air exposure during XPS measurements, which was hardly observed prior to HCl treatment. This result indicates that the unreacted Fe and Fe oxide were completely etched by the HCl treatment. Interestingly, the signal intensity originating from the Si–Si and/or Si–Fe component increased and slightly shifted towards the lower binding energy side after HCl treatment compared to that after Fe film deposition, suggesting that the silicidation reaction of the Si layer surface was induced just after Fe film deposition, despite the absence of heating. In other words, the silicidation reaction proceeds through the deposition of the Fe film onto an atomically flat Si surface, followed by its removal using HCl, without any post-deposition annealing. It is well established that, in silicide systems, the resulting crystal phase is determined by the composition ratio [50]. Therefore, in our process, the silicidation is self-limited due to Fe-supply limited kinetics, which leads to the formation of an abrupt interface. Following SiH4 exposure, the Si–Si and/or Si–Fe signal intensities slightly increased while the oxidation of Fe–silicide remains suppressed. Therefore, to evaluate the impact of changes in the chemical bonding features on the sample surface after the SiH4 exposure, the Si 2p3/2 spectra were also measured at a photoelectron take-off angle of 30°, as indicated in Figure 1g. Angle-resolved analysis of Si2p after SiH4 exposure showed a significant increase in the Si oxide component in the surface-sensitive measurement at a photoelectron take-off angle of 30°. These results might be attributed to the SiH4 decomposition on the Fe–silicide layer surface even at 400 °C, followed by the growth of a Si layer on the silicide layer surface. In the case of Ni silicide, it is known that the silicidation reaction proceeds at low temperature, as the Ni-adamantane structure serves as a precursor for the formation of NiSi2 [51]. Whether a similar mechanism is operative in the Fe silicide system remains an open question, which we aim to address in future studies.
The formation of the ultrathin Fe–silicide layer and Si growth on the silicide layer was confirmed through cross-sectional transmission electron microscopy (TEM) energy-dispersive X-ray spectroscopy (EDX) mapping images, as illustrated in Figure 2a–d. The TEM lamella is prepared by mechanical grinding and polishing followed by Ar ion milling. Despite the Fe film being deposited at room temperature and subsequently etched by HCl, the formation of an ultrathin Fe layer was observed. In contrast, Si signals were detected throughout the sample, irrespective of the film thickness direction. It was confirmed that the top surface of the sample was oxidized, and these results were consistent with those obtained from the XPS analysis. Furthermore, the growth of Si on the Fe surface was confirmed by analyzing the atomic concentration of each element in the film thickness direction, as evaluated from the cross-sectional profile as shown in Figure 2d, indicating the formation of a sandwich structure in which the Fe–silicide layer was sandwiched between the Si layers. It is noteworthy that the atomic concentration ratio of Fe to Si was close to 1:2, suggesting the formation of the FeSi2 phase.

4. Formation and Light Emission Properties of Fe-Silicide Core/Si-Shell Quantum Dots

Based on these results, we applied the same processes as discussed for the SOI substrate, as illustrated in Figure 1, to form a new type of superatom structure, namely, Fe–silicide core/Si–shell QD on an ultrathin SiO2 layer. Figure 3a–f shows the results of the AFM topographic images and the dot height distribution evaluated from each AFM image taken after each process step. Firstly, uniformly sized Si–QDs with an areal density as high as ~1011 cm−2 were self-assembled on a ~3.0 nm thick SiO2 layer/p–Si (100) by precisely controlling the low-pressure chemical vapor deposition (LPCVD) using SiH4 gas, where the average dot height calculated using a log-normal fitting function was estimated to be ~5.1 nm. After the ~1.0 nm thick Fe film deposition, the surface morphology was slightly smeared, and the average dot height slightly decreased to ~3.9 nm. However, after the HCl treatment, the dot height became the same size, which was ~5.1 nm as that of the as-grown Si–QDs. This result can be explained by complete removal of the unreacted Fe and Fe oxide by the HCl treatment. Subsequent SiH4 exposure at 400 °C resulted in a slight increase up to ~5.9 nm with no change in the areal dot density. We also confirmed that no Si deposition occurred on the SiO2 surface at 400 °C because the decomposition temperature of SiH4 was not reached. Therefore, a slight increase in the average dot height after SiH4 exposure can be explained by selective growth of Si on top of the dots. Considering the results of the SOI discussed earlier, these results are expected to form β–FeSi2 core/Si–Shell QDs, as schematically illustrated in Figure 3a′–d′. This indicates that the silicidation reaction of the Si–QDs’ surface occurred immediately after Fe film deposition. Subsequently, the unreactive Fe films were etched by the HCl treatment. Followed by SiH4 exposure, Si was selectively grown on the Fe silicide surface, acting as an outer cladding.
To clarify the crystalline phase of the silicide core, we measured the room temperature PL, using a semiconductor laser with a wavelength of 976 nm at an input power of ~0.33 W/cm2 as an excitation source and a cooled InGaAs detector, as shown in Figure 4a. Although a weak PL signal in the range from ~0.73 to 0.83 eV was observed immediately after the Fe deposition, light emission from the sample was hardly detected after HCl treatment. These results can be explained as follows: immediately after the Fe film deposition, as discussed in Figure 2, β–FeSi2 layer was formed on the Si–QDs surface despite the oxidation. On the other hand, after HCl treatment, non-radiative recombination centers at the β–FeSi2 surface dominated due to the complete etching of the surface oxide. It should be noted that, after SiH4 exposure, light emission with no significant change in wavelength was observed, and the PL intensity drastically increased by a factor of five compared to that after Fe film deposition. This result provides clear evidence for the formation of core/shell dots, indicating that the ultrathin Si cap layer was conformally covered with the dots, resulting in the suppression of defects on the β–FeSi2 surface. In addition, the PL signal was narrower than that of β–FeSi2 NDs with an average height of ~5 nm formed by SiH4 exposure of Fe nanodots, as shown in Figure 4b,c. This result can be interpreted in terms of the formation of very uniformly sized core/shell dots compared to the β–FeSi2 nanodots, as presented in Figure 4e,f, where the full-width and half-maxim value is ~2.4 nm for the β–FeSi2 NDs and ~1.9 nm for the core/shell QDs. To analyze the transition processes, the PL spectrum of the β–FeSi2 core/Si–shell QDs was precisely deconvoluted using four Gaussian curves and compared with that of the β–FeSi2 NDs, as indicated in Figure 4b,c. As discussed in ref. [48], the PL signal of the β–FeSi2 NDs can be deconvoluted into three main components, which peaked at ~0.79 (Component 1), ~0.72 (Component 2), and ~0.84 eV (Component 3) in Figure 4b, in which Comp. 1 can be attributed to the radiative recombination of photogenerated electron–hole pairs through discrete quantized states of β–FeSi2 NDs considering the bandgap of single-crystal β–FeSi2 bulk. In addition, the radiative transition between the higher-order quantized states in the conduction and valence bands of the β–FeSi2 NDs is likely responsible for the higher-energy Comp. 2, as illustrated in Figure 5a. Therefore, Comp. 3 can be attributed to defects in the β–FeSi2 NDs and/or the interface between the SiO2 and NDs. In contrast, the PL spectrum of the β–FeSi2 core/Si–shell QDs was deconvoluted into mainly two components, and the lowest component was hardly detected. However, even though the silicide layer of the core/shell dots is thinner than that of the dots, each PL component is shifted to lower energy compared to the components of the β–FeSi2 NDs. In semiconductor nanostructures, the PL peak energy shifts toward the higher energy side while decreasing in size due to quantum confinement effects. Therefore, this energy shift cannot be explained by the size effect of the semiconductor nanostructures. To evaluate the influence of core size in such core/shell structures, a systematic investigation involving the fabrication of Si-QD with different sized β–FeSi2 core, as demonstrated in ref. [31], is required. Nevertheless, considering the energy band diagram of the β–FeSi2 core/Si–Shell structure, Comp. 1, which has a lower energy emission, might be attributable to radiative recombination between the quantized states in the conduction band of the β–FeSi2 core and the valence band of the Si shell, as illustrated in Figure 5b, because electrons are confined in the conduction band of the β–FeSi2 core, and holes can be confined in the shallow potential well of the valence band of the Si clad and penetrate the β–FeSi2 core for radiative recombination in this system.
The results obtained in this work show that the β–FeSi2 core/Si–shell QDs have narrow PL spectra in the region of 0.74–0.84 eV at room temperature, suggesting that the radiative recombination of photo-generated carriers through quantized states of the core is the dominant pathway for the emission from the dots, reflecting the energy band discontinuity between the Si clad and FeSi2 core. This means that the β–FeSi2 core/Si–shell QDs can be regarded as active elements in optical applications with very few defects at the interface and/or surface, making them one of the most technologically important materials. Furthermore, our samples are highly compatible with the Si–ULSI processing technology. We believe that our findings will pave the way for constructing Si-based photonic processing that exhibits high thermal stability, low power consumption, and room temperature operation.

5. Conclusions

We fabricated highly dense β–FeSi2 core/Si–shell QDs with an areal density as high as ~1011 cm−2 on an ultrathin SiO2 layer and evaluated their room temperature light emission properties. We confirmed a clear light emission in the near-infrared region at room temperature. The origin of this light emission under weak excitation is considered to be the radiative recombination between the quantized states in the conduction band of the β–FeSi2 core and the valence band of the Si shell. Although the β–FeSi₂ core/Si shell QDs fabricated in this study are extremely small in size, their areal density likely results in limited absorption of incident light, leading to low quantum efficiency. Nonetheless, the light emission intensity could be significantly improved by increasing the dot density or by implementing multilayer stacking structures. From a technological point of view, it is quite important that such narrow light emission in the spectral width at room temperature could be achieved using the β–FeSi2 core/Si–shell QDs, which have very few defects at the interface between the QD and its surface. The results lead to the development of Si-based light-emitting devices that are highly compatible with Si–ULSI processing, which was found difficult to realize in silicon photonics.

Author Contributions

Conceptualization, K.M.; methodology, K.M.; formal analysis, K.M.; investigation, Y.Y. and M.A.S.; writing—original draft preparation, K.M.; project administration, A.M. and S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the Grant-in-Aid for Scientific Research (A) 21H04559, the Fund for the Promotion of Joint Intentional Research [Fostering Joint International Research (A)] 18KK0409 of MEXT Japan, and Hibi Science Foundation.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors thank the IHP cleanroom staff epitaxy process team for their excellent support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Soref, R.A.; Lorenzo, J.P. Single-crystal silicon: A new material for 1.3 and 1.6 μm integrated-optical components. Electron. Lett. 1985, 21, 953–954. [Google Scholar] [CrossRef]
  2. Schmidtchen, J.; Splett, A.; Schüppert, B.; Petermann, K.; Burbach, G. Low loss singlemode optical waveguides with large cross-section in silicon-on-insulator. Electron. Lett. 1991, 27, 1486–1488. [Google Scholar] [CrossRef]
  3. Weiss, B.L.; Reed, G.T.; Toh, S.K.; Soref, R.A.; Namavar, F. Optical waveguides in SIMOX structures. IEEE Photonics Technol. Lett. 1991, 3, 19–21. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, A.; Jones, R.; Liao, L.; Samara-Rubio, D.; Rubin, D.; Cohen, O.; Nicolaescu, R.; Paniccia, M. A high-speed silicon optical modulator based on a metal–oxide–semiconductor capacitor. Nature 2004, 427, 615–618. [Google Scholar] [CrossRef]
  5. Liao, L.; Samara-Rubio, D.; Morse, M.; Liu, A.; Hodge, D.; Rubin, D.; Keil, U.D.; Franck, T. High speed silicon mach-zehnder modulator. Opt. Express 2005, 13, 3129–3135. [Google Scholar] [CrossRef]
  6. Liao, L.; Liu, A.; Rubin, D.; Basak, J.; Chetrit, Y.; Nguyen, H.; Cohen, R.; Izhaky, N.; Paniccia, M. 40 Gbit/s silicon optical modulator for high-speed applications. Electron. Lett. 2007, 43, 1196–1197. [Google Scholar] [CrossRef]
  7. Dehlinger, G.; Koester, S.J.; Schaub, J.D.; Chu, J.O.; Ouyang, Q.C.; Grill, A. High-speed Germanium-on-SOI lateral PIN photodiodes. IEEE Photonics Technol. Lett. 2004, 16, 2547–2549. [Google Scholar] [CrossRef]
  8. Ahn, D.; Hong, C.-Y.; Liu, J.; Giziewicz, W.; Beals, M.; Kimerling, L.C.; Michel, J.; Chen, J.; Kärtner, F.X. High performance, waveguide integrated Ge photodetectors. Opt. Express 2007, 15, 3916–3921. [Google Scholar] [CrossRef]
  9. Shekhar, S.; Bogaerts, W.; Chrostowski, L.; Bowers, J.E.; Hochberg, M.; Soref, R.; Shastri, B.J. Roadmapping the next generation of silicon photonics. Nat. Commun. 2024, 10, 751. [Google Scholar] [CrossRef]
  10. Zakharov, N.D.; Talalaev, V.G.; Werner, P.; Tonkikh, A.A.; Cirlin, G.E. Room-temperature light emission from a highly strained Si/Ge superlattice. Appl. Phys. Lett. 2003, 83, 3084–3086. [Google Scholar] [CrossRef]
  11. Jambois, O.; Rinnert, H.; Devaux, X.; Vergnat, M. Photoluminescence and electroluminescence of size-controlled silicon nanocrystallites embedded in SiO2 thin films. J. Appl. Phys. 2005, 98, 046105. [Google Scholar] [CrossRef]
  12. Sun, K.W.; Sue, S.H.; Liu, C.W. Visible photoluminescence from Ge quantum dots. Phys. E Low-Dimens. Syst. Nanostruct. 2005, 28, 525–530. [Google Scholar] [CrossRef]
  13. del Pino, A.P.; Gyorgy, E.; Marcus, I.C.; Roqueta, J.; Alonso, M.I. Effects of pulsed laser radiation on epitaxial self-assembled Ge quantum dots grown on Si substrates. Nanotechnology 2011, 22, 295304. [Google Scholar] [CrossRef] [PubMed]
  14. Gassenq, A.; Guilloy, K.; Pauc, N.; Hartmann, J.-M.; Dias, G.O.; Rouchon, D.; Tardif, S.; Escalante, J.; Duchemin, I.; Niquet, Y.-M.; et al. Study of the light emission in Ge layers and strained membranes on Si substrates. Thin Solid Films 2016, 613, 64–67. [Google Scholar] [CrossRef]
  15. Schlykow, V.; Zaumseil, P.; Schubert, M.A.; Skibitzki, O.; Yamamoto, Y.; Klesse, W.M.; Hou, Y.; Virgilio, M.; De Seta, M.; Di Gaspare, L.; et al. Photoluminescence from GeSn nano-heterostructures. Nanotechnology 2018, 29, 415702. [Google Scholar] [CrossRef]
  16. Thai, Q.M.; Pauc, N.; Aubin, J.; Bertrand, M.; Chrétien, J.; Delaye, V.; Chelnokov, A.; Hartmann, J.-M.; Reboud, V.; Calvo, V. GeSn heterostructure micro-disk laser operating at 230 K. Opt. Express 2018, 26, 32500–32508. [Google Scholar] [CrossRef]
  17. Von den Driesch, N.; Stange, D.; Rainko, D.; Breuer, U.; Capellini, G.; Hartmann, J.-M.; Sigg, H.; Mantl, S.; Grützmacher, D.; Buca, D. Epitaxy of Si-Ge-Sn-based heterostructures for CMOS-integratable light emitters. Solid-State Electron. 2019, 155, 139–143. [Google Scholar] [CrossRef]
  18. Chrétien, J.; Pauc, N.; Pilon, F.A.; Bertrand, M.; Thai, Q.-M.; Casiez, L.; Bernier, N.; Dansas, H.; Gergaud, P.; Delamadeleine, E.; et al. GeSn Lasers Covering a Wide Wavelength Range Thanks to Uniaxial Tensile Strain. ACS Photonics 2019, 6, 2462–2469. [Google Scholar] [CrossRef]
  19. Armand Pilon, F.T.; Lyasota, A.; Niquet, Y.-M.; Reboud, V.; Calvo, V.; Pauc, N.; Widiez, J.; Bonzon, C.; Hartmann, J.-M.; Chelnokov, A.; et al. Lasing in strained germanium microbridges. Nat. Commun. 2019, 10, 2724. [Google Scholar] [CrossRef]
  20. Fadaly, E.M.T.; Dijkstra, A.; Suckert, J.R.; Ziss, D.; van Tilburg, M.A.J.; Mao, C.; Ren, Y.; van Lange, V.T.; Korzun, K.; Kölling, S.; et al. Direct bandgap emission from hexagonal Ge and SiGe alloys. Nature 2020, 580, 205–209. [Google Scholar] [CrossRef]
  21. Ji, Z.-M.; Luo, J.-W.; Li, S.-S. Interface-engineering enhanced light emission from Si/Ge quantum dots. New J. Phys. 2020, 22, 093037. [Google Scholar] [CrossRef]
  22. Jannesari, R.; Schatzl, M.; Hackl, F.; Glaser, M.; Hingerl, K.; Fromherz, T.; Schäffler, F. Commensurate germanium light emitters in silicon-on-insulator photonic crystal slabs. Opt. Express 2014, 22, 25426–25435. [Google Scholar] [CrossRef] [PubMed]
  23. Petykiewicz, J.; Nam, D.; Sukhdeo, D.S.; Gupta, S.; Buckley, S.; Piggott, A.Y.; Vučković, J.; Saraswat, K.C. Direct Bandgap Light Emission from Strained Germanium Nanowires Coupled with High-Q Nanophotonic Cavities. Nano Lett. 2016, 16, 2168–2173. [Google Scholar] [CrossRef] [PubMed]
  24. Reboud, V.; Gassenq, A.; Hartmann, J.M.; Widiez, J.; Virot, L.; Aubin, J.; Guilloy, K.; Tardif, S.; Fédéli, J.M.; Pauc, N.; et al. Germanium based photonic components toward a full silicon/germanium photonic platform. Prog. Cryst. Growth Charact. Mater. 2017, 63, 1–24. [Google Scholar] [CrossRef]
  25. Walters, R.J.; Bourianoff, G.I.; Atwater, H.A. Field-effect electroluminescence in silicon nanocrystals. Nat. Mater. 2005, 4, 143–146. [Google Scholar] [CrossRef]
  26. Xu, X.; Tsuboi, T.; Chiba, T.; Usami, N.; Maruizumi, T.; Shiraki, Y. Silicon-based current-injected light emitting diodes with Ge self-assembled quantum dots embedded in photonic crystal nanocavities. Opt. Express 2012, 20, 14714–14721. [Google Scholar] [CrossRef]
  27. Xu, X.; Usami, N.; Maruizumi, T.; Shiraki, Y. Enhancement of light emission from Ge quantum dots by photonic crystal nanocavities at room-temperature. J. Cryst. Growth 2013, 378, 636. [Google Scholar] [CrossRef]
  28. Xu, X.; Maruizumi, T.; Shiraki, Y. Waveguide-integrated microdisk light-emitting diode and photodetector based on Ge quantum dots. Opt. Express 2014, 22, 3902–3910. [Google Scholar] [CrossRef]
  29. Zeng, C.; Ma, Y.; Zhang, Y.; Li, D.; Huang, Z.; Wang, Y.; Huang, Q.; Li, J.; Zhong, Z.; Yu, J.; et al. Single germanium quantum dot embedded in photonic crystal nanocavity for light emitter on silicon chip. Opt. Express 2015, 23, 22250–22261. [Google Scholar] [CrossRef]
  30. Makihara, K.; Kondo, K.; Ikeda, M.; Ohta, A.; Miyazaki, S. Photoluminescence Study of Si Quantum Dots with Ge Core. ECS Trans. 2014, 64, 365–370. [Google Scholar] [CrossRef]
  31. Yamada, K.; Kondo, K.; Makihara, K.; Ikeda, M.; Ohta, A.; Miyazaki, S. Effect of Ge Core Size on Photoluminescence from Si Quantum Dots with Ge Core. ECS Trans. 2016, 75, 695–700. [Google Scholar] [CrossRef]
  32. Kondo, K.; Makihara, K.; Ikeda, M.; Miyazaki, S. Photoluminescence study of high density Si quantum dots with Ge core. J. Appl. Phys. 2016, 119, 033103. [Google Scholar] [CrossRef]
  33. Makihara, K.; Ikeda, M.; Fujimura, N.; Yamada, K.; Ohta, A.; Miyazaki, S. Electroluminescence of superatom-like Ge-core/Si-shell quantum dots by alternate field-effect-induced carrier injection. Appl. Phys. Express 2018, 11, 011305. [Google Scholar] [CrossRef]
  34. Makihara, K.; Yamamoto, Y.; Imai, Y.; Taoka, N.; Schubert, M.A.; Tillack, B.; Miyazaki, S. Room Temperature Light Emission from Superatom–like Ge–core/Si–shell Quantum Dots. Nanomaterials 2023, 13, 1475. [Google Scholar] [CrossRef]
  35. Bost, M.C.; Mahan, J.E. Optical properties of semiconducting iron disilicide thin films. J. Appl. Phys. 1985, 58, 2696–2703. [Google Scholar] [CrossRef]
  36. Katsumata, H.; Makita, Y.; Kobayashi, N.; Shibata, H.; Hasegawa, M.; Aksenov, I.; Kimura, S.; Obara, A.; Uekusa, S.-I. Optical absorption and photoluminescence studies of β-FeSi2 prepared by heavy implantation of Fe+ ions into Si. J. Appl. Phys. 1996, 80, 5955–5962. [Google Scholar] [CrossRef]
  37. Leong, D.; Harry, M.; Reeson, K.J.; Homewood, K.P. A silicon/iron-disilicide light-emitting diode operating at a wavelength of 1.5 μm. Nature 1997, 387, 686–688. [Google Scholar] [CrossRef]
  38. Suemasu, T.; Negishi, Y.; Takakura, K.; Hasegawa, F.; Chikyow, T. Influence of Si growth temperature for embedding β-FeSi2 and resultant strain in β-FeSi2 on light emission from p-Si/β-FeSi2 particles/n-Si light-emitting diodes. Appl. Phys. Lett. 2001, 79, 1804–1806. [Google Scholar] [CrossRef]
  39. Nakamura, T.; Tatsumi, T.; Morita, K.; Kobayashi, H.; Noguchi, Y.; Kawakubo, Y.; Maeda, Y. Photoluminescence Property of Nano-Composite Phases of β-FeSi2 Nanocrystals Embedded in SiO2. JJAP Conf. Proc. 2015, 3, 011106. [Google Scholar] [CrossRef]
  40. Lourenco, M.A.; Butler, T.M.; Kewell, A.K.; Gwilliam, R.M.; Kirkby, K.J.; Homewood, K.P. Electroluminescence of β-FeSi2 Light Emitting Devices. Jpn. J. Appl. Phys. 2001, 40, 4041–4044. [Google Scholar] [CrossRef]
  41. Martinelli, L.; Grilli, E.; Guzzi, M.; Grimaldi, M.G. Room-temperature electroluminescence of ion-beam-synthesized β-FeSi2 precipitates in silicon. Appl. Phys. Lett. 2003, 83, 794. [Google Scholar] [CrossRef]
  42. Li, C.; Ozawa, Y.; Suemasu, T.; Hasegawa, F. Thermal Enhancement of 1.6 µm Electroluminescence from a Si-Based Light-Emitting Diode with β-FeSi2 Active Region. Jpn. J. Appl. Phys. 2004, 43, L1492. [Google Scholar]
  43. Sunohara, T.; Kobayashi, K.; Suemasu, T. Epitaxial growth and characterization of Si-based light-emitting Si/β-FeSi2 film/Si double heterostructures on Si(001) substrates by molecular beam epitaxy. Thin Solid Films 2006, 508, 371–375. [Google Scholar] [CrossRef]
  44. Miyazaki, S.; Hamamoto, Y.; Yoshida, E.; Ikeda, M.; Hirose, M. Control of Self-Assembling Formation of Nanometer Silicon Dots by Low Pressure Chemical Vapor Deposition. Thin Solid Films 2000, 369, 55–59. [Google Scholar] [CrossRef]
  45. Takami, H.; Makihara, K.; Ikeda, M.; Miyazaki, S. Characterization of Electroluminescence from One-dimensionally Self-Aligned Si-based Quantum Dots. Jpn. J. Appl. Phys. 2013, 52, 04CG08. [Google Scholar] [CrossRef]
  46. Yamada, T.; Makihara, K.; Ohta, A.; Ikeda, M.; Miyazaki, S. Study on electroluminescence from multiply-stacking valency controlled Si quantum dots. Thin Solid Films 2016, 602, 48–51. [Google Scholar] [CrossRef]
  47. Furuhata, H.; Makihara, K.; Shimura, Y.; Fujimori, S.; Imai, Y.; Ohta, A.; Taoka, N.; Miyazaki, S. Study on Silicidation Reaction of Fe-NDs with SiH4. Appl. Phys. Express 2022, 15, 055503. [Google Scholar] [CrossRef]
  48. Miyazaki, S.; Makihara, K. Formation and Characterization of Fe-Silicide Nanodots for Optoelectronic Application. ECS Trans. 2023, 112, 131–137. [Google Scholar] [CrossRef]
  49. Saito, H.; Makihara, K.; Taoka, N.; Miyazaki, S. Formation of β-FeSi2 NDs by SiH4-exposure to Fe-NDs. Jpn. J. Appl. Phys. 2024, 63, 02SP99. [Google Scholar] [CrossRef]
  50. Canali, C.; Majni, G.; Ottaviani, G.; Celotti, G. Phase diagrams and metal-rich silicide formation. J. Appl. Phys. 1979, 50, 255–258. [Google Scholar] [CrossRef]
  51. Ikarashi, N.; Masuzaki, K. Silicide formation process in ultra-thin Ni-silicide film for advanced semiconductor devices: Mechanism of NiSi2 formation at low temperature. J. Phys. Conf. Ser. 2011, 326, 012051. [Google Scholar] [CrossRef]
Figure 1. (ad) Schematic illustration at each process step, and corresponding (a′d′) typical AFM images of (a,a′) SOI surface, (b,b′) taken after Fe film deposition, (c,c′) after HCl treatment, and (d,d′) after subsequent SiH4 exposure. (e) Fe2p, and (f,g) Si2p core–line spectra of Fe/SOI substrate before and after HCl treatment, and after SiH4 exposure taken at photoelectron take-off angle of (e,f) 90° and (g) 30°.
Figure 1. (ad) Schematic illustration at each process step, and corresponding (a′d′) typical AFM images of (a,a′) SOI surface, (b,b′) taken after Fe film deposition, (c,c′) after HCl treatment, and (d,d′) after subsequent SiH4 exposure. (e) Fe2p, and (f,g) Si2p core–line spectra of Fe/SOI substrate before and after HCl treatment, and after SiH4 exposure taken at photoelectron take-off angle of (e,f) 90° and (g) 30°.
Nanomaterials 15 00733 g001
Figure 2. (ac) Cross-sectional EDX mapping images and (d) cross-sectional profile of sample corresponding to Figure 1d′; in EDX mapping images, blue, red, and green colors correspond to Fe, Si, and O, respectively.
Figure 2. (ac) Cross-sectional EDX mapping images and (d) cross-sectional profile of sample corresponding to Figure 1d′; in EDX mapping images, blue, red, and green colors correspond to Fe, Si, and O, respectively.
Nanomaterials 15 00733 g002
Figure 3. (ad) Typical AFM images, (e,f) dot height distribution evaluated from AFM images, and (a′d′) schematic illustrations of (a,a′) pre-grown Si–QDs, (b,b′) after Fe film deposition, (c,c′) after HCl treatment, and (d,d′) subsequent SiH4 exposure, wherein dot height distribution and corresponding curves denote log-normal functions well fitted to measured size distribution.
Figure 3. (ad) Typical AFM images, (e,f) dot height distribution evaluated from AFM images, and (a′d′) schematic illustrations of (a,a′) pre-grown Si–QDs, (b,b′) after Fe film deposition, (c,c′) after HCl treatment, and (d,d′) subsequent SiH4 exposure, wherein dot height distribution and corresponding curves denote log-normal functions well fitted to measured size distribution.
Nanomaterials 15 00733 g003
Figure 4. (ac) Room temperature PL spectra of (a,c) NDs are illustrated in Figure 3, and (b) β–FeSi2–NDs taken under 967 nm light excitations, and (b,c) their deconvoluted spectra evaluated from spectral analysis using Gaussian curve fitting method, where synthetic spectrum of deconvoluted spectra is also shown as dashed line. Dot height distribution of (d) β–FeSi2 NDs and (e) β–FeSi2 core/Si–shell QDs.
Figure 4. (ac) Room temperature PL spectra of (a,c) NDs are illustrated in Figure 3, and (b) β–FeSi2–NDs taken under 967 nm light excitations, and (b,c) their deconvoluted spectra evaluated from spectral analysis using Gaussian curve fitting method, where synthetic spectrum of deconvoluted spectra is also shown as dashed line. Dot height distribution of (d) β–FeSi2 NDs and (e) β–FeSi2 core/Si–shell QDs.
Nanomaterials 15 00733 g004
Figure 5. Energy band diagram of (a) β–FeSi2–NDs and (b) β–FeSi2 core/Si–shell QDs, illustrating radiative transition corresponding to PL components as shown in Figure 4b,c.
Figure 5. Energy band diagram of (a) β–FeSi2–NDs and (b) β–FeSi2 core/Si–shell QDs, illustrating radiative transition corresponding to PL components as shown in Figure 4b,c.
Nanomaterials 15 00733 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Makihara, K.; Yamamoto, Y.; Schubert, M.A.; Mai, A.; Miyazaki, S. Processing and Characterization of High-Density Fe-Silicide/Si Core–Shell Quantum Dots for Light Emission. Nanomaterials 2025, 15, 733. https://doi.org/10.3390/nano15100733

AMA Style

Makihara K, Yamamoto Y, Schubert MA, Mai A, Miyazaki S. Processing and Characterization of High-Density Fe-Silicide/Si Core–Shell Quantum Dots for Light Emission. Nanomaterials. 2025; 15(10):733. https://doi.org/10.3390/nano15100733

Chicago/Turabian Style

Makihara, Katsunori, Yuji Yamamoto, Markus Andreas Schubert, Andreas Mai, and Seiichi Miyazaki. 2025. "Processing and Characterization of High-Density Fe-Silicide/Si Core–Shell Quantum Dots for Light Emission" Nanomaterials 15, no. 10: 733. https://doi.org/10.3390/nano15100733

APA Style

Makihara, K., Yamamoto, Y., Schubert, M. A., Mai, A., & Miyazaki, S. (2025). Processing and Characterization of High-Density Fe-Silicide/Si Core–Shell Quantum Dots for Light Emission. Nanomaterials, 15(10), 733. https://doi.org/10.3390/nano15100733

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop