Next Article in Journal
Investigations on Cavitation Erosion and Wear Resistance of High-Alloy WC Coatings Manufactured by Electric Arc Spraying
Previous Article in Journal
Reassessment of Porosimetry Determinations Using Water Vapor Sorption Measurements for Pastes and Concretes Containing Basaltic Aggregates Compared to the Mercury Intrusion Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optical Properties of GePb Alloy Realized by Ion Beam Technology

1
State Key Laboratory of Semiconductor Physics and Chip Technologies, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China
2
Helmholtz-Zentrum Dresden-Rossendorf, Institute of Ion Beam Physics and Materials Research, Bautzner Landstrasse 400, 01328 Dresden, Germany
3
Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
4
Institut für Werkstoffwissenschaft und Max-Bergmann-Zentrum für Biomaterialien, TUD Technische Universität Dresden, 01062 Dresden, Germany
5
Faculty of Electrical and Computer Engineering, Technische Universität Dresden, Helmholtzstraße 18, 01069 Dresden, Germany
6
Institute of Applied Physics, Technische Universität Dresden, 01062 Dresden, Germany
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(10), 2258; https://doi.org/10.3390/ma18102258
Submission received: 31 March 2025 / Revised: 28 April 2025 / Accepted: 6 May 2025 / Published: 13 May 2025
(This article belongs to the Section Optical and Photonic Materials)

Abstract

:
Incorporating lead (Pb) into the germanium (Ge) lattice emerges as a promising approach for bandgap engineering, enabling luminescence at longer wavelengths and paving the way for enhanced applications in short-wave infrared (SWIR) light sources and photodetectors. In this work, we report on optical properties of GePb alloys fabricated by a complementary metal-oxide semiconductor (CMOS)-compatible process that includes Pb ion implantation followed by solid-phase epitaxial regrowth via flash-lamp annealing. Optical characterization, including photoluminescence spectroscopy and Fourier-transform infrared reflectance spectroscopy, reveals that GePb alloys exhibit a reduced bandgap compared to pure Ge, resulting in longer-wavelength emission, while also providing broadband antireflective properties below 1800 nm wavelengths due to the surface subwavelength nanostructure. These findings position nanostructured GePb as a highly promising candidate for SWIR optoelectronic applications.

1. Introduction

The bandgap engineering of germanium (Ge) can be achieved by ultra-high n-type doping [1,2,3,4,5], introducing tensile strain [6,7] or alloying Ge with larger group-IV atoms, such as tin (Sn) [8,9,10,11,12,13]. These approaches modify the band structure, reducing the bandgap [14], thereby extending Ge emission to longer wavelengths [15] and significantly enhancing its potential for short-wave infrared (SWIR) optoelectronic applications.
Currently, alloying Ge with Sn has achieved great success in SWIR applications, including extended infrared GeSn photodetectors [16,17,18,19] and GeSn light sources [15,20,21], paving the way for on-chip optoelectronic integration. Similar to GeSn alloys, incorporating lead (Pb) into the Ge lattice can also reduce the bandgap and extend the optoelectronic applications to longer wavelengths. Due to its larger atomic radius compared to Sn [22,23], Pb induces a stronger bandgap modification when incorporated into the Ge lattice. First-principle calculations indicate that incorporating just 3.4% Pb can transform Ge into a direct-band structure, and higher Pb concentrations will further shift the absorption edge into the mid-infrared range [24,25,26]. Additionally, simulations show that the optical gain of GePb alloy increases abruptly with higher Pb concentration, indicating great promise in on-chip light-source applications [26].
However, obtaining GePb alloys with high Pb content remains challenging due to the limited equilibrium solid solubility of Pb in the Ge lattice, which is less than 0.5% at room temperature, thus requiring non-equilibrium fabrication methods [27,28]. Furthermore, the large atomic radius difference (23.2%) between Ge and Pb atoms limits the material quality of the regrowth of the GePb layer on the Ge substrate during the annealing process [29]. Ion implantation is a typical method to achieve non-equilibrium doping or alloying above the solid solubility. However, the high defect density in the implanted and annealed material significantly enhances nonradiative recombination, which quenches the photoluminescence (PL) [30,31] and limits optoelectronic applications.
In this work, we investigate the optical properties of GePb alloys, featuring a broadband antireflective subwavelength nanocavity layer using a CMOS-compatible process based on ion implantation combined with ultra-fast flash-lamp annealing (FLA). The Pb concentration of our GePb alloy is estimated to be 0.7%. While lead-free technologies are preferred in many CMOS applications, the minute concentrations of Pb in our GePb alloys minimize contamination concerns. We observed room-temperature PL in the GePb alloys, with spectral extension beyond 1900 nm, indicating low amounts of non-radiative defects [32] and broadband antireflection properties of GePb samples ascribed to the antireflective subwavelength nanocavity layer. We performed detailed characterization of the bandgap engineering through low-temperature PL measurements comparing these results with a reference Ge-virgin sample. PL emission from the GePb alloys was detectable from 1800 nm to 2060 nm at liquid nitrogen temperature (77 K). This work serves as a proof-of-concept for bandgap engineering in GePb materials. With wavelength-extended photoluminescence and broadband antireflection properties, the GePb material holds promise for SWIR optoelectronic applications.

2. Materials and Methods

The Ge wafer used in this work was a commercial intrinsic 4-inch (100) Ge wafer with a resistivity of 50 Ω·cm and a thickness of 300 μm. This double-side polished wafer was capped with a 40 nm thick SiO2 layer deposited via plasma-enhanced chemical vapor deposition (PECVD) before ion implantation. Pb ions were implanted into the Ge wafer at a fluence of 2 × 1015 cm−2 and an energy of 200 keV at room temperature, resulting in the formation of a GePb-implanted layer beneath the surface. During the ion implantation, the spot size of the ion beam was 5 mm × 5 mm. After Pb ion implantation, the GePb wafer was cut into 10 mm × 10 mm for subsequent flash-lamp annealing (FLA). We performed FLA on the implanted GePb samples in a continuous flow of N2 with firstly preheating at 180 °C for 60 s (heating rate of 3 K/s) and subsequently a flash with an energy density of 78 J/cm2 for 3.2 ms (heating rate of 105 K/s). The wafer temperature during FLA is estimated to be approximately 50 °C below the melting point of Ge. This estimation is based on the energy density of the flash (78 J/cm2) and the heating conditions, which were optimized to achieve solid-phase epitaxial regrowth without reaching the melting point of Ge.
The Raman scattering spectra were recorded in the backscattering geometry of HORIBA micro-Raman system using 100× objective (Olympus, Hachioji-shi, Japan), 532 nm pump laser (HORIBA Ltd., Palaiseau, France), and HORIBA Syncerity Si-CCD detector (HORIBA, Kyoto, Japan). Reflectance spectra were recorded in the Bruker VERTEX 70v Fourier-transform infrared (FTIR) spectroscopy system (Bruker, Billerica, MA, USA) with a Au mirror as the reference of 100% reflection. Photoluminescence (PL) spectra were recorded in a home-made micro-PL system equipped with a 532 nm laser, a 50× visible-infrared objective (Olympus, Hachioji-shi, Japan), and the HORIBA Symphony II InGaAs-CCD detector (HORIBA, Kyoto, Japan). Rutherford backscattering spectrometry (RBS) was carried out using the 1.7 MeV He+ beam of the Rossendorf van de Graff accelerator (Ion Beam Center, Helmholtz-Zentrum Dresden-Rossendorf, Dresden, Germany).
To characterize the microstructure of the GePb sample, bright-field transmission electron microscopy (TEM) imaging and high-resolution TEM imaging were performed with an image-Cs-corrected Titan 80–300 microscope (FEI, Eindhoven, The Netherlands) operated at an accelerating voltage of 300 kV. Prior to TEM analysis, the specimen mounted in a high-visibility low-background holder was placed for 8 s into a model 1020 Plasma Cleaner (Fischione, Export, PA, USA) to remove potential contamination. Cross-sectional TEM lamella preparation was performed by in situ lift-out using a Helios 5 CX focused ion beam (FIB) device (Thermo Fisher, Brno, Czech Republic). To protect the sample surface, a carbon cap layer was deposited via electron-beam-induced deposition, followed by Ga-FIB-assisted precursor decomposition. Afterwards, the TEM lamella was prepared using a 30-keV Ga-FIB with adapted currents. Its transfer to a 3-post copper lift-out grid (Omniprobe, High Wycombe, UK) was performed with an EasyLift EX nanomanipulator (Thermo Fisher, Brno, Czech Republic). To minimize sidewall damage, Ga ions with only 5-keV energy were used for the final thinning of the TEM lamella to electron transparency. Top-down scanning electron microscopy (SEM) imaging of the flash-lamp-annealed GePb sample was performed with an S-4800 microscope (Hitachi, Tokyo, Japan) operated at an accelerating voltage of 5 kV.

3. Results and Discussion

Figure 1 displays the micro-Raman spectra obtained from virgin Ge, as-implanted GePb, and flash-lamp-annealed GePb samples. The main peak near 300.5 cm−1 in the reference Ge-virgin sample corresponds to the Ge-Ge transverse optical (TO) phonon mode. It depends on the Ge-Ge bond length of the Ge lattice, which can be affected by the presence of alloying atoms and the strain conditions within the Ge lattice [4]. After Pb ion implantation, the TO phonon mode of Ge disappears, indicating an amorphous structure due to implantation-related crystal damage. In the FLA-treated GePb sample, the TO phonon mode is located at 300.0 cm−1, showing a slight shift toward lower wavenumbers compared to pristine Ge. This shift is attributed to the increased Ge-Ge bond length due to the larger Pb atoms incorporated into the lattice. The full width at half maximum (FWHM) of the TO phonon mode in the FLA-treated GePb sample is 5.5 cm−1, significantly broader than that of pristine Ge (2.7 cm−1), indicating the presence of residual lattice damage in the annealed GePb layer.
We utilized Rutherford backscattering spectrometry (RBS) in a random direction to confirm successful Pb implantation and analyze the distribution of Pb before and after the annealing process (see Figure 2). The SiO2 capping layer was removed by HF etching before RBS measurement. In both the as-implanted (Asimpl.) and FLA-treated GePb samples, Pb contribution appears as a peak in the spectra at an energy between 1450 and 1600 keV. The Pb profile shows an intensity decrease after FLA, which indicates the loss of the Pb element in the annealed sample. This can be ascribed to the flash-lamp annealing process being carried out on a millisecond timescale and the Pb alloying concentration exceeding the solid solubility. During the annealing process, Pb atoms can redistribute, resulting in an overall decrease in concentration. Some of these Pb atoms may diffuse to the surface and are subsequently removed by HF etching. As demonstrated by Raman spectroscopy, the Pb-implanted Ge layer becomes crystalline. The effective crystallization during the FLA process was later confirmed by a cross-sectional TEM study. Unfortunately, the actual Pb concentration cannot be calculated using the RBS profile due to the porous surface formed during ion implantation. However, based on simulations using the Stopping and Range of Ions in Matter (SRIM) software (SRIM-2013) [33], combined with RBS measurements, we estimate the nominal Pb concentration to be approximately 0.7%. Further details on the SRIM simulation and Pb concentration estimation are provided in the Supplementary Materials.
Figure 3a–e summarize cross-sectional transmission electron microscopy (TEM) results of the as-implanted GePb and flash-lamp-annealed GePb samples. In particular, the overview bright-field TEM image in Figure 3a and the magnified view in Figure 3c show that, after Pb ion implantation, the GePb surface structure is fully amorphous, with a clear interface between the amorphous GePb-implanted layer (appearing light gray) and the single-crystalline Ge substrate (appearing dark gray). Regarding morphology, the GePb layer is not homogeneous but shows a cavity structure within the first 130 nm below the SiO2 capping layer and a 60 nm homogeneous amorphous GePb layer between this cavity structure and the single-crystalline Ge substrate. Based on the TEM images, we estimate that the maximum implantation depth of Pb reaches approximately 190 nm below the surface. The relatively high atomic weight of Pb and the high implantation energy result in the formation of the cavity-rich surface structure during implantation.
This typical surface morphology can also be observed in Ge implanted with heavy ions (such as Mn [34], In [35], and Sb [36]) at room temperature and arises from the formation and accumulation of vacancy-related clusters, which minimize the dangling bond density [37,38]. During the ion implantation process, the ion beam collides with the Ge lattice, creating a vacancy-rich region close to the surface. Vacancy clustering in this near-surface region accelerates ion collisions with the Ge substrate [39], leading to the growth of cavity structures perpendicular to the sample surface [40]. However, there is not yet a conclusive explanation for the various nanostructures observed in Ge upon implantation [41]. Furthermore, Ref [42] demonstrates the potential to enhance the ordering of self-organized surface nanostructures and achieve large-area uniformity by increasing the implantation fluence.
Similar surface morphology is also observed in GeSn alloys fabricated through room-temperature ion implantation [43,44,45]. However, pulsed-laser annealing destroys these cavity structures due to the melting and liquid-phase epitaxial regrowth process, hindering the realization of group-IV Ge alloys with surface morphology. In this work, Pb’s significantly higher atomic weight compared to Sn leads to the formation of high-aspect-ratio cavity structures. In addition, the ultra-fast solid-phase epitaxial regrowth achieved by FLA preserves these high-aspect-ratio cavity structures during recrystallization.
As illustrated in Figure 3b,d, the flash-lamp annealing process yields solid-phase epitaxial regrowth for the 60 nm homogeneous amorphous GePb layer. However, for the sidewalls of the GePb cavity structure (shown in detail in Figure 3e), epitaxial regrowth occurs only in the lower quarter up to one-third, where the diffractogram indicates a single-crystalline GePb structure. In contrast, the upper cavity sidewalls exhibit a polycrystalline structure, indicating random nucleation during the regrowth process. Such random nucleation may be due to an inhomogeneous temperature distribution within the cavity structure, which might have a faster cooling rate than the GePb layer, and the absence of regrowth seeds during the recrystallization process [46], thus resulting in only polycrystal formation in this region (Figure 3e). High-resolution TEM images of the as-implanted and FLA-treated GePb samples, highlighting the clear contrast between the amorphous layer and the epitaxially regrown crystalline layer, are presented in the Supplementary Materials (Figure S2). The surface morphology of the FLA-treated sample is shown in the top-view SEM image in Figure 3f. While the black regions in the SEM image correspond to cavities in the GePb surface structure, the bright regions represent the corresponding side walls. While TEM provides a clear view of the nanocavity distribution and crystal structure in the vertical direction, it is limited in revealing the lateral distribution of the nanocavities. This is because the FIB-prepared sample is laminar, making it difficult to observe details such as the shape and diameter distribution of the nanocavities. In contrast, the top-view SEM image offers a comprehensive perspective on these lateral characteristics, providing valuable complementary information.
Here, we achieved partial solid-phase epitaxial recrystallization using an ultra-fast FLA process, which occurs on a millisecond scale. This allowed us to attain recrystallization while preserving the surface nanostructure. This surface morphology, featuring cavity structures with diameters in the tens of nanometers, exhibits antireflective properties over a broadband wavelength range, as confirmed by following optical characterization, positioning GePb as a promising material for optoelectronic applications.
After implantation, the GePb sample exhibited a dark color visible to the naked eye (Figure 4a, right), indicating broadband antireflective properties of subwavelength nanostructures led by Pb implantation. In contrast, the polished Ge reference wafer (Figure 4a, left) shows high reflectance in the visible range. Figure 4b shows the optical reflectance spectra of the Ge-virgin sample, as-implanted GePb, and flash-lamp-annealed GePb samples. The Ge-virgin reference sample exhibits a characteristic reflection increase at the edge of its bandgap, attributed to backscattered light from the back surface of the Ge substrate and high transmission below its energy bandgap [47]. Following Pb implantation, a significant decrease in reflectance for wavelengths shorter than the bandgap energy was observed, indicating broadband antireflective properties in the GePb samples. After flash-lamp annealing, the reflectance remained substantially lower than that of the Ge-virgin sample within the same wavelength range, suggesting successful preservation of the cavity surface structure during solid-phase epitaxial regrowth via flash-lamp annealing.
In this work, the size of our nanostructures is much smaller than the wavelength, and the mechanism differs from typical “light trapping” structures [48,49], which generally have features on the order of the wavelength. The antireflective effect in the GePb sample can be attributed to the smooth gradient in the refractive index at the air–substrate interface, facilitated by the subwavelength nanocavity structure on the GePb surface (as shown in Figure 3). Polished Ge wafers exhibit a reflectance exceeding 35% across wavelengths of 350–1800 nm, primarily due to the significant refractive index contrast between Ge (refractive index ~4) and air (refractive index ~1), which causes Fresnel reflections at the polished surface [47]. Subwavelength nanostructures mitigate this issue by creating a smooth refractive index gradient at the interface [50,51], allowing the surface to function as a broadband antireflective layer on germanium [52,53]. In this work, the subwavelength nanocavity structure induced by Pb implantation enables the GePb surface layer to function effectively as a broadband antireflective layer. Such a layer reduces reflectance from the Ge surface and enhances device performance in optoelectronic applications, including photodetectors [54] and solar cells [47]. While the primary goal of adding Pb is to shrink the bandgap and extend the optical activity into the SWIR region, the synergistic effect of reduced reflectance and bandgap modification makes GePb a promising candidate for optoelectronic applications.
However, characterizing the bandgap engineering of GePb using absorptance data, which combines reflectance and transmission measurements, is highly challenging for our GePb material. This difficulty arises primarily due to the overlapping absorption between the GePb layer and the Ge substrate. The GePb layer is only 200 nm thick, whereas the Ge substrate is 300 μm thick and exhibits strong absorption below the bandgap. The Ge substrate has a transmission of less than 40% beyond 2 μm, likely caused by sub-bandgap impurity absorption, as reported in Ref. [55]. This overlap complicates isolating the absorption signature of the implanted layer.
Typically, high-fluence ion implantation results in a high concentration of defects in the material, which are challenging to remove during subsequent ultra-fast annealing processes. These remaining detects strongly contribute to nonradiative recombination, limiting the application of high-fluence implantation in light source applications. In this work, we successfully conducted PL measurements on the FLA-treated GePb sample under 532 nm laser excitation, and a new emission peak can be observed under 77 K, which might indicate the bandgap engineering of Ge realized by Pb alloying.
In Figure 5, we present the normalized PL signal from Ge-virgin and FLA-treated GePb samples under room temperature (a) and 77 K (b). For the measurements under room temperature (Figure 5a), the PL spectrum of the FLA-treated GePb sample exhibits a spectral broadening mainly toward the longer wavelength (red shift), indicating a bandgap engineering through Pb alloying, which can reduce the bandgap of Ge and shift the PL emission toward the longer wavelength [26]. However, the main PL emission peak is still near 1550 nm, indicating that the PL signal is still strongly influenced by the contributions of Ge substrate. This can be ascribed to the thickness of the Pb alloyed layer, which is less than 200 nm thick with much stronger defect-related nonradiative recombination than Ge substrate. The strong PL emission from the defect-free Ge substrate overlaps with the PL contribution from the GePb layer, making it challenging to distinguish but still shows a peak extension toward the longer wavelength. The signal decrease after 2060 nm is due to the responsivity cutoff of the InGaAs array detector used in the PL measurement system.
In Figure 5b, we compare the normalized PL spectra from the Ge-virgin sample and the FLA-treated GePb sample under 77 K. The direct-band transition near 1400 nm from both Ge and GePb layers is not dominant because carriers are frozen in the indirect-band L valley in the Ge conduction band. For the Ge-virgin sample at 77 K, the sharp PL peak at 1732 nm is attributed to the indirect-band transition from Ge, which shifts to 1700 nm in the GePb layer, possibly due to energy band modification from compressive strain conditions originating from the Ge/GePb interface where the Pb concentration is lower than the peak concentration. However, a new emission peak from 1800 to 2060 nm (detector cutoff) can be observed in our FLA-treated GePb sample, which is not observed in the pristine Ge sample. This provides evidence of Pb alloying into the Ge lattice and reducing the band gap of the L valley, making it a promising candidate for extended SWIR light source application.
The GePb-related PL emission cannot be resolved at room temperature, which can be attributed to the low PL intensity resulting from the limited thickness of the GePb layer and the strong overlap with the broad PL emission tail from the Ge substrate. This overlap leads to an extension of the overall PL spectrum toward longer wavelengths, but no distinct emission peak from GePb can be identified. As the temperature decreases, the PL contribution from the phonon sideband (PSB) of the Ge indirect bandgap transition is significantly reduced, and the zero-phonon line (ZPL) becomes narrower, in agreement with the Maxwell–Boltzmann probability distribution function [56]. As a result, the spectral overlap between the Ge and GePb emissions is minimized, allowing a distinct PL peak from GePb to be resolved.
Photoluminescence measurements of the GePb-FLA sample at various temperatures are shown in Figure 6a. As the measurement temperature decreases, the PL contribution from the direct bandgap transition of Ge substrate near 1550 nm shows both blue shift and significant intensity decreases due to carriers being frozen in the indirect band L valley of the Ge conduction band. The PL emission from the GePb layer becomes evident when the sample temperature drops below 123 K. This emission spectrum appears as a broad peak starting from 1800 nm and extending to the detector cutoff at 2060 nm, indicating bandgap modification via Pb alloying into the Ge lattice.
A detailed PL comparison of the GePb sample at 123 K and 77 K is shown in Figure 6b. With a temperature decrease from 123 K to 77 K, the PL emission peak between 1800 and 2060 nm shifts to shorter wavelengths, and the PL intensity increases as the temperature decreases, which is consistent with the behavior of the indirect-band transition in Ge [57], and the longer emission wavelength than virgin Ge also agrees well with the theorical predicted bandgap engineering of Pb alloyed in Ge [26], indicating that the PL emission peak between 1800 and 2060 nm might originate from the indirect-band transition of bandgap-engineered GePb material. The relatively larger full width at half maximum (FWHM) of the PL peak from the GePb layer can be attributed to inhomogeneous Pb concentration in the implanted layer, which typically follows a Gaussian distribution. The sharp decrease beyond 2060 nm is due to the cut-off of the PL measurement system.
We further discuss and aim to exclude the possibility that the PL signal originates from structures potentially present in our sample, such as Ge nanocrystals, polycrystalline Ge, and defects in Ge. This is achieved by comparing our PL results with previously reported PL characteristics from these structures:
  • Ge Nanocrystals: As reported in Refs. [58,59], Ge nanocrystals exhibit a higher bandgap than single-crystal Ge, which is particle-size-dependent. Their PL emission is strongly broadened, extending toward shorter wavelengths, and can cover the visible range up to the Ge bandgap at room temperature. In contrast, the photoluminescence observed in our work predominantly extends toward longer wavelengths, which is inconsistent with this origin.
  • Polycrystalline Ge: According to Ref. [56], the PL emission of polycrystalline Ge shifts slightly to shorter wavelengths compared to single-crystal Ge. While our GePb material contains polycrystalline Ge, no such blue-shifted PL spectra are observed in our results.
  • Defects in Ge: To the best of our knowledge, the only reported defect-related luminescence in Ge is threading dislocation luminescence, which produces a sharp PL peak at 2.4 μm and does not shift with temperature (Refs. [56,60,61]). Since this emission lies outside our measurement range, it cannot explain our PL results. Furthermore, most defect-related PL in semiconductors arises from fixed energy levels, producing narrow emission lines (Refs. [62,63]), which is not consistent with our observations. The new emission peak observed in our FLA-treated GePb sample is distributed around 1800–2000 nm at 77 K and shifts with temperature. This behavior cannot be attributed to defect-related luminescence.

4. Conclusions

In summary, we have shown the extension of the luminescence emission to longer wavelengths in Ge using a CMOS-compatible process involving Pb ion implantation and flash-lamp annealing. We fabricated GePb alloys integrated with the antireflective subwavelength nanostructure via ultra-fast solid-phase epitaxial regrowth. Photoluminescence spectroscopy at varying temperatures revealed bandgap modification and extended emission wavelengths in the GePb alloys. Notably, our ion-beam-implanted and flash-lamp-annealed GePb alloys exhibit two significant optical characteristics: reduced optical reflectance due to the subwavelength cavity structure and bandgap shrinkage toward longer wavelengths attributed to Pb alloying. These findings position flash-lamp-annealed GePb alloys as promising candidates for various optoelectronic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18102258/s1, Figure S1. Pb distribution profile simulated by the SRIM software. Figure S2. High-resolution TEM images of as-implanted GePb (a, b) and FLA-treated GePb (c, d). Panels (b) and (d) are enlarged views of the quadratic regions in panels (a) and (c), and indicate the amorphous structure of as-implanted GePb and the single-crystalline structure of FLA-treated GePb, respectively. The fast Fourier transforms of panels (b) and (d) are presented as the corresponding insets.

Author Contributions

Conceptualization, S.W. and Y.B.; methodology, S.W.; validation, S.W. and Y.B.; formal analysis, S.W.; investigation, S.W., Y.-H.Z., O.S., M.S.S., S.P., R.H., A.W., L.H. and Y.B.; resources, M.H., S.Z., J.-W.L. and Y.B.; data curation, S.W.; writing—original draft preparation, S.W. and Y.B.; writing—review and editing, O.S., S.P., R.H., L.H., M.H., S.Z., J.-W.L. and Y.B.; visualization, S.W.; supervision, Y.B. and J.-W.L.; project administration, Y.B.; funding acquisition, Y.B. All authors have read and agreed to the published version of the manuscript.

Funding

Support by the Ion Beam Center (IBC) at HZDR is gratefully acknowledged. This work was financially supported by the German Research Foundation (DFG, BE7060/4-1) and the Hi-acts Use Case Initiatives 2023 of the Helmholtz Association. The authors thank Annette Kunz for TEM specimen preparation. Furthermore, the funding of TEM Talos by the German Federal Ministry of Education and Research (BMBF), Grant No. 03SF0451, in the framework of HEMCP is acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Spitzer, W.G.; Trumbore, F.A.; Logan, R.A. Properties of Heavily Doped n-Type Germanium. J. Appl. Phys. 1961, 32, 1822–1830. [Google Scholar] [CrossRef]
  2. Camacho-Aguilera, R.E.; Cai, Y.; Patel, N.; Bessette, J.T.; Romagnoli, M.; Kimerling, L.C.; Michel, J. An Electrically Pumped Germanium Laser. Opt. Express 2012, 20, 11316. [Google Scholar] [CrossRef]
  3. Prucnal, S.; Berencén, Y.; Wang, M.; Rebohle, L.; Böttger, R.; Fischer, I.A.; Augel, L.; Oehme, M.; Schulze, J.; Voelskow, M.; et al. Ex Situ n+ Doping of GeSn Alloys via Non-Equilibrium Processing. Semicond. Sci. Technol. 2018, 33, 065008. [Google Scholar] [CrossRef]
  4. Prucnal, S.; Berencén, Y.; Wang, M.; Rebohle, L.; Kudrawiec, R.; Polak, M.; Zviagin, V.; Schmidt-Grund, R.; Grundmann, M.; Grenzer, J.; et al. Band Gap Renormalization in N-Type GeSn Alloys Made by Ion Implantation and Flash Lamp Annealing. J. Appl. Phys. 2019, 125, 203105. [Google Scholar] [CrossRef]
  5. Carraro, C.; Milazzo, R.; Sgarbossa, F.; Fontana, D.; Maggioni, G.; Raniero, W.; Scarpa, D.; Baldassarre, L.; Ortolani, M.; Andrighetto, A.; et al. N-Type Heavy Doping with Ultralow Resistivity in Ge by Sb Deposition and Pulsed Laser Melting. Appl. Surf. Sci. 2020, 509, 145229. [Google Scholar] [CrossRef]
  6. Sánchez-Pérez, J.R.; Boztug, C.; Chen, F.; Sudradjat, F.F.; Paskiewicz, D.M.; Jacobson, R.; Lagally, M.G.; Paiella, R. Direct-Bandgap Light-Emitting Germanium in Tensilely Strained Nanomembranes. Proc. Natl. Acad. Sci. USA 2011, 108, 18893–18898. [Google Scholar] [CrossRef]
  7. Jain, J.R.; Hryciw, A.; Baer, T.M.; Miller, D.A.B.; Brongersma, M.L.; Howe, R.T. A Micromachining-Based Technology for Enhancing Germanium Light Emission via Tensile Strain. Nat. Photon 2012, 6, 398–405. [Google Scholar] [CrossRef]
  8. Atalla, M.R.M.; Assali, S.; Koelling, S.; Attiaoui, A.; Moutanabbir, O. High-Bandwidth Extended-SWIR GeSn Photodetectors on Silicon Achieving Ultrafast Broadband Spectroscopic Response. ACS Photonics 2022, 9, 1425–1433. [Google Scholar] [CrossRef]
  9. Talamas Simola, E.; Kiyek, V.; Ballabio, A.; Schlykow, V.; Frigerio, J.; Zucchetti, C.; De Iacovo, A.; Colace, L.; Yamamoto, Y.; Capellini, G.; et al. CMOS-Compatible Bias-Tunable Dual-Band Detector Based on GeSn/Ge/Si Coupled Photodiodes. ACS Photonics 2021, 8, 2166–2173. [Google Scholar] [CrossRef]
  10. Prucnal, S.; Berencén, Y.; Wang, M.; Grenzer, J.; Voelskow, M.; Hübner, R.; Yamamoto, Y.; Scheit, A.; Bärwolf, F.; Zviagin, V.; et al. Strain and Band-Gap Engineering in Ge–Sn Alloys via P Doping. Phys. Rev. Appl. 2018, 10, 064055. [Google Scholar] [CrossRef]
  11. Casiez, L.; Cardoux, C.; Acosta Alba, P.; Bernier, N.; Richy, J.; Pauc, N.; Calvo, V.; Coudurier, N.; Rodriguez, P.; Concepción, O.; et al. Ex-Situ n-Type Doped Carrier-Injection Layers in Direct Bandgap GeSn LEDs. Mater. Sci. Semicond. Process. 2024, 182, 108654. [Google Scholar] [CrossRef]
  12. Di Russo, E.; Sgarbossa, F.; Ranieri, P.; Maggioni, G.; Ndiaye, S.; Duguay, S.; Vurpillot, F.; Rigutti, L.; Rouvière, J.-L.; Morandi, V.; et al. Synthesis of Relaxed Ge0.9Sn0.1/Ge by Nanosecond Pulsed Laser Melting. Appl. Surf. Sci. 2023, 612, 155817. [Google Scholar] [CrossRef]
  13. Fontana, D.; Sgarbossa, F.; Milazzo, R.; Di Russo, E.; Galluccio, E.; De Salvador, D.; Duffy, R.; Napolitani, E. Ex-Situ n-Type Heavy Doping of Ge1-xSnx Epilayers by Surface Sb Deposition and Pulsed Laser Melting. Appl. Surf. Sci. 2022, 600, 154112. [Google Scholar] [CrossRef]
  14. Li, M.; Zheng, J.; Liu, X.; Zhu, Y.; Niu, C.; Pang, Y.; Liu, Z.; Zuo, Y.; Cheng, B. Sn Composition Graded GeSn Photodetectors on Si Substrate with Cutoff Wavelength of 3.3 μm for Mid-Infrared Si Photonics. Appl. Phys. Lett. 2022, 120, 121103. [Google Scholar] [CrossRef]
  15. Wirths, S.; Geiger, R.; Von Den Driesch, N.; Mussler, G.; Stoica, T.; Mantl, S.; Ikonic, Z.; Luysberg, M.; Chiussi, S.; Hartmann, J.M.; et al. Lasing in Direct-Bandgap GeSn Alloy Grown on Si. Nat. Photon 2015, 9, 88–92. [Google Scholar] [CrossRef]
  16. Liu, X.; Zheng, J.; Niu, C.; Liu, T.; Huang, Q.; Li, M.; Zhang, D.; Pang, Y.; Liu, Z.; Zuo, Y.; et al. Sn Content Gradient GeSn with Strain Controlled for High Performance GeSn Mid-Infrared Photodetectors. Photon. Res. 2022, 10, 1567. [Google Scholar] [CrossRef]
  17. Li, M.; Zheng, J.; Liu, X.; Niu, C.; Zhu, Y.; Pang, Y.; Liu, Z.; Yang, Y.; Zuo, Y.; Cheng, B. GeSn Resonance Cavity Enhanced Photodetector with Gold Bottom Reflector for the L Band Optical Communication. Opt. Lett. 2022, 47, 4315. [Google Scholar] [CrossRef]
  18. Tsai, C.-H.; Lin, K.-C.; Cheng, C.-Y.; Lee, K.-C.; Cheng, H.H.; Chang, G.-E. GeSn Lateral P-i-n Waveguide Photodetectors for Mid-Infrared Integrated Photonics. Opt. Lett. 2021, 46, 864. [Google Scholar] [CrossRef]
  19. Wen, S.; Shaikh, M.S.; Steuer, O.; Prucnal, S.; Grenzer, J.; Hübner, R.; Turek, M.; Pyszniak, K.; Reiter, S.; Fischer, I.A.; et al. Room-Temperature Extended Short-Wave Infrared GeSn Photodetectors Realized by Ion Beam Techniques. Appl. Phys. Lett. 2023, 123, 081109. [Google Scholar] [CrossRef]
  20. Huang, B.-J.; Chang, C.-Y.; Hsieh, Y.-D.; Soref, R.A.; Sun, G.; Cheng, H.-H.; Chang, G.-E. Electrically Injected GeSn Vertical-Cavity Surface Emitters on Silicon-on-Insulator Platforms. ACS Photonics 2019, 6, 1931–1938. [Google Scholar] [CrossRef]
  21. Wu, S.; Wang, Z.; Zhang, L.; Chen, Q.; Wen, S.; Lee, K.H.; Bao, S.; Fan, W.; Seng, T.C.; Luo, J.-W. Enhanced Light Emission of Germanium Light-Emitting-Diode on 150 Mm Germanium-on-Insulator (GOI). Opt. Express 2023, 31, 17921. [Google Scholar] [CrossRef] [PubMed]
  22. Slater, J.C. Atomic Radii in Crystals. J. Chem. Phys. 1964, 41, 3199–3204. [Google Scholar] [CrossRef]
  23. Suresh, C.H.; Koga, N. A Consistent Approach toward Atomic Radii. J. Phys. Chem. A 2001, 105, 5940–5944. [Google Scholar] [CrossRef]
  24. Huang, W.; Cheng, B.; Xue, C.; Li, C. Comparative Studies of Clustering Effect, Electronic and Optical Properties for GePb and GeSn Alloys with Low Pb and Sn Concentration. Phys. B Condens. Matter 2014, 443, 43–48. [Google Scholar] [CrossRef]
  25. Broderick, C.A.; O’Halloran, E.J.; Kirwan, A.C.; Dunne, M.D.; Tanner, D.S.P.; Schulz, S.; O’Reilly, E.P. Multi-Scale Electronic Structure Analysis of Direct-Gap Group-IV Alloys: Implications for Device Applications. In Proceedings of the 2019 IEEE 16th International Conference on Group IV Photonics (GFP), Singapore, 28–30 August 2019; IEEE: Singapore, 2019; pp. 1–2. [Google Scholar]
  26. Huang, W.; Cheng, B.; Xue, C.; Yang, H. The Band Structure and Optical Gain of a New IV-Group Alloy GePb: A First-Principles Calculation. J. Alloys Compd. 2017, 701, 816–821. [Google Scholar] [CrossRef]
  27. Liu, X.; Zheng, J.; Zhao, Y.; Li, M.; Peng, L.; Wan, F.; Niu, C.; Liu, Z.; Zuo, Y.; Xue, C.; et al. Germanium Lead Alloy on Insulator Grown by Rapid Melting Growth. J. Alloys Compd. 2021, 864, 158798. [Google Scholar] [CrossRef]
  28. Liu, X.; Zheng, J.; Li, X.; Niu, C.; Peng, L.; Wan, F.; Liu, Z.; Zuo, Y.; Xue, C.; Cheng, B. Investigation of Lead Surface Segregation during Germanium–Lead Epitaxial Growth. J. Mater. Sci. 2020, 55, 4762–4768. [Google Scholar] [CrossRef]
  29. Zhou, Q.; Chan, T.K.; Lim, S.L.; Zhan, C.; Osipowicz, T.; Gong, X.; Tok, E.S.; Yeo, Y.-C. Single Crystalline Germanium-Lead Alloy on Germanium Substrate Formed by Pulsed Laser Epitaxy. ECS Solid State Lett. 2014, 3, P91–P93. [Google Scholar] [CrossRef]
  30. Harding, R.E.; Davies, G.; Hayama, S.; Coleman, P.G.; Burrows, C.P.; Wong-Leung, J. Photoluminescence Response of Ion-Implanted Silicon. Appl. Phys. Lett. 2006, 89, 181917. [Google Scholar] [CrossRef]
  31. Harding, R.; Davies, G.; Coleman, P.G.; Burrows, C.P.; Wong-Leung, J. Study of Defects in Ion-Implanted Silicon Using Photoluminescence and Positron Annihilation. Phys. B Condens. Matter 2003, 340–342, 738–742. [Google Scholar] [CrossRef]
  32. Tran, T.T.; Hudspeth, Q.; Liu, Y.; Smillie, L.A.; Wang, B.; Bruce, R.A.; Mathews, J.; Warrender, J.M.; Williams, J.S. Ion Beam Synthesis and Photoluminescence Study of Supersaturated Fully-Relaxed Ge-Sn Alloys. Mater. Sci. Eng. B 2020, 262, 114702. [Google Scholar] [CrossRef]
  33. Ziegler, J.F.; Ziegler, M.D.; Biersack, J.P. SRIM–The stopping and range of ions in matter. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 1818–1823. [Google Scholar] [CrossRef]
  34. Ottaviano, L.; Verna, A.; Grossi, V.; Parisse, P.; Piperno, S.; Passacantando, M.; Impellizzeri, G.; Priolo, F. Surface Morphology of Mn+ Implanted Ge(100): A Systematic Investigation as a Function of the Implantation Substrate Temperature. Surf. Sci. 2007, 601, 2623–2627. [Google Scholar] [CrossRef]
  35. Stepanov, A.L.; Nuzhdin, V.I.; Valeev, V.F.; Konovalov, D.A.; Rogov, A.M. Modification of the Surface of Germanium and the Formation of a Porous Layer During Implantation with Indium Ions. Nano Biotechnol. Rep. 2023, 18, S144–S150. [Google Scholar] [CrossRef]
  36. Janssens, T.; Huyghebaert, C.; Vanhaeren, D.; Winderickx, G.; Satta, A.; Meuris, M.; Vandervorst, W. Heavy Ion Implantation in Ge: Dramatic Radiation Induced Morphology in Ge. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. Process. Meas. Phenom. 2006, 24, 510–514. [Google Scholar] [CrossRef]
  37. Romano, L.; Impellizzeri, G.; Bosco, L.; Ruffino, F.; Miritello, M.; Grimaldi, M.G. Nanoporosity Induced by Ion Implantation in Deposited Amorphous Ge Thin Films. J. Appl. Phys. 2012, 111, 113515. [Google Scholar] [CrossRef]
  38. Darby, B.L.; Yates, B.R.; Rudawski, N.G.; Jones, K.S.; Kontos, A.; Elliman, R.G. Insights for Void Formation in Ion-Implanted Ge. Thin Solid Film. 2011, 519, 5962–5965. [Google Scholar] [CrossRef]
  39. Rogov, A.M.; Nuzhdin, V.I.; Valeev, V.F.; Stepanov, A.L. Formation of Porous Germanium Layers with Various Surface Morphology in Dependence on Mass of Implanted Ions. Compos. Commun. 2020, 19, 6–10. [Google Scholar] [CrossRef]
  40. Romano, L.; Impellizzeri, G.; Tomasello, M.V.; Giannazzo, F.; Spinella, C.; Grimaldi, M.G. Nanostructuring in Ge by Self-Ion Implantation. J. Appl. Phys. 2010, 107, 084314. [Google Scholar] [CrossRef]
  41. Rudawski, N.G.; Jones, K.S. Nanostructured Germanium Prepared via Ion Beam Modification. J. Mater. Res. 2013, 28, 1633–1645. [Google Scholar] [CrossRef]
  42. Mollick, S.A.; Kumar, M.; Singh, R.; Satpati, B.; Ghose, D.; Som, T. Gold-decorated highly ordered self-organized grating-like nanostructures on Ge surface: Kelvin probe force microscopy and conductive atomic force microscopy studies. Nanotechnology 2016, 27, 435302. [Google Scholar] [CrossRef] [PubMed]
  43. Gao, K.; Prucnal, S.; Huebner, R.; Baehtz, C.; Skorupa, I.; Wang, Y.; Skorupa, W.; Helm, M.; Zhou, S. Ge1−xSnx Alloys Synthesized by Ion Implantation and Pulsed Laser Melting. Appl. Phys. Lett. 2014, 105, 042107. [Google Scholar] [CrossRef]
  44. Tran, T.T.; Alkhaldi, H.S.; Gandhi, H.H.; Pastor, D.; Huston, L.Q.; Wong-Leung, J.; Aziz, M.J.; Williams, J.S. Suppression of Ion-Implantation Induced Porosity in Germanium by a Silicon Dioxide Capping Layer. Appl. Phys. Lett. 2016, 109, 082106. [Google Scholar] [CrossRef]
  45. Tran, T.T.; Pastor, D.; Gandhi, H.H.; Smillie, L.A.; Akey, A.J.; Aziz, M.J.; Williams, J.S. Synthesis of Ge1−xSnx Alloys by Ion Implantation and Pulsed Laser Melting: Towards a Group IV Direct Bandgap Material. J. Appl. Phys. 2016, 119, 183102. [Google Scholar] [CrossRef]
  46. Prucnal, S.; Liu, F.; Voelskow, M.; Vines, L.; Rebohle, L.; Lang, D.; Berencén, Y.; Andric, S.; Boettger, R.; Helm, M.; et al. Ultra-Doped n-Type Germanium Thin Films for Sensing in the Mid-Infrared. Sci. Rep. 2016, 6, 27643. [Google Scholar] [CrossRef]
  47. Leem, J.W.; Song, Y.M.; Yu, J.S. Broadband Antireflective Germanium Surfaces Based on Subwavelength Structures for Photovoltaic Cell Applications. Opt. Express 2011, 19, 26308. [Google Scholar] [CrossRef]
  48. Iyengar, V.V.; Nayak, B.K.; Gupta, M.C. Optical Properties of Silicon Light Trapping Structures for Photovoltaics. Sol. Energy Mater. Sol. Cells 2010, 94, 2251–2257. [Google Scholar] [CrossRef]
  49. Battaglia, C.; Hsu, C.-M.; Söderström, K.; Escarré, J.; Haug, F.-J.; Charrière, M.; Boccard, M.; Despeisse, M.; Alexander, D.T.L.; Cantoni, M.; et al. Light Trapping in Solar Cells: Can Periodic Beat Random? ACS Nano 2012, 6, 2790–2797. [Google Scholar] [CrossRef]
  50. Han, K.; Chang, C.-H. Numerical Modeling of Sub-Wavelength Anti-Reflective Structures for Solar Module Applications. Nanomaterials 2014, 4, 87–128. [Google Scholar] [CrossRef]
  51. Choi, H.-J.; Huh, D.; Jun, J.; Lee, H. A Review on the Fabrication and Applications of Sub-Wavelength Anti-Reflective Surfaces Based on Biomimetics. Appl. Spectrosc. Rev. 2019, 54, 719–735. [Google Scholar] [CrossRef]
  52. Gonzalez, F.L.; Gordon, M.J. Bio-Inspired, Sub-Wavelength Surface Structures for Ultra-Broadband, Omni-Directional Anti-Reflection in the Mid and Far IR. Opt. Express 2014, 22, 12808. [Google Scholar] [CrossRef] [PubMed]
  53. Ting, C.-J.; Chen, C.-F.; Chou, C.P. Antireflection Subwavelength Structures Analyzed by Using the Finite Difference Time Domain Method. Optik 2009, 120, 814–817. [Google Scholar] [CrossRef]
  54. An, S.; Liao, Y.; Shin, S.; Kim, M. Black Germanium Photodetector Exceeds External Quantum Efficiency of 160%. Adv. Mater. Technol. 2022, 7, 2100912. [Google Scholar] [CrossRef]
  55. Zanatta, A.R. Revisiting the Optical Bandgap of Semiconductors and the Proposal of a Unified Methodology to Its Determination. Sci. Rep. 2019, 9, 11225. [Google Scholar] [CrossRef]
  56. Kittler, M.; Arguirov, T.; Oehme, M.; Yamamoto, Y.; Tillack, B.; Abrosimov, N.V. Photoluminescence Study of Ge Containing Crystal Defects. Phys. Status Solidi (a) 2011, 208, 754–759. [Google Scholar] [CrossRef]
  57. Lieten, R.R.; Bustillo, K.; Smets, T.; Simoen, E.; Ager, J.W.; Haller, E.E.; Locquet, J.-P. Photoluminescence of Bulk Germanium. Phys. Rev. B 2012, 86, 035204. [Google Scholar] [CrossRef]
  58. Nataraj, L.; Xu, F.; Cloutier, S.G. Direct-Bandgap Luminescence at Room-Temperature from Highly-Strained Germanium Nanocrystals. Opt. Express 2010, 18, 7085. [Google Scholar] [CrossRef]
  59. Carolan, D. Recent Advances in Germanium Nanocrystals: Synthesis, Optical Properties and Applications. Prog. Mater. Sci. 2017, 90, 128–158. [Google Scholar] [CrossRef]
  60. Steinmann, E.A. Dislocation Luminescence of Germanium. Cryst. Res. Technol. 1981, 16, 247–250. [Google Scholar] [CrossRef]
  61. Shevchenko, S.A.; Tereshchenko, A.N. Photoluminescence in Germanium with a Quasi-Equilibrium Dislocation Structure. Phys. Solid State 2007, 49, 28–33. [Google Scholar] [CrossRef]
  62. Street, R.A.; Johnson, N.M.; Gibbons, J.F. Defect Luminescence in Cw Laser-Annealed Silicon. J. Appl. Phys. 1979, 50, 8201–8203. [Google Scholar] [CrossRef]
  63. Reshchikov, M.A.; Morkoç, H. Luminescence Properties of Defects in GaN. J. Appl. Phys. 2005, 97, 061301. [Google Scholar] [CrossRef]
Figure 1. Raman measurement spectra of reference Ge-virgin sample, as-implanted (Asimpl.) GePb, and flash-lamp-annealed GePb samples.
Figure 1. Raman measurement spectra of reference Ge-virgin sample, as-implanted (Asimpl.) GePb, and flash-lamp-annealed GePb samples.
Materials 18 02258 g001
Figure 2. The results of RBS measurement along the random direction of reference Ge-virgin, as-implanted GePb, and flash-lamp-annealed GePb samples, while SiO2-capped layers of all of the samples are removed before RBS measurement.
Figure 2. The results of RBS measurement along the random direction of reference Ge-virgin, as-implanted GePb, and flash-lamp-annealed GePb samples, while SiO2-capped layers of all of the samples are removed before RBS measurement.
Materials 18 02258 g002
Figure 3. (a,b) Overview bright-field TEM images and (c,d) magnified views of as-implanted GePb (a,c) and flash-lamp-annealed GePb (b,d). The featureless columns in the image indicate the cavity structure, while the light gray and dark gray areas represent the amorphous GePb (a-GePb) and crystalline GePb (c-GePb), respectively. It should be noted that the contrast variations within and below the cavity structure are not due to the Pb implantation but, rather, are caused by different thinning rates during focused-ion-beam-based preparation of the TEM lamellae. (e) High-resolution TEM image (left) for the rectangular region marked in panel (d) together with the fast Fourier transforms (right) for the two quadratic regions. (f) SEM image showing the surface morphology of the flash-lamp-annealed GePb sample. The SiO2 capping layer was removed before SEM measurement.
Figure 3. (a,b) Overview bright-field TEM images and (c,d) magnified views of as-implanted GePb (a,c) and flash-lamp-annealed GePb (b,d). The featureless columns in the image indicate the cavity structure, while the light gray and dark gray areas represent the amorphous GePb (a-GePb) and crystalline GePb (c-GePb), respectively. It should be noted that the contrast variations within and below the cavity structure are not due to the Pb implantation but, rather, are caused by different thinning rates during focused-ion-beam-based preparation of the TEM lamellae. (e) High-resolution TEM image (left) for the rectangular region marked in panel (d) together with the fast Fourier transforms (right) for the two quadratic regions. (f) SEM image showing the surface morphology of the flash-lamp-annealed GePb sample. The SiO2 capping layer was removed before SEM measurement.
Materials 18 02258 g003
Figure 4. (a) Image of Ge-virgin substrate (left) and 200 keV 2 × 1015/cm2 Pb-implanted sample (right). After implantation, the GePb sample exhibited a dark color visible to the naked eye, indicating broadband antireflective properties. (b) Reflectance spectra of Ge-virgin, as-implanted GePb, and flash-lamp-annealed GePb samples. The direct bandgap and indirect bandgap of Ge are shown as EgΓ and EgL. The decreased reflectance in the GePb samples indicates broadband antireflective properties due to the cavity structures.
Figure 4. (a) Image of Ge-virgin substrate (left) and 200 keV 2 × 1015/cm2 Pb-implanted sample (right). After implantation, the GePb sample exhibited a dark color visible to the naked eye, indicating broadband antireflective properties. (b) Reflectance spectra of Ge-virgin, as-implanted GePb, and flash-lamp-annealed GePb samples. The direct bandgap and indirect bandgap of Ge are shown as EgΓ and EgL. The decreased reflectance in the GePb samples indicates broadband antireflective properties due to the cavity structures.
Materials 18 02258 g004
Figure 5. Normalized PL spectrum comparison of Ge-virgin and GePb-FLA samples at room temperature (a) and 77 K (b). The spectra show that the PL signal of the GePb-FLA sample has contributions from both the GePb layer and the Ge substrate.
Figure 5. Normalized PL spectrum comparison of Ge-virgin and GePb-FLA samples at room temperature (a) and 77 K (b). The spectra show that the PL signal of the GePb-FLA sample has contributions from both the GePb layer and the Ge substrate.
Materials 18 02258 g005
Figure 6. (a) Photoluminescence measurement of GePb-FLA sample at various temperatures. (b) The detailed PL comparison of the GePb sample at 123 K and 77 K.
Figure 6. (a) Photoluminescence measurement of GePb-FLA sample at various temperatures. (b) The detailed PL comparison of the GePb sample at 123 K and 77 K.
Materials 18 02258 g006
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

Wen, S.; Zhu, Y.-H.; Steuer, O.; Shaikh, M.S.; Prucnal, S.; Hübner, R.; Worbs, A.; He, L.; Helm, M.; Zhou, S.; et al. Optical Properties of GePb Alloy Realized by Ion Beam Technology. Materials 2025, 18, 2258. https://doi.org/10.3390/ma18102258

AMA Style

Wen S, Zhu Y-H, Steuer O, Shaikh MS, Prucnal S, Hübner R, Worbs A, He L, Helm M, Zhou S, et al. Optical Properties of GePb Alloy Realized by Ion Beam Technology. Materials. 2025; 18(10):2258. https://doi.org/10.3390/ma18102258

Chicago/Turabian Style

Wen, Shuyu, Yuan-Hao Zhu, Oliver Steuer, Mohd Saif Shaikh, Slawomir Prucnal, René Hübner, Andreas Worbs, Li He, Manfred Helm, Shengqiang Zhou, and et al. 2025. "Optical Properties of GePb Alloy Realized by Ion Beam Technology" Materials 18, no. 10: 2258. https://doi.org/10.3390/ma18102258

APA Style

Wen, S., Zhu, Y.-H., Steuer, O., Shaikh, M. S., Prucnal, S., Hübner, R., Worbs, A., He, L., Helm, M., Zhou, S., Luo, J.-W., & Berencén, Y. (2025). Optical Properties of GePb Alloy Realized by Ion Beam Technology. Materials, 18(10), 2258. https://doi.org/10.3390/ma18102258

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