A New Look at Calcium Digermanide CaGe2: A High-Performing Semimetal Transparent Conducting Material for Ge Optoelectronics
Abstract
Following a recently manifested guide of how to team up infrared transparency and high electrical conductivity within semimetal materials [C. Cui Prog. Mater. Sci. 2023, 136, 101112], we evaluate an applicability of the calcium digermanide (CaGe2) thin film electrodes for the advanced Ge-based optical devices. Rigorous growth experiments were conducted to define the optimal annealing treatment and thickness of the Ca-Ge mixture for producing stable CaGe2 layers with high figure of merit (FOM) as transparent conducting material. Ab-initio electronic band structure calculations and optical modeling confirmed CaGe2 semimetal nature, which is responsible for a demonstrated high FOM. To test CaGe2 electrodes under actual conditions, a planar Ge photodetector (PD) with metal-semiconductor-metal structure was fabricated, where CaGe2/Ge interface acts as Schottky barrier. The resulting Ge PD with semimetal electrodes outperformed commercially available Ge devices in terms of both photoresponse magnitude and operated spectral range. Moreover, by using femtosecond-laser projection lithography, a mesh CaGe2 electrode with the relative broadband transmittance of 90% and sheet resistance of 20 /sq. was demonstrated, which further enhanced Ge PD photoresponse. Thus, obtained results suggest that CaGe2 thin films have a great potential in numerous applications promoting the era of advanced Ge optoelectronics.
I Introduction
The modern civil, scientific and military applications raise new grand challenges for photonic and optoelectronic devices beyond scalability, low energy consumption and high photoelectric conversion yield, which are axiomatic golden standards. Among them flexible sensors, multicolor photodetectors (PDs) and their counterpart niche of the selectively blinds ones are worthy of special mention Xie and Yan (2017); Cao et al. (2020); Xie et al. (2019). These application fields are responsible not only for a rapid discovery of the novel light-sensitive materials at the cutting edge of the two-dimensional and topological condensed matter Wu and Gong (2021); Yan et al. (2018); Ezhilmaran et al. (2021); Huo et al. (2022); Yao et al. (2017); Zhang et al. (2022a); Liu et al. (2020); Wang et al. (2017), but for an associated breakthrough of the transparent conducting materials (TCMs) Moreira et al. (2022); Singh et al. (2022); Stoner et al. (2019); Liu and Alshareef (2021), which are essential parts of any optoelectronic device.
Transparent conducting oxides (TCOs) no longer could meet all mentioned requirements, despite a comprehensive list of the available materials Spencer et al. (2022); Jaffray et al. (2022) to replace the most utilized indium-tin-oxide with its high cost and brittleness He and Tjong (2016). In addition, widely used band engineering and doping close to the solubility limit Cai and Wei (2021) hardly can result in simultaneous low optical losses and high electrical conductivity due to free-carriers absorption and impurity-enhanced electron scattering, respectively Zhang et al. (2016). Owing to thickness reduction down to tens of nanometers, further accompanied by nanostructuring of all sorts, conventional thin metal films were granted a second chance Zhang et al. (2022b); Paeng et al. (2015); Zhang et al. (2021); Qiao et al. (2023); Nam et al. (2019). Unfortunately, there is a significant conductivity drop partially associated with island-like growth, which assumes needs for additional seeding layers deposition to improve electrical properties Park et al. (2022); Martinez-Cercos et al. (2021), nothing to say about electrical losses induced by different metal film processing toward enhanced optical transparency resulted from defects introducing Zhao et al. (2015). Quite different approach was introduced by Zhang (2016) with original materials screening based on high carrier effective mass rather than high concentration Zhang et al. (2016). This concept allows plasma energy to be shifted deeply below visible range with much lower free-carriers absorption compared to conventional metals. The systems of interests became correlated metals and semimetals Ok et al. (2021); Boileau et al. (2022), which infrared transparency has been recently validated for the thin films of calcium disilicide (CaSi2) and tungsten ditelluride (WTe2) both possessing layered crystal structure and classified as trivial and type-II Weyl semimetals, respectively Shevlyagin et al. (2022a); Cui et al. (2023). High intrinsic transparency in the optical telecommunication spectral range of the former material was shown in addition to a very low sheet resistance, which resulted in a record near-infrared (NIR) figure of merit (FOM) competitive with state-of-the-art TCOs and other TCMs. Moreover, a prototype photovoltaic device utilizing CaSi2 top electrode instead of conventional metal-finger contacts was demonstrated resulting in the enhanced photovoltaic performance and clearly confirmed reliability of the semimetal approach Shevlyagin et al. (2022b).
It is a common knowledge that Ge is a more attractive NIR optoelectronics platform in replacing Si owing to its almost twice-narrower band gap Miller (2009). That is why, developing the TCMs for that purpose is of high interest. Unfortunately, there are no reports on Ge integration with semimetals except for PDs with transparent electrodes made of graphene, Kwon et al. (2022); Jiang et al. (2022) which is semimetal to a certain extent. However, as it is often the case for Me-IV alloys and compounds (Me = Ca, Mg; IV = Si, Ge, Sn), the solution lies on the surface. Thus, we propose to use calcium digermanide (CaGe2) films in a similar way to CaSi2 and Si. Despite the same crystallography (Zintl phases) Beekman et al. (2019), CaGe2 is less investigated in comparison with CaSi2. Little is known about its electrical properties Evers and Weiss (1974) with no available data on optical investigations. Currently, CaGe2 is used to produce 2D derivatives of Ge (germanene, germanane, polygermine etc.) by topochemical reaction Jiang et al. (2016); Vogg et al. (2000a); Rosli et al. (2020); Liu et al. (2019); Chia et al. (2021) just like in the case of CaSi2 and silicene, which is Si-based graphene analogue Wang et al. (2022); Ryan et al. (2019).
In this work, we report on successful growth of the CaGe2 films on the transparent insulating (Al2O3) and semiconducting (Ge) substrates. Ambient stability of the produced films depending on growth conditions was assessed by means of Raman spectroscopy and X-ray diffraction (XRD) methods. Based on electrical and optical measurements accompanied by first principles electronic band structure calculations and optical modeling, CaGe2 was comprehensively characterized as a candidate for TCMs. It turned out, that among CaGe2 polymorph modifications Yaokawa et al. (2021, 2018), only h2-CaGe2 is stable under ambient conditions, while hR6-CaGe2 tends to be formed as a primary phase in thin and/or annealed at low temperature Ca-Ge films. Next, it was shown that h2-CaGe2 possesses semimetal behavior, which determines its high IR transparency and high conductivity similar to CaSi2. After growth conditions optimization, CaGe2 film demonstrates maximal optical transmittance close to 80% at 2.25 m wavelength, while its sheet resistance is as low as 13 /sq. In view of these features, CaGe2 film in the form of finger electrodes was grown on Ge(001) substrate to produce Ge metal-semiconductor-metal (MSM) PD with back-to-back two CaGe2/Ge Schottky barriers. The resulting Ge MSM PD demonstrates expanded photoresponse spectrum down to 2300 nm and enhanced sensitivity of 0.8 A/W under a small reverse bias compared with conventional Ge PD. Finally, to compensate for low transparency at optical frequencies, femtosecond (fs) laser perforation was used to produce a CaGe2 mesh electrode. Even the highest perforation ratio of 92% has only moderate influence on the resultant sheet resistance, while optical transparency in the (400-1000) nm range was enhanced by a factor of 3. As a result, CaGe2 electrode could reach FOM of 0.75 that much higher or comparable with the currently applied near-IR and middle-IR TCMs Gao et al. (2021); Khamh et al. (2018); Tong et al. (2016); Wang et al. (2019); Fukumoto et al. (2022) and Ge PD with perforated CaGe2 top electrode surpassed peak value photoresponse of 1 A/W at 1600 nm wavelength while overall improvement in photosensitivity for a wide photon band (400-2200) nm was confirmed. Of great importance is that either the semimetal nature of CaGe2 or laser perforation have no crucial influence on Ge PD resulting response speed (tens of microseconds), which is comparable with other Ge-based heterojunction devices.
II Results and Discussion
II.1 Influence of growth conditions on ambient stability of the Ca-Ge layers

We started our examinations from the attempts to grow CaGe2 films on Al2O3 (sapphire) substrate. In doing so, a solid phase epitaxy (SPE) method was chosen suggesting some similarities between CaGe2 thin films and CaSi2 transparent conducting layers obtained previously Shevlyagin et al. (2022a). Schematic illustration of the CaGe2 SPE, which is a two-stage process, is pictured in Figure 1a. A set of the samples was obtained with varied Ca-Ge thickness (10-120) nm and annealing temperature (600-850)oC. All grown films can be divided into two groups based on their stability, which is directly associated with phase composition driven by the growth conditions. Stable after air exposure and storage for 1 year samples demonstrate simultaneous partial optical transparency at least in the red spectral region (Figure 1d), sufficient electrical conductivity to supply the LED by passing a DC current through it (Figure 1e) and relatively smooth surface with RMS roughness not exceeding 3 nm in accordance with atomic force microscopy (AFM) data (Figure 1b). The latter is of great importance for developing TCMs with high FOM values, since pronounced surface relief could deteriorate both transparency and electrical conductivity, to say nothing of difficulties with materials processing towards real electrode engineering (lithography, patterning etc.). Scanning electron microscopy (SEM) investigations show that stable Ca-Ge layers consist of large grains and form continuous films (Figure 1c), while energy dispersive X-ray spectroscopy (EDX) confirms stoichiometry of CaGe2 (Figure S1 in Supporting information). For unstable CaGe2 films, in most cases their degradation took place within a few hours depending on its thickness except for the thinnest ones exhibiting some changes in their appearance obvious by a naked eye just after growth ending. The readers are further addressed to Supporting information for optical microscopy data and planar SEM images (Figure S2), which are helpful to express and preliminary examine the CaGe2 films stability, while rigorous phase identification for the grown Ca-Ge layers probed with XRD and Raman method will be given below.
It is known that CaGe2 can exist in several polymorph modifications, with hexagonal (space group #186), two trigonal (space groups #166 and #164) and monoclinic (space group #12) crystal lattices being the most frequently observed Yaokawa et al. (2018, 2021); Tobash and Bobev (2007). The former two phases referred to as hR6 (or 6R) and h2 (or 2H) were experimentally observed including thin epitaxial films, while the latter two are metastable. In addition, it was reported that fluorine diffusion and crystal lattice stress release can promote stabilization of the other hexagonal (h4 or 4H) and trigonal (hR3 or 3R) CaGe2 polymorphs. In this view, the Ca-Ge phase diagram is quite similar to the Ca-Si one in terms of the variety of phases and compounds. However, while both h3R and h6R trigonal CaSi2 polymorphs are (i) stable, (ii) semimetal in nature and (iii) tolerant to degradation under ambient conditions, it is not the case for CaGe2. Additional challenges arise, since h2- and hR6-CaGe2 can be hardly separated by means of Raman spectroscopy and XRD examination Vogg et al. (2000b); Hara et al. (2021); Arguilla et al. (2017). For example, Figure 1h represents a typical XRD pattern of the decomposed film obtained. The observed peaks can be attributed as diffraction from CaGe2 and CaO crystal planes. However, it was further specified that GeH (germanane) and CaO formation took place rather than both hR6-CaGe2 decomposition and h2-CaGe2 preservation. The trick is that GeH and h2-CaGe2 diffraction peaks are hardly distinguishable Itoh et al. (2022); Pinchuk et al. (2014), which can be clearly seen by comparing the two XRD patterns corresponding to stable and decomposed Ge-Ca layers (Figure 1h). The only difference is in the observed CaO related peak in the latter case. The presence of the GeH in the decomposed film instead of the h2 phase is supported by Raman measurements presented in Figure 1g. The broad phonon bands centered at 275 cm-1 (Ge-Ge bonds) and 225 cm-1 (Ge-H bonds) can be categorically assigned to GeH (the pink graph) Bianco et al. (2013). This spectrum is in a marked contrast with that of measured from the stable CaGe2 film (the green graph). At least four relatively narrow peaks were resolved, assigned as Eg and A1g phonon modes of h2-CaGe2. Thus, XRD measurements allowed detecting CaO presence, while Raman spectroscopy did the same for GeH. As a result, hR6-CaGe2+H2OCaO+2GeH can be tentatively suggested as an origin of the observed decomposition of thin CaGe2 layers after air exposure. Concerning growth conditions, which lead to CaGe2 decomposition, it can be stated that (i) thin (15 nm) films show this tendency regardless annealing temperature, (ii) increase in total Ge-Ca thickness results in the presence of both h2 and hR6 phases and (iii) further increase in thickness opens a stability window for CaGe2 in h2 modification with no hR6 traces, which suggest single-phase film growth. Too low annealing temperature results in CaGe formation (insulating phase), while too high temperature treatment does in re-evaporation of the deposited layers.
To outline some intermediate results on experiments of growth stable CaGe2 films, we illustrated it by a schematic phase diagram shown in Figure 1f, which reflects chemical composition of the obtained samples addressed by Raman and XRD measurements in relation to growth conditions. In brief, it appears to be that a stable and single-phase h2-CaGe2 film could not be grown thinner than 20 nm and below annealing temperature of 750oC (region I). Outside the optimal ranges, the following processes make sense: phase coexistence (II), monogermanide formation (region III), total decomposition (region IV) and film re-evaporation (region V).
II.2 Semimetal nature of the CaGe2 and its influence on the obtained TCM figure of merit

After resolving the stability issue, optical and electrical properties of the h2-CaGe2 films can be discussed from the view of both modeling and some practical aspects toward the TCMs performance. This screening was performed by first principles calculations, while extracted information is of high value in supporting optical and electrical measurements. We started for the band structure calculations of the fully relaxed h2-CaGe2 primitive cell and the resulting electron band diagram is shown in Figure 2a. There are multiple crossings at the Fermi level for both valence and electron bands. One can see at least one electron and one hole pockets above Fermi level (holes energy is counted inversely to that of for electrons) tinged with red and green, respectively. In addition, some peculiarities of the h2-CaGe2 band structure are of great importance. First, CaGe2 in h2 modification is a trivial indirect-type semimetal Markov et al. (2019), which means that electron and valence band extreme points located at the different k-points ( and M, respectively) with the so-called negative band gap of about 0.6 eV (the sum of the electrons and holes pockets maximal energy offset with respect to Fermi level). Secondly, one can observe that with the exception of mentioned carriers pockets, there is a wide energy gap diving below and above Fermi level to about 0.6 and 0.3 eV, respectively. It resulted in a very low electron density of states (DOS) near Fermi levels compared to pure metals, but significantly larger than for semiconductors even under degeneracy doping. On the one hand, this electronic bands configuration allows demonstrating very low resistivity in the order of 10-5-10-4 cm for h2-CaGe2 thin films (see Supporting information S3 for Hall measurements performed on the films of different thickness) compared with bulk single-crystals Evers and Weiss (1974). From the other hand, low electronic DOS should results in the low optical DOS as well. Calculated complex dielectric function of the h2-CaGe2 phase is plotted in Figure 2b and explains well the low optical losses in the IR spectral range up to 0.8 eV.
To provide an experimental verification of the ab initio calculations, optical transparency and sheet resistance were measured for 100-nm thick CaGe2 film on Al2O3 substrate obtained under the growth conditions, which correspond to formation of a stable Ca-Ge phase (region I; Figure 1a). A typical experimental transmittance spectrum of the CaGe2/Al2O3 sample is presented in Figure 2c together with the optical modeling results based on calculated complex dielectric function of the h2-CaGe2 and the data available for Al2O3. In doing this, obtained dielectric constants within Fresnel law were used to calculate optical response of the CaGe2/Al2O3 system assuming a smooth and sharp interface between materials and zero scattering. One can see a close agreement between theoretical and experimental evaluations. In addition, relatively thick (100 nm) CaGe2 is semitransparent from the near-infrared (NIR) to middle-infrared (MIR) spectral ranges, while optical transparency of the CaGe2/Al2O3 system reaches 60% at 2500 nm.

Next, rigorous investigations were performed to define CaGe2 growth conditions (thickness and annealing temperature), which lead to the highest FOM value in accordance with the following expression: FOM = -1/[Rln(T)], where Rsheet and T are sheet resistance and optical transmittance (at the selected wavelength or averaged over the specific range), respectively Gordon (2000). The results of this screening is shown as 3D mapping in Figure 2d. It was found that the transparency window of the h2-CaGe2 itself is located in the near-IR (1-3 m) spectral range. The measured sheet resistance varied from 8 to 30 /sq. depending on both film thickness and annealing conditions. After all, 60 nm thick CaGe2 film annealed at 750oC demonstrated the highest relative optical transparency of 78% at 2.25 m and moderate sheet resistance of 16 /sq. These parameters give the FOM values of the 0.33 and 0.13 at the wavelength of maximal transmittance (2250 nm) and for an transparency averaged over (400-7000) nm range, respectively. Obtained FOM value for the h2-CaGe2 is competitive with other state-of-the-art TCMs currently applied for NIR-MIR applications Gao et al. (2021); Khamh et al. (2018); Tong et al. (2016); Wang et al. (2019); Fukumoto et al. (2022).
II.3 CaGe2/Ge planar MSM NIR-SWIR photodetector: fabrication and characterization
To test h2-CaGe2 TCM under the real photovoltaic device operations, we grew a corresponding thin film on the Ge(111) substrate instead of Al2O3 transparent one using the same optimized conditions found previously (60 nm thick CaGe2 layer annealed at 750oC). The deposition and annealing of the Ge-Ca bilayer was done through a tantalum mask for in situ formation of the two CaGe2 pads. Thus, we obtained a MSM photodetector acting as two back-to-back Schottky junctions with a schematic design pictured in Figure 3a. Raman measurements suggest a good crystallinity of the grown semimetal layers with observed phonon bands corresponding only to CaGe2 and Ge with no traces of film decomposition and other Ca-Ge alloys. Figure 3b demonstrates dark I-V characteristic of the planar MSM structure with clear Schottky behavior. Under the assumption of the thermionic-field emission regime Katz et al. (2001), a Schottky barrier height was calculated to be 0.5 eV, while reverse saturation current is equal to 50 nA. Under laser illumination at 1550 nm, CaGe2/Ge/CaGe2 MSM diode demonstrates no photocurrent generation under zero-bias condition. However, even a small bias voltage results in pronounced photocurrent, wherein current flow is enhanced by two orders of magnitude at the applied 2V under illumination compared to dark conditions. Thus, a typical operation of the MSM PD with symmetrical Schottky contacts was confirmed. Next, a room temperature photoresponse at the bias voltage of 2V was measured for the CaGe2/Ge/CaGe2 MSM PD. Results obtained are plotted in Figure 3c together with characteristics of the commercial Ge PD. It is obvious that application of the CaGe2 transparent electrodes resulted in the spectral range expanded into the shortwave IR region below the band gap of Ge (0.67 eV). We estimated a photoelectric threshold energy to be 0.54 eV (2300 nm), which well correlates with Schottky barrier height obtained from the I-V curve. Moreover, there is the photoresponse enhancement in the (1200-1900) nm range owing to a high transparency of the CaGe2 electrodes, which result in higher photogeneration occurring in Ge. Thus, a simple but effective approach of advancing Ge-based photodiode performance in terms of maximal sensitivity and red shifting of the cut-off wavelength is demonstrated.
II.4 Improvement of the CaGe2 TCM performance with laser projection lithography

Despite a high NIR-MIR TCM performance, the grown CaGe2 films have low optical transparency in the visible spectral range. Fortunately, it can be tuned by making a perforated mesh electrode by direct fs-laser patterning. For this purpose, 60 nm thick CaGe2 film was processed using a flat-top square-shape laser beam as schematically illustrated in Figure 4a. The laser fluence was kept slightly above the single-pulse ablation threshold of the corresponding CaGe2 film (see detailed description of the ablation threshold measurements in Supporting information - Figure S4). Typical morphology of the 60-nm thick CaGe2 film grown on Al2O3 substrate after laser processing is demonstrated by top-view SEM image in Figure 4b, with the photograph of the resulted Al2O3/CaGe2 sample after laser perforation clearly demonstrating the enhanced optical transparency compared to non-perforated areas (Figure 4c). The sidewalls of the CaGe2 mesh are smooth and no traces of the ejected submicron-sized particles can be found, which suggest high suitability of the CaGe2 material for laser patterning. It is important to note, that a maximal rim height of the square-shape openings was about 90 nm (Figure S5), which is only 1.7 higher than initial CaGe2 film thickness indicating that under chosen laser patterning conditions the outward radial mass transfer of the molten Ca-Ge is not much intense.
Representative optical transparency spectra of the 60-nm thick CaGe2 film before and after laser perforation are presented in Figure 4d. The most pronounced changes are observed in the (400-1500) nm range, which are characterized by 3 times higher transparency of the perforated sample in comparison to untreated one, while transparency in the (1500-7000) nm range are intrinsically high for CaGe2 material owing to its semimetal nature. Raman measurements confirmed (not shown) that even at high perforation ratio, laser ablation does not strongly affect crystallinity of the remaining sections of the CaGe2 films. Compared with continuous films, a mesh CaGe2 electrode with an optimal perforation ratio (determined to be 78%) demonstrates six-fold increases in the FOM value reaching 0.75 (see Figure 4e) due to an optical transparency averaged over (400-7000) nm beating 90% after laser processing. The lower perforation ratio is not enough to reach high optical transparency. On the contrary, extensive perforation rate above 80% decreases FOM value owing to drastically increased sheet resistance up to 32 /sq., which is 3 times higher compared with initial smooth CaGe2 film, however being still acceptable for TCM applications.
Finally, to confirm and highlight the advantage of fs-laser projection lithography in tuning optical properties of CaGe2 film, a vertical Ge PD was fabricated on a sapphire substrate with a perforated transparent electrode (face contact) and thin Al film acting as a back electrode. The schematic view of the resulting vertical Ge Schottky PD is shown in Figure 4f assuming light irradiation from the CaGe2 side. As the most illustrative test, we next directly compared the photoresponse characteristics of the two Ge PDs with a flat and perforated CaGe2 electrodes (Figure 4g). One can see that the latter device generally follows the trend observed for the optical transmittance measurement. For instance, integrated over (600-2000) nm range photoresponse and peak photoresponse value (at 1600 nm) were enhanced by 85% and 44% (the room-temperature photoresponse exceeds 1 A/W under -2V bias voltage), respectively, while an expanded operation spectral range was observed for the Ge device with perforated top electrode. Of great importance, integrating the Ge PD platform with CaGe2 electrodes has no negative influence on the light detection, which is linearly dependent from the irradiation intensity regardless of CaGe2 electrode type (see inset in Figure 4g). In addition, we investigated the response speed of the device with a perforated electrode. The relative balance of the Ge PD device versus a switching frequency of pulsed IR light irradiation (@1550 nm laser) is presented in Figure 4h with a relatively high 3dB frequency of 22 kHz, which claims for Ge PD with transparent conducting electrode made from semimetal film to be capable of detecting fast optical signals. The response speed was further evaluated by analyzing the rising and falling edges of the photoresponse time curve under a cycle conditions, which are normally calculated as the time intervals for the response to rise from 10% to 90% and vice versa (Figure 4i). A fast rise/fall time (/) of 27/47 s was obtained. We could attribute this relatively quick response speed to the low density of trap centers at the Ge/CaGe2 heterointerface, which is additionally confirmed by the photoresponse versus incident light dependence, and a very effective separation of photogenerated carriers by the built-in electric field formed at the CaGe2/Ge Schottky junction area. To sum up, the critical parameters for the developed vertical Ge PD with Schottky type contact made of perforated film of semimetalic CaGe2 were compared with other Ge Schottky type PDs Kwon et al. (2022); Jiang et al. (2022); Falcone et al. (2022); Huang et al. (2021, 2018); Cho et al. (2016); An et al. (2022); Dushaq et al. (2017); Ang et al. (2008); Zumuukhorol et al. (2019); Yin et al. (2022); Chang et al. (2019); Zeng et al. (2013); Kim et al. (2021); Yang et al. (2017); Luo et al. (2019); Zhao et al. (2020); Xiong et al. (2022) with an emphasis on MSM structures and listed in Table 1.
Structure |
|
|
|
|
|
Reference | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
|
2 V | 0.6 - 2 m |
|
0.31 mA/cm2 |
|
This work | ||||||||||
|
2 V | 1.3 - 1.8 m |
|
2 mA/cm2 | - | [63] | ||||||||||
|
1 V |
|
|
33 mA/cm2 | - | [64] | ||||||||||
|
0.2 V | 0.8-1.65 m |
|
1.4 mA/cm2 | - | [65] | ||||||||||
|
1 V | selected wavelength |
|
100 mA/cm2 | - | [66] | ||||||||||
|
1 V | selected wavelength |
|
63 mA/cm2 | - | [35] | ||||||||||
|
0.5 V | 1.5-2 m |
|
0.616 mA | - | [67] | ||||||||||
|
1 V | selected wavelength |
|
0.76 mA/cm2 | - | [68] | ||||||||||
|
1 V | selected wavelength |
|
100 mA/cm2 | 15 GHz | [69] | ||||||||||
|
2 V | 1.53-1.61 m |
|
2 mA/cm2 | - | [70] | ||||||||||
|
1V | 1.064-1.85 m |
|
1.6 mA/cm2 | - | [36] | ||||||||||
|
zero bias | selected wavelength |
|
- | 12.1 s | [71] | ||||||||||
|
10 V | 0.3-1.8 m |
|
- | 250 Hz | [72] | ||||||||||
|
zero bias | 1.2-1.6 m |
|
- | 23 s/108 s | [73] | ||||||||||
|
2 V | 0.5-1.65 m |
|
- | - | [74] | ||||||||||
|
1 V | 0.35-1.65 m |
|
- | 5.6 ms/3.5 ms | [75] | ||||||||||
|
zero bias | selected wavelength |
|
- | 25.4 s/38.5 s | [76] | ||||||||||
|
1 V | selected wavelength |
|
- | 68 s/70 s | [77] | ||||||||||
|
zero bias | 0.35-1.55 m |
|
- | 1.4 s/4.1 s | [78] |
As can be clearly seen, such phenomena as transparent conducting electrodes have rarely been addressed concerning the applications in Ge PDs except for some works focusing on graphene, ITO and MXenes integration, nothing to say about much fewer reports on Ge/semimetal or Ge/topological insulator heterojunction PDs. The current work demonstrates that low dark current, high responsivity in wide wavelength range and fast response speed is achievable by introducing semimetal CaGe2 transparent conducting electrode into Ge optoelectronics. In particular, spectral operation range of the proposed Ge PD is the broadest than that of other photodetectors of the same type (Schottky or MSM). Concerning both photoresponsivity and operation speed, fabricated PD is obviously outperform commercially available Ge PD, while demonstrating competitive characteristics even compared to black Ge based PD. Moreover, a self-powered operation (zero bias driving voltage) is not anyhow restricted by CaGe2 application and could be potentially achievable by the widely known symmetry breaking approach Wang et al. (2020); Zhou et al. (2018) for Schottky type PDs, which could be a question of future work. Thus, obtained results suggest that laser-perforated CaGe2 thin films have a great potential as transparent conducting materials for Ge optoelectronic devices and applications.
III Conclusion
In the present work, a CaGe2 commonly used in chemical reactions toward producing 2D Ge structures and derivatives was examined as a candidate for transparent conducting material to be applied in Ge-based optoelectronics. Obtained CaGe2 films demonstrated electrical and optical properties quite similar to its vis-a-vis CaSi2. Semimetal topology of the electronic bands provided high both optical transparency and electrical conductivity, which were suggested to be the common features of the Ca-IV2 (IV=Si, Ge) Zintl phases. However, experimental investigations revealed some critical differences in stability and optical performance of the CaGe2 layers. The former is resulted in relatively narrow ranges of the growth conditions, which prevent ambient induced decomposition and promote a single-phase CaGe2 film formation. The latter is in the redshift of the transparency window to the MIR region compared to CaSi2 film of equal thickness. Nevertheless, after finding the optimal growth conditions, CaGe2 layers were used to produce planar metal-semiconductor-metal photodetector with two back-to-back Ge/CaGe2 Schottky junctions. The resulting Ge MSM PD with CaGe2 top transparent electrodes outperformed commercial Ge PD in both absolute photoresponse value operated spectral range. Next, a low intrinsic visible transparency of the CaGe2 film was improved by a film perforation with fs-laser projection lithography. A mesh electrode with perforation ratio of 80% reaches FOM value of 0.75 in the wide spectral range of (400-7000) nm, which makes it competitive with other TCMs. Finally, patterned CaGe2 thin film was used towards advanced Ge PDs developing, which demonstrated low dark current, high responsivity in wide wavelength range and fast response speed compared to state-of-the-art Ge-based optoelectronic devices.
Of great importance is that we demonstrated a new application of the CaGe2 beyond just using it as precursor material for germanane, germanene and their polymorphs and derivatives production, while results obtained could become a great step toward mass Ge optoelectronics.
IV Experimental Section
All CaGe2 thin films were grown in a turbo-pumped vacuum chamber with a base pressure of 10-6 mbar. Solid phase epitaxy was used to form CaGe2 layers in a way similar to CaSi2 growth, which can be found elsewhere Shevlyagin et al. (2022a). First, a Ge-Ca bilayer (K-cells, evaporation rates of 25 nm/min) was deposited onto substrate kept at room temperature followed by thermal annealing.
Phase composition and crystallinity of all samples were examined by Raman spectroscopy setup (NTEGRA SPECTRA II) and X-ray diffraction method (XRD) working in 2/ mode with Cu Kα radiation source, parallel beam optics (Rigaku SmartLab). In addition, visual control of the sample surface modification after air exposure was carried out with an optical microscopy.
Combination of the four-point probe setup for sheet resistance measurements (Teslatron PV) and Fourier Transform Infrared spectrometer (Bruker Vertex V80) coupled to the IR microscope (Bruker Hyperion 2000) for optical properties evaluation were used to calculate FOM values of Ca-Ge films as TCM using the following expression: FOM = -1/[Rln(T)], where Rsheet and T are sheet resistance and optical transmittance (at the selected wavelength or averaged over the specific range), respectively Gordon (2000).
For a deeper insight into semimetal electronic structure of the CaGe2 and checking its optical properties, first principles calculations and multilayer optical modeling were performed. Further particulars of the calculations are as follows. The density-functional theory (DFT) Hohenberg and Kohn (1964); Kohn and Sham (1965) calculations were performed with the package VASP Kresse and Furthmüller (1996a, b). The generalized-gradient approximation (GGA) to the exchange-correlation functional was used Perdew et al. (1996). Non-spherical contributions from the gradient corrections were included Adler (1962); Wiser (1963). The cut-off energy of 600 eV and gamma-centered k-points mesh of 20?20?8 were used. Break condition for self-consistency loop was of 10-8 eV. Convergence with respect to cut-off energy and k-points density was performed. Calculation of frequency-dependent dielectric constants is implemented in VASP code as follows: the imaginary part of dielectric tensor is determined by a summation over empty band states using the equation Gajdoš et al. (2006):
(1) |
where is the vacuum dielectric constant, is the volume, and represents the valence and conduction bands respectively, is the energy of the incident phonon, u is the vector defining the polarization of the incident electric field, ur is the momentum operator, and and are the wave functions of the conduction and valence band at the k point, respectively. The real part of dielectric tensor is obtained by the Kramers-Kronig relation:
(2) |
where P denotes the principle value. Obtained dielectric constants and Fresnel law were used to calculate reflecting and refracting layers response on radiation assuming a smooth and sharp interface between materials and zero scattering.
For the device applications, a CaGe2 layer of optimal thickness and thermal annealing was grown on Ge(001) and Al2O3(0001) substrates to fabricate Ge planar and vertical Schottky MSM photodetectors. The latter CaGe2 film was delicately perforated using a fs-laser projection lithography with second-harmonic (515 nm) 180 fs laser pulses generated by regeneratively amplified Yb:KGW system at 50 KHz maximal repetition rate. The output Gaussian laser beam was shaped to flat-top square-shape beam with the lateral size of 25x25 m2 to perform uniform laser ablation of square-shape surface area of the films. To obtain such intensity distribution in the focal plane of the dry microscope objective with a numerical aperture (NA) of 0.42, the output laser beam was expanded, while its central part with nearly uniform intensity profile was passed through the square-shape pinhole drilled in the aluminum foil by direct laser ablation Zhizhchenko et al. (2020). A 4 optical system consisting of a lens and a focusing objective was further used to project with magnification the uniform square-shape intensity profile at the output of the pinhole to the focal plane of the objective. Laser perforation was followed by p-Ge (500 nm) and Al (50 nm) films deposition to complete vertical photodetector structure.
The room-temperature current-voltage (I-V) and photoresponse spectra of the resulting planar MSM PD with CaGe2/Ge Schottky barriers were measured in a standard manner described elsewhere Shevlyagin et al. (2020). To detect the response speed of the resulted Ge PD, pulsed optical signals with varied frequencies were produced by modulation (Thorlabs optical chopper) of monochromatic light with 1550 nm wavelength (Hamamatsu Xe lamp and monochromator Solar Tii, MS3504i), while the output photocurrent was recorded with an oscilloscope (DPO2012B, Tektronix).
V Acknowledgements
The work was supported by the Russian Science Foundation under the Grant #23-49-10044 (https://rscf.ru/project/23-49-10044/). Convergence verification in k-points density and cut-off energy for ab initio calculations were performed on HPC-cluster ”Akademic V.M. Matrosov” (Irkutsk Supercomputer Center of SB RAS, https://hpc.icc.ru). Calculation of frequency-dependent dielectric constants was performed using the equipment of Shared Resource Center ”Far Eastern Computing Resource” (Shared Resource Center ”Far Eastern Computing Resource” IACP FEB RAS, https://cc.dvo.ru).
VI Conflict of Interest
The authors declare no conflict of interest.
References
- Xie and Yan (2017) C. Xie and F. Yan, Small 13, 1701822 (2017).
- Cao et al. (2020) G. Cao, F. Wang, M. Peng, X. Shao, B. Yang, W. Hu, X. Li, J. Chen, Y. Shan, P. Wu, et al., Advanced Electronic Materials 6, 1901007 (2020).
- Xie et al. (2019) C. Xie, X.-T. Lu, X.-W. Tong, Z.-X. Zhang, F.-X. Liang, L. Liang, L.-B. Luo, and Y.-C. Wu, Advanced Functional Materials 29, 1806006 (2019).
- Wu and Gong (2021) J. Z. Wu and M. Gong, Advanced Photonics Research 2, 2100015 (2021).
- Yan et al. (2018) F. Yan, Z. Wei, X. Wei, Q. Lv, W. Zhu, and K. Wang, Small Methods 2, 1700349 (2018).
- Ezhilmaran et al. (2021) B. Ezhilmaran, A. Patra, S. Benny, M. Sreelakshmi, V. Akshay, S. V. Bhat, and C. S. Rout, Journal of Materials Chemistry C 9, 6122 (2021).
- Huo et al. (2022) Z. Huo, Y. Wei, Y. Wang, Z. L. Wang, and Q. Sun, Advanced Functional Materials 32, 2206900 (2022).
- Yao et al. (2017) J. Yao, Z. Zheng, and G. Yang, Advanced Functional Materials 27, 1701823 (2017).
- Zhang et al. (2022a) L. Zhang, X. Han, P. Wen, S. Zhang, Z. Zheng, J. Li, and W. Gao, ACS Applied Nano Materials 5, 6523 (2022a).
- Liu et al. (2020) J. Liu, F. Xia, D. Xiao, F. J. Garcia de Abajo, and D. Sun, Nature materials 19, 830 (2020).
- Wang et al. (2017) Q. Wang, C.-Z. Li, S. Ge, J.-G. Li, W. Lu, J. Lai, X. Liu, J. Ma, D.-P. Yu, Z.-M. Liao, et al., Nano letters 17, 834 (2017).
- Moreira et al. (2022) M. Moreira, J. Afonso, J. Crepelliere, D. Lenoble, and P. Lunca-Popa, Journal of Materials Science 57, 3114 (2022).
- Singh et al. (2022) M. Singh, S. Rana, and A. K. Singh, Colloid and Interface Science Communications 46, 100563 (2022).
- Stoner et al. (2019) J. L. Stoner, P. A. Murgatroyd, M. O’Sullivan, M. S. Dyer, T. D. Manning, J. B. Claridge, M. J. Rosseinsky, and J. Alaria, Advanced Functional Materials 29, 1808609 (2019).
- Liu and Alshareef (2021) Z. Liu and H. N. Alshareef, Advanced electronic materials 7, 2100295 (2021).
- Spencer et al. (2022) J. A. Spencer, A. L. Mock, A. G. Jacobs, M. Schubert, Y. Zhang, and M. J. Tadjer, Applied Physics Reviews 9 (2022).
- Jaffray et al. (2022) W. Jaffray, S. Saha, V. M. Shalaev, A. Boltasseva, and M. Ferrera, Advances in Optics and Photonics 14, 148 (2022).
- He and Tjong (2016) L. He and S. C. Tjong, Materials Science and Engineering: R: Reports 109, 1 (2016).
- Cai and Wei (2021) X. Cai and S.-H. Wei, Applied Physics Letters 119 (2021).
- Zhang et al. (2016) L. Zhang, Y. Zhou, L. Guo, W. Zhao, A. Barnes, H.-T. Zhang, C. Eaton, Y. Zheng, M. Brahlek, H. F. Haneef, et al., Nature materials 15, 204 (2016).
- Zhang et al. (2022b) T.-R. Zhang, Y.-N. Wang, Y.-F. Liu, and J. Feng, Applied Physics Letters 120 (2022b).
- Paeng et al. (2015) D. Paeng, J.-H. Yoo, J. Yeo, D. Lee, E. Kim, S. H. Ko, and C. P. Grigoropoulos, Advanced Materials (Deerfield Beach, Fla.) 27, 2762 (2015).
- Zhang et al. (2021) Q. Zhang, D. Huang, D. Qi, W. Zhou, L. Wang, Z. Zhang, S. Chen, S. Dai, and H. Zheng, Optics & Laser Technology 140, 107056 (2021).
- Qiao et al. (2023) J. Qiao, J. Huang, L. Zhao, D. Yuan, P. Wang, and S. Xu, Advanced Materials Technologies 8, 2201209 (2023).
- Nam et al. (2019) V. B. Nam, J. Shin, Y. Yoon, T. T. Giang, J. Kwon, Y. D. Suh, J. Yeo, S. Hong, S. H. Ko, and D. Lee, Advanced Functional Materials 29, 1806895 (2019).
- Park et al. (2022) Y.-B. Park, C. Jeong, and L. J. Guo, Advanced Electronic Materials 8, 2100970 (2022).
- Martinez-Cercos et al. (2021) D. Martinez-Cercos, B. Paulillo, R. A. Maniyara, A. Rezikyan, I. Bhattacharyya, P. Mazumder, and V. Pruneri, ACS Applied Materials & Interfaces 13, 46990 (2021).
- Zhao et al. (2015) G. Zhao, W. Wang, T.-S. Bae, S.-G. Lee, C. Mun, S. Lee, H. Yu, G.-H. Lee, M. Song, and J. Yun, Nature communications 6, 8830 (2015).
- Ok et al. (2021) J. M. Ok, N. Mohanta, J. Zhang, S. Yoon, S. Okamoto, E. S. Choi, H. Zhou, M. Briggeman, P. Irvin, A. R. Lupini, et al., Science Advances 7, eabf9631 (2021).
- Boileau et al. (2022) A. Boileau, S. Hurand, F. Baudouin, U. Lüders, M. Dallocchio, B. Bérini, A. Cheikh, A. David, F. Paumier, T. Girardeau, et al., Advanced Functional Materials 32, 2108047 (2022).
- Shevlyagin et al. (2022a) A. Shevlyagin, N. Galkin, K. Galkin, E. Subbotin, V. Il yaschenko, A. Gerasimenko, and I. Tkachenko, Journal of Alloys and Compounds 910, 164893 (2022a).
- Cui et al. (2023) C. Cui, Q. Ding, S. Yu, C. Yu, D. Jiang, C. Hu, Z. Gu, and J. Zhu, Progress in Materials Science , 101112 (2023).
- Shevlyagin et al. (2022b) A. V. Shevlyagin, V. M. Il yaschenko, A. A. Kuchmizhak, E. V. Mitsai, A. V. Amosov, S. A. Balagan, and S. A. Kulinich, Materials 15, 6637 (2022b).
- Miller (2009) D. A. Miller, Proceedings of the IEEE 97, 1166 (2009).
- Kwon et al. (2022) M. G. Kwon, C. Kim, K. E. Chang, T. J. Yoo, S.-Y. Kim, H. J. Hwang, S. Lee, and B. H. Lee, APL Photonics 7 (2022).
- Jiang et al. (2022) H. Jiang, B. Li, Y. Wei, S. Feng, Z. Di, Z. Xue, D. Sun, and C. Liu, Nanotechnology 33, 345204 (2022).
- Beekman et al. (2019) M. Beekman, S. M. Kauzlarich, L. Doherty, and G. S. Nolas, Materials 12, 1139 (2019).
- Evers and Weiss (1974) J. Evers and A. Weiss, Materials Research Bulletin 9, 549 (1974).
- Jiang et al. (2016) S. Jiang, M. Q. Arguilla, N. D. Cultrara, and J. E. Goldberger, Chemistry of Materials 28, 4735 (2016).
- Vogg et al. (2000a) G. Vogg, M. S. Brandt, and M. Stutzmann, Advanced materials 12, 1278 (2000a).
- Rosli et al. (2020) N. F. Rosli, N. Rohaizad, J. Sturala, A. C. Fisher, R. D. Webster, and M. Pumera, Advanced Functional Materials 30, 1910186 (2020).
- Liu et al. (2019) N. Liu, G. Bo, Y. Liu, X. Xu, Y. Du, and S. X. Dou, Small 15, 1805147 (2019).
- Chia et al. (2021) H. L. Chia, J. Sturala, R. D. Webster, and M. Pumera, Advanced Functional Materials 31, 2011125 (2021).
- Wang et al. (2022) X. Wang, J. Feng, F. Hou, L. Dong, C. Long, D. Li, and J. Liang, Chemical Communications 58, 5717 (2022).
- Ryan et al. (2019) B. J. Ryan, M. P. Hanrahan, Y. Wang, U. Ramesh, C. K. Nyamekye, R. D. Nelson, Z. Liu, C. Huang, B. Whitehead, J. Wang, et al., Chemistry of Materials 32, 795 (2019).
- Yaokawa et al. (2021) R. Yaokawa, A. Nagoya, and H. Nakano, Journal of Solid State Chemistry 295, 121919 (2021).
- Yaokawa et al. (2018) R. Yaokawa, A. Nagoya, K. Mukai, and H. Nakano, Acta Materialia 151, 347 (2018).
- Gao et al. (2021) G. Gao, L. Tong, L. Yang, C. Sun, L. Xu, F. Xia, F. Geng, J. Xue, H. Gong, and J. Zhu, Applied Physics Letters 118 (2021).
- Khamh et al. (2018) H. Khamh, E. Sachet, K. Kelly, J.-P. Maria, and S. Franzen, Journal of Materials Chemistry C 6, 8326 (2018).
- Tong et al. (2016) C. Tong, J. Yun, Y.-J. Chen, D. Ji, Q. Gan, and W. A. Anderson, ACS applied materials & interfaces 8, 3985 (2016).
- Wang et al. (2019) Z. Wang, C. Chen, K. Wu, H. Chong, and H. Ye, physica status solidi (a) 216, 1700794 (2019).
- Fukumoto et al. (2022) M. Fukumoto, Y. Hirose, B. A. Williamson, S. Nakao, K. Kimura, K. Hayashi, Y. Sugisawa, D. Sekiba, D. O. Scanlon, and T. Hasegawa, Advanced Functional Materials 32, 2110832 (2022).
- Tobash and Bobev (2007) P. H. Tobash and S. Bobev, Journal of Solid State Chemistry 180, 1575 (2007).
- Vogg et al. (2000b) G. Vogg, M. Brandt, M. Stutzmann, I. Genchev, A. Bergmaier, L. Görgens, and G. Dollinger, Journal of crystal growth 212, 148 (2000b).
- Hara et al. (2021) K. O. Hara, S. Kunieda, J. Yamanaka, K. Arimoto, M. Itoh, and M. Kurosawa, Materials Science in Semiconductor Processing 132, 105928 (2021).
- Arguilla et al. (2017) M. Arguilla, N. Cultrara, M. Scudder, S. Jiang, R. Ross, and J. Goldberger, Journal of Materials Chemistry C 5, 11259 (2017).
- Itoh et al. (2022) M. Itoh, M. Araidai, A. Ohta, O. Nakatsuka, and M. Kurosawa, Japanese Journal of Applied Physics 61, SC1048 (2022).
- Pinchuk et al. (2014) I. V. Pinchuk, P. M. Odenthal, A. S. Ahmed, W. Amamou, J. E. Goldberger, and R. K. Kawakami, Journal of Materials Research 29, 410 (2014).
- Bianco et al. (2013) E. Bianco, S. Butler, S. Jiang, O. D. Restrepo, W. Windl, and J. E. Goldberger, ACS nano 7, 4414 (2013).
- Markov et al. (2019) M. Markov, S. E. Rezaei, S. N. Sadeghi, K. Esfarjani, and M. Zebarjadi, Physical Review Materials 3, 095401 (2019).
- Gordon (2000) R. G. Gordon, MRS bulletin 25, 52 (2000).
- Katz et al. (2001) O. Katz, V. Garber, B. Meyler, G. Bahir, and J. Salzman, Applied physics letters 79, 1417 (2001).
- Falcone et al. (2022) V. Falcone, A. Ballabio, A. Barzaghi, C. Zucchetti, L. Anzi, F. Bottegoni, J. Frigerio, R. Sordan, P. Biagioni, and G. Isella, APL Photonics 7 (2022).
- Huang et al. (2021) Z. Huang, S. Ke, J. Zhou, Y. Zhao, W. Huang, S. Chen, and C. Li, Chinese Physics B 30, 037303 (2021).
- Huang et al. (2018) Z. Huang, Y. Mao, A. Chang, H. Hong, C. Li, S. Chen, W. Huang, and J. Wang, Applied Physics Express 11, 102203 (2018).
- Cho et al. (2016) M. Cho, J.-H. Seo, M. Kim, J. Lee, D. Liu, W. Zhou, Z. Yu, and Z. Ma, Journal of Vacuum Science & Technology B 34 (2016).
- An et al. (2022) S. An, Y. Liao, S. Shin, and M. Kim, Advanced Materials Technologies 7, 2100912 (2022).
- Dushaq et al. (2017) G. Dushaq, A. Nayfeh, and M. Rasras, Optics express 25, 32110 (2017).
- Ang et al. (2008) K.-W. Ang, M.-B. Yu, S.-Y. Zhu, K.-T. Chua, G.-Q. Lo, and D.-L. Kwong, IEEE Electron Device Letters 29, 708 (2008).
- Zumuukhorol et al. (2019) M. Zumuukhorol, S. Boldbaatar, K.-H. Shim, C.-J. Choi, Z. Khurelbaatar, and S.-N. Lee, Journal of the Korean Physical Society 74, 713 (2019).
- Yin et al. (2022) Q. Yin, G. Si, J. Li, S. Wali, J. Ren, J. Guo, and H. Zhang, Nanotechnology 33, 255502 (2022).
- Chang et al. (2019) K. E. Chang, C. Kim, T. J. Yoo, M. G. Kwon, S. Heo, S.-Y. Kim, Y. Hyun, J. I. Yoo, H. C. Ko, and B. H. Lee, Advanced Electronic Materials 5, 1800957 (2019).
- Zeng et al. (2013) L.-H. Zeng, M.-Z. Wang, H. Hu, B. Nie, Y.-Q. Yu, C.-Y. Wu, L. Wang, J.-G. Hu, C. Xie, F.-X. Liang, et al., ACS applied materials & interfaces 5, 9362 (2013).
- Kim et al. (2021) C. Kim, T. J. Yoo, K. E. Chang, M. G. Kwon, H. J. Hwang, and B. H. Lee, Nanophotonics 10, 1573 (2021).
- Yang et al. (2017) F. Yang, H. Cong, K. Yu, L. Zhou, N. Wang, Z. Liu, C. Li, Q. Wang, and B. Cheng, ACS Applied Materials & Interfaces 9, 13422 (2017).
- Luo et al. (2019) L.-B. Luo, D. Wang, C. Xie, J.-G. Hu, X.-Y. Zhao, and F.-X. Liang, Advanced Functional Materials 29, 1900849 (2019).
- Zhao et al. (2020) M. Zhao, Z. Xue, W. Zhu, G. Wang, S. Tang, Z. Liu, Q. Guo, D. Chen, P. K. Chu, G. Ding, et al., ACS applied materials & interfaces 12, 15606 (2020).
- Xiong et al. (2022) G. Xiong, G. Zhang, X. Yang, and W. Feng, Advanced Electronic Materials 8, 2200620 (2022).
- Wang et al. (2020) Q. Wang, C. Zhou, and Y. Chai, Nanoscale 12, 8109 (2020).
- Zhou et al. (2018) C. Zhou, S. Raju, B. Li, M. Chan, Y. Chai, and C. Y. Yang, Advanced Functional Materials 28, 1802954 (2018).
- Hohenberg and Kohn (1964) P. Hohenberg and W. Kohn, Phys. Rev 136, B864 (1964).
- Kohn and Sham (1965) W. Kohn and L. J. Sham, Physical review 140, A1133 (1965).
- Kresse and Furthmüller (1996a) G. Kresse and J. Furthmüller, Computational materials science 6, 15 (1996a).
- Kresse and Furthmüller (1996b) G. Kresse and J. Furthmüller, Physical review B 54, 11169 (1996b).
- Perdew et al. (1996) J. P. Perdew, K. Burke, and M. Ernzerhof, Physical review letters 77, 3865 (1996).
- Adler (1962) S. L. Adler, Physical Review 126, 413 (1962).
- Wiser (1963) N. Wiser, Physical Review 129, 62 (1963).
- Gajdoš et al. (2006) M. Gajdoš, K. Hummer, G. Kresse, J. Furthmüller, and F. Bechstedt, Physical Review B 73, 045112 (2006).
- Zhizhchenko et al. (2020) A. Y. Zhizhchenko, P. Tonkaev, D. Gets, A. Larin, D. Zuev, S. Starikov, E. V. Pustovalov, A. M. Zakharenko, S. A. Kulinich, S. Juodkazis, et al., Small 16, 2000410 (2020).
- Shevlyagin et al. (2020) A. Shevlyagin, I. Chernev, N. Galkin, A. Gerasimenko, A. Gutakovskii, H. Hoshida, Y. Terai, N. Nishikawa, and K. Ohdaira, Solar Energy 211, 383 (2020).