Open Access Article
G. Gurieva
*a,
D. M. Többens
a,
A. Manjon Sanz
b,
V. Rotaru
cd,
C. Hennig
ef,
M. Kirkham
b,
M. Guc
c and
S. Schorr
ag
aHelmholtz-Zentrum Berlin für Materialien und Energie (HZB), 14109 Berlin, Germany. E-mail: galina.gurieva@helmholtz-berlin.de
bNeutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
cCatalonia Institute for Energy Research (IREC), 08930 Barcelona, Spain
dFacultat de Fisica, Universitat de Barcelona (UB), 08028 Barcelona, Spain
eHelmholtz-Zentrum Dresden-Rossendorf (HZDR), Institute of Resource Ecology, Dresden, Germany
fRossendorf Beamline (BM20), European Synchrotron Radiation Facility, Grenoble, France
gFreie Universität Berlin, Institute of Geological Sciences, 12249 Berlin, Germany
First published on 28th November 2025
Developing low-cost, sustainable, and environmentally friendly top absorber layers for tandem solar cells is essential to advancing photovoltaic technologies and accelerating the transition to renewable energy. In this work, we explore the potential of tetravalent (Cu2Zn(GexSi1−x)Se4) cation mutations in chalcogenide compound semiconductors with the aim of finding a material with increased band gap and reduced structural disorder. A combination of high-resolution synchrotron powder diffraction and neutron powder diffraction was used to determine the atomic positions and monoclinic angles in monoclinic wurtz–kesterite type Cu2Zn(GexSi1−x)Se4 mixed crystals as well as to determine the cation distribution in the crystal structure of Ge-rich kesterite-type and Si-rich wurtz–kesterite type mixed crystals. These investigations enabled us to deduce the structural transition scenario within the Cu2Zn(GexSi1−x)Se4 series. The transition occurs via a region where two phases with different crystal structures, tetragonal and monoclinic and thus a different distortion of the coordination tetrahedra, but the same cation distribution within the element specific cation sites co-exist. Thus, the structural transition between the kesterite and the wurtz–kesterite structure within the Cu2Zn(GexSi1−x)Se4 series is a distortion driven transition. The study identifies cation mutation in quaternary chalcogenides as a promising strategy beyond chalcopyrites and kesterites for low cost and environmentally friendly top absorbers in tandem solar cells.
Due to its non-toxicity and abundance of the constituent elements, Cu2ZnSn(S,Se)4 (CZTSSe) is a promising low-cost alternative as absorber layer in critical-raw material free thin film solar cells. A number of new world records were announced recently, the highest being 15.1% PCE (power-conversion-efficiency) in 2024.2,3 But the low open circuit voltage with respect to the band gap phenomenon continues to be a critical issue. The absorber band tailing caused by the exceptionally high density of CuZn and ZnCu anti-sites causing Cu/Zn disorder4 is believed to be one of the reasons for the limited open-circuit voltage in CZTSSe-based devices.5
One of the ways previously shown to completely block structural disorder, especially the Cu/Zn disorder,4 is a change of the atomic crystal structure, e.g. see ref. 6. In doing so, the stacking sequence of lattice planes perpendicular to the crystallographic c-axis has to change, but keeping the structural building block, the Cu2ZnCIVX tetrahedron (see Fig. 1a). Different cation and anion mutation series were studied for this purpose.7 One example could be substituting Zn2+ by Mn2+ in Cu2ZnSnSe4.8 In the resulting series, Cu2(Zn1−xMnx)SnSe4, both end members adopt different crystal structures: Cu2ZnSnSe4 crystallizes in the kesterite-type structure whereas Cu2MnSnSe4 adopts the stannite-type crystal structure, and both are tetragonal.
Another strategy would be to substitute the tetravalent cation in the Cu2Zn(GexSi1−x)Se4 semiconductor solid solution; in this case, the structural change is even more drastic from a tetragonal structure in Cu2ZnGeSe4,9 which was shown to crystallize in a kesterite type structure (space group I
) (Fig. 1b), to a monoclinic structure in Cu2ZnSiSe4 (space group Pn),10 which was recently shown by multiple energy anomalous diffraction (MEAD) to be wurtz–kesterite (Fig. 1c). A comprehensive comparison of these two structures can be found in ref. 11, highlighting the changes happening to the Cu2ZnCIVX anion tetrahedron which is the building block for both structures.
In one of our previous studies,12 we have shown the crystal structure of Cu2ZnGexSi1−xSe4 mixed crystals in the regions 0 ≤ x ≤ 0.4 and 0.6 ≤ x < 1, while the region in the middle of the series (0.4 < x < 0.6) was still an open question. Here, we studied the crystal structure, cation distribution and intrinsic point defect scenario through the whole Cu2Zn(GexSi1−x)Se4 series of powder samples by synchrotron X-ray and neutron diffraction at large scale facilities. Applying neutron diffraction enables us to differentiate the isoelectronic cations Cu+, Zn2+ and Ge4+ in the crystal structure analysis, which is essential for obtaining insights into the structural phase transition mechanism as well as being able to detect and even more quantify the Cu/Zn disorder. The detailed analysis of the Raman spectra allowed to define the evolution of the main vibrational modes in the solid solutions and to confirm the phase content of the powders. The band gap energy of the studied mixed crystals was deduced from diffuse reflectance measurements.
A proper understanding of the mechanism for the structural phase transition in this solid solution is of uttermost importance for the potential engineering of devices based on these materials. Several questions should be answered in order to achieve such an understanding. These include the details of the transition mechanism from kesterite to wurtz–kesterite structure, the possibility of the existence of Cu/Zn disorder in the wurtz–kesterite structure, and the potential influence over the changes in the band gap.
038 eV was used; the wavelength was refined to 0.72769(1) Å using LaB6 NIST 660c as a calibration standard. The samples were mounted in glass capillaries of 0.3 mm diameter; in order to reduce X-ray absorption, they were mixed with ≈70% ground glass. Data were collected in the 2θ range 3–65°, at 3 different positions along the length of each capillary to further improve particle statistics.
| Sample name | Cu/(Zn + Ge + Si) | Zn/(Ge + Si) | Ge/(Si + Ge) | Chemical formula | Typea 15 | Secondary phases |
|---|---|---|---|---|---|---|
| a G: Zn2+IV + Zn2+I; F: Cu+IV + Cu+I + Zn2+I; I: Cu+IV + 3Cu+I; E: 2VCu + IV4+Zn; J: IV4+Cu + 3VCu. | ||||||
| Ge100 (840C) | 1.04 (1) | 1.12 (1) | 1 | Cu2.09Zn1.06Ge0.95Se4 | G–F | — |
| Ge90Si10(840C) | 1.03 (1) | 1.11 (1) | 0.90 (1) | Cu2.07Zn1.06Ge0.86Si0.10Se4 | G–F | — |
| Ge80Si20(840C) | 1.06 (1) | 1.05 (1) | 0.80 (1) | Cu2.10Zn1.02Ge0.77Si0.20Se4 | F–I | ZnSe |
| Ge70Si30(840C) | 1.01 (1) | 1.07 (1) | 0.68 (1) | Cu2.03Zn1.04Ge0.66Si0.31Se4 | G–F | ZnSe |
| Ge60Si40(840C) | 1.03 (1) | 1.04 (1) | 0.58 (1) | Cu2.06Zn1.02Ge0.57Si0.41Se4 | F–I | ZnSe |
| Ge55Si45(840C) | 1.05 (1) | 1.09 (1) | 0.55 (1) | Cu2.10Zn1.04Ge0.53Si0.43Se4 | F–I | ZnSe, CuSe, Si |
| Ge55Si45(900C) | 1.07 (1) | 1.12 (1) | 0.55 (1) | Cu2.13Zn1.05Ge0.52Si0.42Se4 | F–I | ZnSe |
| Ge50Si50(840C) | 1.05 (1) | 1.09 (1) | 0.50 (1) | Cu2.10Zn1.04Ge0.48Si0.48Se4 | F–I | ZnSe |
| Ge50Si50(900C) | 1.08 (1) | 1.10 (1) | 0.49 (1) | Cu2.14Zn1.04Ge0.47Si0.48Se4 | F–I | ZnSe |
| Ge45Si55(840C) | 1.05 (1) | 1.08 (1) | 0.44 (1) | Cu2.10Zn1.03Ge0.42Si0.54Se4 | F–I | ZnSe |
| Ge45Si55(900C) | 1.08 (1) | 1.09 (1) | 0.43 (1) | Cu2.14Zn1.04Ge0.41Si0.54Se4 | F–I | ZnSe |
| Ge40Si60(900C) | 0.97 (1) | 0.98 (1) | 0.37 (1) | Cu1.95Zn1.00Ge0.37Si0.64Se4 | E–J | ZnSe, CuSe |
| Ge30Si70(900C) | 0.98 (1) | 0.97 (1) | 0.28 (1) | Cu1.96Zn0.99Ge0.28Si0.74Se4 | E–J | ZnSe, CuSe |
| Ge20Si80(900C) | 0.95 (1) | 0.95 (1) | 0.17 (1) | Cu1.92Zn0.98Ge0.18Si0.85Se4 | E–J | ZnSe, CuSe |
| Ge10Si90(900C) | 0.98 (1) | 0.97 (1) | 0.09 (1) | Cu1.97Zn0.98Ge0.09Si0.93Se4 | E–J | ZnSe, CuSe |
| Ge0(900C) | 0.96 (1) | 0.95 (1) | 0 | Cu1.93Zn0.98Si1.03Se4 | E–J | Cu2Se, ZnSe |
To deepen the understanding of the phase formation, Raman spectra of Cu2Zn(GexSi1−x)Se4 solid solutions measured under different excitation wavelengths were analyzed. Continuous changes in Raman spectra are observed with the variation of the Ge/(Ge + Si) ratio in the Cu2Zn(GexSi1−x)Se4 solid solutions independent of the applied excitation wavelength (Fig. 2). Raman spectra of the end members of the solid solution exhibit their characteristic Raman peaks. It is noted here that in the case of Cu2ZnGeSe4 and Cu2ZnSiSe4, the measurements performed under 785 nm and 532 nm excitation wavelengths, respectively, are close to resonant conditions. This enables a more detailed analysis of the spectra of the end members. However, it also introduces additional complexity when analyzing the evolution of the Raman spectra across the entire compositional range of solid solutions. In this context, Raman measurements performed under a 442 nm excitation wavelength provide a clearer resolution of spectral variations while avoiding resonance effects, which is preferable when studying the wide compositional range of Cu2Zn(GexSi1−x)Se4. In the case of the Cu2ZnGeSe4 compound, the corresponding peaks are seen at 177, 205, 270, and 282 cm−1 (peaks indicated with blue color),17 whereas for the Cu2ZnSiSe4 compound characteristic peaks are seen at 170, 179, 211, 222, and 227 cm−1 (peaks indicated with green color).18 In the case of solid solutions, mainly a two-mode behavior is observed for different peaks, which is typical of solid solutions of complex compounds.19 Moreover, some new peaks are detected in the range 220–260 cm−1 (indicated with red color). One of these peaks, at 250 cm−1, can be assigned to the ZnSe secondary phase, the presence of which was also detected by WDX analysis, and the 442 nm excitation wavelength is close to resonant for this compound allowing us to detect even the small concentrations of this secondary phase.20 However, this peak is also observed in the spectra measured under 532 nm excitation (with lower relative intensity), which is not sensitive to the small amount of the ZnSe secondary phase. This suggests a more complex nature of the band at 250 cm−1, where two peaks (one from ZnSe and another from Cu2Zn(GexSi1−x)Se4 solid solution) can be overlapped. Additionally, a peak at 234 cm−1 cannot be assigned to any expected secondary phase (ZnSe or CuxSe). This suggests that both peaks at 234 and 250 cm−1 can be assigned to new Raman modes related to solid solution, and some of the modes of end point compounds have a three-mode behavior. Similar behavior has been reported in the case of Cu2ZnSn(S,Se)4 solid solutions with anions substitutions,21 but not in the case of quaternary compounds with substituted cations, which can be a specific behavior for the solid solutions with significant change of structure of the end point members.
![]() | ||
| Fig. 2 Raman spectra of Cu2Zn(GexSi1−x)Se4 solid solution measured at 442, 532 and 785 nm excitation wavelengths. | ||
Further analysis of the Raman peaks of the solid solution allows the detection of two specific spectral ranges where the peaks related to the end members are dominating: (1) 197–211 cm−1 region for the Cu2ZnGeSe4-related peaks; (2) 216–234 cm−1 region for the Cu2ZnSiSe4-related peaks. Calculation of the relative integrated intensity of these two spectral ranges directly shows modifications in the Raman spectra with the change of the Ge/(Ge + Si) ratio (see Fig. 3). Considering the slopes of the graph, three distinct regions can be identified, corresponding to Si-rich (Ge/(Ge + Si) ≤ 0.4) samples, Ge-rich (Ge/(Ge + Si) > 0.55) samples, as well as an intermediate region of the samples' compositional range. This leads to the assumption about different structures in the Ge- and Si-rich regions and coexistence of both kesterite and wurtz–kesterite structures in the middle compositional region, which is further developed by in-depth structural analysis.
Refinement of the crystal structure was done by a full pattern Rietveld refinement22 using the FullProf Suite software package.23 The kesterite type structure (I
) with Cu on Wyckoff position 2a: (0, 0, 0) and 2c: (0, ½, ¼), Zn on 2d: (0, ½, ¾), Ge on 2b: (½, ½, 0) and Se on 8g: (x, y, z) was used as the structural model in the refinement of the germanium rich mixed crystals9 and wurtz–kesterite type structure (Pn) with all of the cations and anions found on general positions (x, y, z)10 for the silicon rich mixed crystals. The initial refinement of the synchrotron data used the Thompson–Cox–Hastings pseudo-Voigt algorithm for peak shape in combination with anisotropic broadening modelled by 2nd order symmetry adapted spherical harmonics, nominal composition cation occupations, neutral atomic scattering factors and isotropic Debye–Waller factors with identical values each for all selenium and metal atoms in each structure. In the monoclinic structure, the x and z positions of the (Ge, Si) cation site were kept fixed at values of 0 and 5/8, respectively, in order to avoid random drift, as the origin of these directions is not fixed by symmetry in space group Pn. Due to the very good counting statistics, even very minor impurities gave rise to observable Bragg peaks. These were modelled as additional phases with structure parameters taken from the ICSD or, if not identifiable, excluded from the refined data. The results of Rietveld refinement of these data can be found in the SI.
In the subsequent Rietveld refinement of the neutron diffraction data, the monoclinic angle and the cation and anion positions were fixed to the values obtained from the refinement of synchrotron X-ray diffraction data. The refinement of site occupancy factors (SOFs) was done without any chemical constraints. Three neutron powder diffraction patterns and the corresponding Rietveld analysis are shown exemplarily (see Fig. 4) for the end members Cu2ZnGeSe4 (applying the kesterite-type structure) and Cu2ZnSiSe4 (applying the wurtz–kesterite-type structure) as well as a Cu2ZnGexSi1−xSe4 mixed crystal with Ge/(Ge + Si) = 0.45 where both structure types were applied. For the latter, it can be clearly seen in the diffractogram (Fig. 4c) that Bragg peaks attributed to both kesterite and wurtz–kesterite structures are present. For mixed crystals in the 0.45 ≤ Ge/(Ge + Si) ≤ 0.55 region, both the tetragonal and monoclinic structures were refined. Thus, it was concluded that in the compositional region of the structural phase transition mixed crystals with the same chemical composition but different crystal structure, either the kesterite or the wurtz–kesterite structure coexist.
The lattice parameters, determined by Rietveld analysis of the diffraction data, follow Vegard's law24 for both structural regions. In order to follow the lattice constant parameter development through the whole series, the effective lattice parameter aeff, adapted from ref. 1, was calculated according to eqn (1) (see Fig. 5a).
![]() | (1) |
The steady, linear increase of aeff suggests that mixed crystals in the 0.45 ≤ Ge/(Ge + Si) ≤ 0.55 region have indeed a homogeneous chemical composition, while two structural phases are present simultaneously.7
The lattice parameter for the end members and the mixed crystals of the Cu2Zn(GexSi1−x)Se4 series are presented in Table 2 and Fig. 5b. The details of the refinement can be found in the SI.
| Sample name | atetr, Å | ctetr, Å | amono, Å | bmono, Å | cmono, Å | βmono, ° | aeff tetr, Å | aeff mono, Å |
|---|---|---|---|---|---|---|---|---|
| Ge100(840C) | 5.610 (1) | 11.043 (2) | 5.580 (3) | |||||
| Ge90Si10(840C) | 5.606 (1) | 11.031 (2) | 5.576 (3) | |||||
| Ge80Si20(840C) | 5.599 (1) | 11.017 (2) | 5.569 (3) | |||||
| Ge70Si30(840C) | 5.600 (1) | 10.999 (2) | 5.566 (3) | |||||
| Ge60Si40(840C) | 5.595 (1) | 10.983 (2) | 5.560 (3) | |||||
| Ge55Si45(840C) | 5.595 (1) | 10.975 (2) | 5.559 (3) | |||||
| Ge55Si45(900C) | 5.595 (1) | 10.976 (2) | 5.559 (3) | |||||
| Ge50Si50(840C) | 5.594 (1) | 10.967 (2) | 5.557 (3) | |||||
| Ge50Si50(900C) | 5.593 (1) | 10.968 (2) | 5.556 (3) | |||||
| Ge45Si55(840C) | 5.592 (1) | 10.959 (2) | 7.846 (1) | 6.759 (1) | 6.471 (1) | 90.0160 (1) | 5.554 (3) | 5.557 (3) |
| Ge45Si55(900C) | 5.592 (1) | 10.959 (2) | 7.845 (1) | 6.758 (1) | 6.470 (1) | 90.0131 (1) | 5.554 (3) | 5.556 (3) |
| Ge40Si60(900C) | 7.847 (1) | 6.757 (1) | 6.469 (1) | 90.0167 (1) | 5.556 (3) | |||
| Ge30Si70(900C) | 7.840 (1) | 6.750 (1) | 6.464 (1) | 90.0179 (1) | 5.551 (3) | |||
| Ge20Si80(900C) | 7.835 (1) | 6.743 (1) | 6.459 (1) | 90.0256 (1) | 5.546 (3) | |||
| Ge10Si90(900C) | 7.828 (1) | 6.736 (1) | 6.454 (1) | 90.0473 (1) | 5.541 (3) | |||
| Ge0(900C) | 7.822 (1) | 6.730 (1) | 6.450 (1) | 90.0593 (1) | 5.537 (3) |
The average neutron scattering length analysis method25 was applied to determine the distribution of the cations Cu+, Zn2+ Ge4+ and Si4+ on the four structural sites of the kesterite and wurtz–kesterite type structures. The experimental average neutron scattering lengths (b) of the cation structural sites were calculated according to eqn (2) for the kesterite type structure and eqn (3) for the wurtz–kesterite type structure using the cation site occupancy factors SOF2a, SOF2b, SOF2c and SOF2d as well as SOF2a1, SOF2a2, SOF2a3 and SOF2a4 determined by Rietveld analysis of neutron diffraction data.
![]() | (2) |
![]() | (3) |
The calculated average neutron scattering length is derived from a cation distribution model taking into account the experimentally determined total amounts of each element, as obtained by WDX analysis, and assuming that all four cation sites are fully occupied. A comparison of the experimental and calculated average neutron scattering length is shown in Fig. 6. The presence of intrinsic point defects is included in the cation distribution model. For example, the decrease of the average neutron scattering length of the Cu site 2a in the kesterite type or 2a1 in the wurtz–kesterite type structure indicates the presence of a less scattering species on that structural site, such as a copper vacancy (VCu) or the anti-site defect ZnCu. On the other hand, the increase of the average scattering length on the Zinc site 2d or 2a3 site in the kesterite or wurtz–kesterite structure, respectively, would indicate the presence of a species scattering more, e.g. Cu or Ge forming an anti-site. It is important that the sum of a cation species on the different cation sites is in good agreement with the chemical composition of the phase determined by WDX analysis. The resulting cation distributions for all the mixed crystals in the Cu2Zn(GexSi1−x)Se4 series are presented in Fig. 7. The detailed cation distribution model is presented in Table 3. The results for the end member Cu2ZnGeSe4 are in agreement to ref. 9 and those for the other end member Cu2ZnSiSe4 are in agreement with ref. 10. In the Cu2Zn(GexSi1—x)Se4 mixed crystals, both Si and Ge are randomly distributed on the 2b/2a4 sites of the kesterite/wurtz–kesterite structure also across the structural transition. In the Cu-rich mixed crystals with 0.58 ≤ Ge/(Ge + Si) ≤ 1, there is a small amount of copper on these sites forming anti-sites CuGe and CuSi, respectively, corresponding to the off-stoichiometry type scenario.15 The 2a/2a1 site is occupied by Cu in agreement with the kesterite and wurtz–kesterite crystal structure definition, with a small presence of VCu in Cu-poor mixed crystals, which is also in a very good agreement with the off-stoichiometry type scenario.15 Additionally, copper and zinc interstitials (Cui and Zni) are present in Cu-rich and Zn-rich mixed crystals. It can be summarized that all the off-stoichiometry related point defects found experimentally (see Table 4 and Fig. 8) are in agreement with the off-stoichiometry type model15 and were found to not be influenced by the structural phase transition.
| Sample name | 2a/2a1 | 2c/2a2 | 2d/2a3 | 2b/2a4 | Interstitials | ||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| V | Cu | V | Cu | Zn | Cu | Zn | Ge | Si | Cu | Zn | Ge | Si | Cui | Zni | |
| Ge100(840C) | 1.00 | 0.87 | 0.13 | 0.13 | 0.87 | 0.05 | 0.95 | 0.05 | 0.05 | ||||||
| Ge90Si10(840C) | 1.00 | 0.80 | 0.20 | 0.20 | 0.80 | 0.03 | 0.01 | 0.86 | 0.10 | 0.03 | 0.05 | ||||
| Ge80Si20(840C) | 1.00 | 0.80 | 0.20 | 0.20 | 0.80 | 0.03 | 0.77 | 0.20 | 0.07 | 0.02 | |||||
| Ge70Si30(840C) | 1.00 | 0.90 | 0.10 | 0.10 | 0.90 | 0.02 | 0.01 | 0.66 | 0.31 | 0.02 | 0.03 | ||||
| Ge60Si40(840C) | 1.00 | 0.89 | 0.11 | 0.11 | 0.89 | 0.02 | 0.57 | 0.41 | 0.03 | 0.02 | |||||
| Ge55Si45(840C) | 1.00 | 0.75 | 0.25 | 0.25 | 0.75 | 0.06 | 0.52 | 0.42 | 0.06 | 0.04 | |||||
| Ge55Si45(900C) | 1.00 | 0.75 | 0.25 | 0.25 | 0.75 | 0.06 | 0.52 | 0.42 | 0.08 | 0.05 | |||||
| Ge50Si50(840C) | 0.05 | 0.95 | 0.70 | 0.30 | 0.30 | 0.70 | 0.06 | 0.47 | 0.48 | 0.06 | 0.04 | ||||
| Ge50Si50(900C) | 0.05 | 0.95 | 0.70 | 0.30 | 0.30 | 0.70 | 0.06 | 0.47 | 0.48 | 0.09 | 0.04 | ||||
| Ge45Si55(840C) | 1.00 | 0.70 | 0.30 | 0.30 | 0.70 | 0.05 | 0.42 | 0.54 | 0.06 | 0.03 | |||||
| Ge45Si55(900C) | 1.00 | 0.70 | 0.30 | 0.30 | 0.70 | 0.05 | 0.42 | 0.54 | 0.09 | 0.04 | |||||
| Ge40Si60(900C) | 0.03 | 0.97 | 0.01 | 0.88 | 0.11 | 0.10 | 0.90 | 0.36 | 0.64 | ||||||
| Ge30Si70(900C) | 1.00 | 0.04 | 0.89 | 0.07 | 0.07 | 0.92 | 0.01 | 0.26 | 0.74 | ||||||
| Ge20Si80(900C) | 0.04 | 0.96 | 0.03 | 0.91 | 0.05 | 0.05 | 0.93 | 0.02 | 0.15 | 0.85 | |||||
| Ge10Si90(900C) | 1.00 | 0.03 | 0.92 | 0.05 | 0.05 | 0.93 | 0.02 | 0.07 | 0.93 | ||||||
| Ge0(900C) | 1.00 | 0.06 | 0.94 | 0.98 | 0.02 | 1.00 | |||||||||
| Sample name | CuIV | err+ | err− | ZnIV | err+ | err− | CuI | err | Zni | err | Cu/Zn disorder | err |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Ge100(840C) | 1.352 | 0.201 | 0.317 | 0.201 | 0.029 | 0.201 | 1.324 | 0.029 | 1.525 | 0.029 | 3.740 | 0.471 |
| Ge90Si10(840C) | 0.923 | 0.346 | 0.029 | 0.375 | 0.144 | 0.375 | 0.923 | 0.029 | 1.298 | 0.029 | 5.769 | 0.169 |
| Ge80Si20(840C) | 0.984 | 0.434 | 0.087 | 1.969 | 0.029 | 0.463 | 0.029 | 5.790 | 0.751 | |||
| Ge70Si30(840C) | 0.464 | 0.232 | 0.029 | 0.319 | 0.157 | 0.319 | 0.435 | 0.029 | 0.783 | 0.029 | 2.899 | 0.657 |
| Ge60Si40(840C) | 0.640 | 0.349 | 0.029 | 0.960 | 0.029 | 0.465 | 0.029 | 3.199 | 0.816 | |||
| Ge55Si45(840C) | 1.368 | 0.524 | 0.378 | 1.688 | 0.029 | 1.222 | 0.029 | 7.276 | 0.632 | |||
| Ge55Si45(900C) | 1.746 | 0.262 | 0.143 | 2.241 | 0.058 | 1.455 | 0.029 | 7.276 | 1.996 | |||
| Ge50Si50(840C) | 1.399 | 0.029 | 0.903 | 1.865 | 0.029 | 1.195 | 0.029 | 8.741 | 1.253 | |||
| Ge50Si50(900C) | 1.603 | 0.029 | 0.105 | 2.507 | 0.058 | 1.137 | 0.029 | 8.743 | 1.984 | |||
| Ge45Si55(840C) | 1.226 | 0.029 | 0.759 | 1.838 | 0.029 | 0.963 | 0.029 | 8.754 | 1.084 | |||
| Ge45Si55(900C) | 1.546 | 0.029 | 0.963 | 2.509 | 0.058 | 0.929 | 0.029 | 8.753 | 1.392 |
| Sample name | VCu | err | IVZn | err+ | err− | IVCu | err+ | err− | Cu/Zn disorder | err |
|---|---|---|---|---|---|---|---|---|---|---|
| Ge40Si60(900C) | 0.979 | 0.477 | 0.126 | 0.025 | 0.025 | 0.126 | 0.025 | 0.050 | 2.510 | 0.253 |
| Ge30Si70(900C) | 0.931 | 0.252 | 0.327 | 0.101 | 0.050 | 0.076 | 0.025 | 0.025 | 1.761 | 0.344 |
| Ge20Si90(900C) | 1.816 | 0.555 | 0.454 | 0.252 | 0.025 | 0.303 | 0.076 | 0.010 | 1.261 | 0.423 |
| Ge10Si90(900C) | 0.834 | 0.177 | 0.405 | 0.126 | 0.051 | 0.025 | 0.003 | 0.003 | 1.264 | 0.391 |
| Ge0(900C) | 1.546 | 0.912 | 0.633 | 0.329 | 0.025 | 0.101 | 0.051 | 0.025 | 0 | 0 |
The 2c and 2d sites of the kesterite type structure in Ge-rich mixed crystals show a very typical picture, where equal amounts of CuZn and ZnCu anti-site defects are present – the Cu/Zn disorder,4 which is often observed in kesterite-type Ge-containing quaternary chalcogenides.9 On the other hand, wurtz–kesterite type Cu2ZnSiSe4, with an off-stoichiometric composition Cu1.93Zn0.98Si1.03Se4 does not show any presence of Cu/Zn disorder, which is in good agreement with a previously published MEAD study.10 With the increasing amount of Ge in Cu2Zn(GexSi1−x)Se4 mixed crystals though, we see the appearance and increase of Cu/Zn disorder even in the wurtz–kesterite type structure (CuZn and ZnCu on the sites 2a2 and 2a3). Interestingly enough, it was possible to describe the cation distribution in both structural phases (kesterite and wurtz–kesterite) within the two-phase region of the same sample with the same distribution of the cations on the corresponding sites (see Fig. 6b). In this case, the Cu/Zn disorder was the strongest, with the fraction of CuZn on 2d/2a3 and ZnCu on 2c/2a2 reaching as high as 0.3 (see Table 3 and Fig. 8).
It is suggested that the structural phase transition from the kesterite type to the wurtz–kesterite type structure occurs via the co-existence of two types of domains, tetragonal and monoclinic ones, with the same chemical composition, composed of the same building blocks – anion/cation tetrahedra which are differently distorted, in this way making it a distortion driven transition. This distortion becomes visible in the bond angles within the IVSe4 tetrahedron (IV stands for the four-valent cation), see Fig. 9. The deviation of the Se–Ge–Se bond angles from the bond angle in an ideal (regular) tetrahedron (109.45°) is small in the case of Cu2ZnGeSe4 and the tetragonal mixed crystals but increases in the tetragonal phase of the 2-phase region. On the other hand, the SiSe4 coordination tetrahedron is heavily distorted resulting in six different Se–Si–Se bond angles in Cu2ZnSiSe4.
![]() | ||
| Fig. 9 Deviations from ideal bond angle (absolute values) in the anion tetrahedron around the four-valent cation (Ge and Si) with dependence on the Ge/(Ge + Si) ratio. | ||
In order to correlate the structural changes as well as the obtained point defect concentrations with the optoelectronic properties of these materials, the optical band gap energy Eg was determined by diffuse reflectance spectroscopy (DRS). Tauc plots were obtained by plotting (F(R) × hν)2 versus the photon energy,26 and the linear part of the curve was extrapolated to the baseline, and the optical band gap was extracted from the value of intersection (Fig. S1 in the SI). An estimation of the band gap energy in the same way was previously reported for Cu2ZnSnS4 (CZTS), Cu2ZnSnSe4 (CZTSe) and Cu2ZnGeSe4 (CZGSe) powder samples,11 proving the applicability of this approach. It is found that the band gap energy Eg within the Cu2Zn(GexSi1−x)Se4 series is changing in the range of 1.31 eV to 2.22 eV. The obtained Eg values can be found in Table 5 and Fig. 10. Both Ge-rich tetragonal and Si-rich monoclinic parts of the solid solution show a bowing behaviour according to eqn (4), typically observed in compound semiconductors.11
| Eg(x) = Ex=0g·(1 − x) + Ex=1g·x + b(1 − x)x with x = Ge/(Ge + Si) | (4) |
| Sample name | Eg, eV |
|---|---|
| Ge100(840C) | 1.31 (5) |
| Ge90Si10(840C) | 1.36 (5) |
| Ge80Si20(840C) | 1.41 (5) |
| Ge70Si30(840C) | 1.47 (5) |
| Ge60Si40(840C) | 1.54 (5) |
| Ge55Si45(840C) | 1.56 (5) |
| Ge55Si45(900C) | 1.57 (5) |
| Ge50Si50(840C) | 1.62 (5) |
| Ge50Si50(900C) | 1.61 (5) |
| Ge45Si55(840C) | 1.70 (5) |
| Ge45Si55(900C) | 1.71 (5) |
| Ge40Si60(900C) | 1.78 (5) |
| Ge30Si70(900C) | 1.86 (5) |
| Ge20Si80(900C) | 1.95 (5) |
| Ge10Si90(900C) | 2.04 (5) |
| Ge0(900C) | 2.22 (5) |
![]() | ||
| Fig. 10 Band gap energies of the Cu2Zn(GexSi1−x)Se4 solid solution with dependence on the cation ratio Ge/(Ge + Si). | ||
The bowing parameter b was determined to be b = −1.32 for Si-rich monoclinic mixed crystals and b = −0.48 for Ge-rich tetragonal mixed crystals, respectively. This is in a very good agreement with the fact that Vegard's law is followed in solid solutions only as long as the structure does not change. Thus, with two different bowing parameters for two different structures, two solid solutions were derived. In the region of interest for the tandem applications, the band gap of 1.6–1.8 eV (ref. 27) is also achieved within the investigated series. It partially overlaps with the two-phase region, which could have potential consequences for the efficiency of the fabricated devices.
The results of this study highlight the importance of considering alternative materials beyond the known chalcopyrites (Cu(In,Ga)Se2) or kesterites in the search for low-cost and environmentally friendly top absorber layers and demonstrate that cation mutation in quaternary chalcogenides is a promising path towards the development of highly efficient tandem solar cells.
). In Cu2Zn(GexSi1−x)Se4 mixed crystals with Ge/(Ge + Si) = 0.45–0.55, two phases with the same chemical composition but different crystal structure co-exist. The IV-valent cations (Si4+ an Ge4+) are randomly distributed on the same structural site (in both structure types) through the whole Cu2Zn(GexSi1−x)Se4. The off-stoichiometry type related intrinsic point defect scenario is not influenced by the phase transition. The structural phase transition occurs in the following way: the cations shift from special positions (in the kesterite type structure) to more general positions (in the wurtz–kesterite type structure). The distortion of the anion (and cation) tetrahedra increases, which results in a symmetry reduction, making it a distortion driven transition. The band gap range of 1.31 eV to 2.22 eV can be covered by the samples from Cu2Zn(GexSi1−x)Se4 series, which makes it promising for tandem applications.
| This journal is © The Royal Society of Chemistry 2025 |