Copper antimony chalcogenides CuSbCh2 (Ch=S, Se) are an emerging family of absorbers studied for thin-film solar cells. These non-toxic and Earth-abundant materials show a layered low-dimensional chalcostibite crystal structure, leading to interesting optoelectronic properties for applications in photovoltaic (PV) devices. This research update describes the CuSbCh2 crystallographic structures, synthesis methods, competing phases, band structures, optoelectronic properties, point defects, carrier dynamics, and interface band offsets, based on experimental and theoretical data. Correlations between these absorber properties and PV device performance are discussed, and opportunities for further increase in the efficiency of the chalcostibite PV devices are highlighted.

The most studied materials for the thin-film photovoltaic (PV) technologies are Cu(In,Ga)Se2 (CIGS, 22.6% one-sun energy conversion efficiency) and CdTe (22.1% efficiency).1 Nevertheless, the future environmental benefits and potential scale of deployment of these solar cells may be limited by their elemental scarcity (In, Ga, Te) and toxicity (Cd).2 These considerations have led to increased interest in emerging light-absorber materials with less toxic and more abundant elements, including Cu2SnS3 (CTS),3 Cu2O,4 Sb2Se3,5 Cu3N,6,7 SnS,8,9 ZnSnN2,10 and Cu2ZnSn(S,Se)4 (CZTSSe).11 Among these, the highest PV efficiency (certified 12.6%1,12) has been reported for the CZTSSe absorber. However, its further improvements may be hindered by chemical and crystallographic complexity, pointing researchers toward less complex materials.

Most of the emerging absorbers, like CZTSSe, have a three-dimensional (3D) crystal structure, where dangling bonds in grain boundaries (GBs) can act as recombination centers, causing efficiency loss.13,14 In contrast, the low-dimensional crystal structure of some potential absorbers, such as 1D-Sb2Se3,5,15 1D-SbSeI,16 2D-CuSbS2,17 2D-CuBiS2,18 and 2D-CuSbSe2,19 can be oriented hypothetically in such a way to decrease the number of dangling bonds, even at the GBs. In turn, this can minimize photoexcited charge-carrier recombination losses, which is one of the major limiting factors for high-efficiency thin-film solar cells.

One class of such non-toxic Earth-abundant less-complex and low-dimensional absorbers is ternary copper chalcogenides, such as CuBiS218 and CuSbCh2 (where Ch=S, Se) with a chalcostibite crystal structure.17,19 Despite a relatively small amount of research, these layered semiconductors have shown promising PV device efficiencies approaching 5%.19–21 The PV-relevant properties of these materials include slightly indirect bandgaps of 1.1–1.6 eV (that is, well-suited for single-junction terrestrial solar cell applications) and strong optical absorption coefficient >105 cm−1 at 0.5 eV above the bandgap (slightly better than CIGSe, CdTe, and CZTSSe, for example).19,21–23 Many other characteristics that make these chalcostibite materials worth exploring will be discussed in this research update.

A good review of the PV applications of CuSbS2 (CAS) was published three years ago,24 but much CuSbS2 research activity has occurred since that time, and there are no review papers about CuSbSe2 (CASe) as a PV absorber. Thus, in this article, we focus on the most recent research efforts (past three years) on CuSbCh2 (Ch=S, Se) as a promising family of absorber materials for thin-film solar cell applications. This paper summarizes both the currently used methods to prepare these materials and their most important PV-relevant properties, as compared to other thin-film PV absorbers. We also discuss the factors that have limited current CuSbCh2-based solar cell performance, as well as some strategies that have been used to address them.

As mentioned earlier, most of the absorbers used in solar cells have a tetrahedrally bonded lattice, where different atoms are linked to one another in 3D diamond-like structures (e.g., Si, CdTe, CuInS2, and Cu2ZnSnS4), as can be seen in Fig. 1(a). However, the CuSbCh2 chalcostibites show a very different 2D-like layered crystal structure, despite the same 1:1:2 stoichiometry as in CuInS2 chalcopyrites, and similar Sb/In ionic radii.25 In contrast to the four-fold coordinated chalcopyrite, in the chalcostibite structure the Cu atoms are four-fold coordinated, whereas the Sb atoms are in a distorted three-fold coordination [Fig. 1(b)]. This difference in the metal coordination number makes chalcostibites “line compounds”: namely, the small deviations from the nominal CuSbCh2 stoichiometry (Cu/Sb = 1) lead to phase impurities, indirectly suggesting a smaller propensity to cation disorder for the stoichiometric material. This is in contrast to the tetrahedrally coordinated CIGS, CTZS, and CTS, which are known to show cation disorder and to tolerate high levels of off-stoichiometry (up to ∼20%).25 The crystallographic differences between the 3D (e.g., CuInS2), 2D (e.g., CuSbS2), and 1D (e.g., Sb2Se3) absorbers can be seen in Fig. 1.

FIG. 1.

Crystallographic structure of (a) CuInS2, (b) CuSbS2, and (c) Sb2Se3. The atoms are not at comparable scales with respect to each other, but the unit cells are, as defined by the 15-Å scale bar on the bottom of the figure. The crystal structure pictures were created using Mercury 3.9 (Copyright CCDC 2001-2016) and the CIF files from Inorganic Crystal Structure Database (Ref. 46).

FIG. 1.

Crystallographic structure of (a) CuInS2, (b) CuSbS2, and (c) Sb2Se3. The atoms are not at comparable scales with respect to each other, but the unit cells are, as defined by the 15-Å scale bar on the bottom of the figure. The crystal structure pictures were created using Mercury 3.9 (Copyright CCDC 2001-2016) and the CIF files from Inorganic Crystal Structure Database (Ref. 46).

Close modal

As can be seen in Fig. 1(b), the CuSbCh2 crystal structure is formed by stacking layers composed of SbSe2 and CuSe3 motifs along the “c”-axis, with interlayer distances of ∼7.10 Å (for Ch=S) and ∼7.29 Å (for Ch=Se).26 Theoretical calculations reveal that these layers (along with other characteristics) are responsible for the increase in the density of states (DOS) in CuSbCh2 and that each trivalent Sb atom has a non-bonding electron pair from its 5s orbital. These factors are responsible for the larger optical extinction coefficient of CuSbCh2 than in CIGS, for example, but they also lead to higher effective masses and less efficient collection of the photogenerated charge carriers.21,26–28

Another interesting property that results from the low-dimensional structure is the possibility of crystallographically orienting the layers in different ways with respect to the substrate. Such preferential orientation may create chemically inert surfaces in 2D structures and benign grain boundaries (GBs) in 1D structures, both with fewer dangling bonds compared to the 3D structures. The benign surfaces or GBs can reduce the carrier recombination loss, improving the efficiency of thin-film solar cells. The preferential orientation is also important to the charge transport inside CuSbCh2, which is easier within the sheets than between them. Within the CuSbCh2 sheets, the carriers are transported along chemical bonds, whereas in the other case, they are required to hop between the layers, as in Sb2Se3.19,21,22 Thus, the crystallographic orientation control is vital for reduced carrier recombination and improved carrier transport in CuSbCh2 PV devices.19,21,22

A wide variety of techniques has been used to grow CuSbCh2 and they can be grouped into physical and chemical methods. Physical methods include thermal evaporation29 and co-sputtering,19,30,31 whereas chemical methods include chemical bath deposition,32,33 spray pyrolysis,34 spin coating,20,35 solution processing,36 electrodeposition,37,38 and solvo-/hydro-thermal synthesis.39,40 Below, we focus mostly on the physical vapor deposition methods for CuSbCh2 growth.

The control of the CuSbCh2 composition during growth is very important because small deviations from the nominal stoichiometry more easily lead to phase impurities than in CIGS and CZTSSe.25 For this reason, self-regulated growth (a.k.a. adsorption-controlled growth) is the most promising approach to achieve reproducible synthesis of this material, regardless of the deposition method. This kind of CuSbCh2 growth control has been demonstrated for CuSbCh2 by radio-frequency (RF) magnetron co-sputtering of Cu2Ch and Sb2Ch3 targets in excess of Sb2Ch3 vapor.19,21,22,41 For CAS, it was found that it is possible to grow highly stoichiometric phase-pure thin films at the substrate temperature of 350 °C in a wide range of Sb2S3 over-flux.22 The CAS films showed some Sb2S3 impurities at lower temperatures or decomposed to Cu12Sb4S13 at higher temperatures.

To determine the stability region for the deposition of stoichiometric phase-pure CAS films, the phase diagram has been calculated as a function of temperature and Sb2S3 partial pressure [Fig. 2(a)].22 It was concluded that for a given substrate temperature, phase-pure CuSbS2 can be grown in a ∼102 Torr dynamic range of Sb2S3 partial pressure. In turn, for a given Sb2S3 partial pressure, the region of stability extends ∼70 °C of substrate temperature in agreement with experimental data. Overall, these results suggest that no fine control of the Sb2S3 precursor flux or substrate temperature is required within this region of CuSbS2 self-regulated growth. Similar conclusions hold for CuSbSe2 adsorption-controlled growth in an Sb2Se3-rich atmosphere, but at slightly higher temperatures (∼380 °C), and with somewhat different competing phases [Sb2Se3, Cu3SbSe3, and Cu2Se in Fig. 2(b)].

FIG. 2.

Calculated phase diagrams for (a) Cu—Sb—S and (b) Cu—Sb—Se material systems, as a function of temperature and vapor pressure of Sb2Ch3, outlining the adsorption-limited self-regulated growth window for phase-pure CuSbCh2 (orange region). Figures “(a)” and “(b)” were adapted from A. W. Welch et al., Sol. Energy Mater. Sol. Cells 132, 499 (2014). Copyright 2014 Elsevier B.V. and A. W. Welch et al., Appl. Phys. Express 8, 82301 (2015). Copyright 2015 The Japan Society of Applied Physics.

FIG. 2.

Calculated phase diagrams for (a) Cu—Sb—S and (b) Cu—Sb—Se material systems, as a function of temperature and vapor pressure of Sb2Ch3, outlining the adsorption-limited self-regulated growth window for phase-pure CuSbCh2 (orange region). Figures “(a)” and “(b)” were adapted from A. W. Welch et al., Sol. Energy Mater. Sol. Cells 132, 499 (2014). Copyright 2014 Elsevier B.V. and A. W. Welch et al., Appl. Phys. Express 8, 82301 (2015). Copyright 2015 The Japan Society of Applied Physics.

Close modal

It is worth mentioning that the CuSbCh2 films grown by different methods often show small grains and/or low crystallinity, encouraging the use of post-deposition thermal treatments (TTs) to improve film quality. Such a treatment has been demonstrated under atmospheric conditions,30,35,42 H2S gas and S vapor,20 or Sb2S3 vapor.30,31,43,44 For example, taking into account the calculated stability of the CuSbCh2 phase under an Sb2Ch3-rich environment [Fig. 2(a)], it has been shown that the TT of the sputtered CAS thin films under Sb2S3 vapor increases their grain size without affecting the composition or phase purity. This TT also improved the structural quality and optoelectronic properties of the films, leading to more reproducible and efficient TT-CuSbS2 PV devices.44 In this context, it is also important to note that CuSbCh2 has a lower melting point of 480 °C (for Ch=Se) and 535 °C (for Ch=S), compared to CIGS, CZTSSe, and CdTe (∼1000 °C21,45). This should lead to grain growth at a lower temperature, making CuSbCh2 a suitable absorber for the fabrication of flexible thin-film solar cells on polymer substrates.

Based on the experimental and calculated data shown in Fig. 2, the most likely impurities found in CuSbCh2 (Ch=S or Se) are the following phases: Sb2Ch3, Cu1.8-2Ch, Cu12Sb4S13 (for the Cu-Sb-S system), and Cu3SbSe3 (for the Cu—Sb—Se system). As shown by the simulated X-ray diffraction (XRD) reference patterns in Fig. 3, there are several regions of superposition between the diffraction peaks of the CuSbCh2 and impurity phases. For CAS [Fig. 3(a)], the most difficult impurity to identify by XRD is Cu12Sb4S13 (12-4-13 phase) because its main peak is at the same position as the CAS (200)/(013) peaks. Thus, a common mistake made in the literature is attributing an increase in the (200)/(013) peaks compared to the (111)/(104) peaks to preferential orientation of CAS, instead of the actual presence of the very stable 12-4-13 secondary phase.

FIG. 3.

Simulated X-ray diffraction reference patterns for the most likely impurities found in (a) CuSbS2 and (b) CuSbSe2 films after growth or thermal treatment. The following X-ray diffraction patterns were taken from Inorganic Crystal Structures Database (ICSD®)46 and International Centre for Diffraction Data (ICDD®)–Powder Diffraction File (PDF™): CuSbS2 (ICSD code 85133), C1.8S (ICSD code 69756), Cu12Sb4S13 (ICSD code 25707), Sb2S3 (ICSD code 22176), CuSbSe2 (ICDD-PDF card no. 75-992), Cu1.8Se (ICDD-PDF card no. 4-839), Cu3SbSe3 (ICDD-PDF card no. 50-1346), and Sb2Se3 (ICDD-PDF card no. 65-2433).

FIG. 3.

Simulated X-ray diffraction reference patterns for the most likely impurities found in (a) CuSbS2 and (b) CuSbSe2 films after growth or thermal treatment. The following X-ray diffraction patterns were taken from Inorganic Crystal Structures Database (ICSD®)46 and International Centre for Diffraction Data (ICDD®)–Powder Diffraction File (PDF™): CuSbS2 (ICSD code 85133), C1.8S (ICSD code 69756), Cu12Sb4S13 (ICSD code 25707), Sb2S3 (ICSD code 22176), CuSbSe2 (ICDD-PDF card no. 75-992), Cu1.8Se (ICDD-PDF card no. 4-839), Cu3SbSe3 (ICDD-PDF card no. 50-1346), and Sb2Se3 (ICDD-PDF card no. 65-2433).

Close modal

The Raman spectroscopy can be useful for determining the presence of this impurity because its main peak is ∼20 cm−1 higher compared to the CuSbS2 main peak at 332 cm−1.47 Also, the CuSbS2 thin film with the Cu12Sb4S13 phase impurity often has higher conductivity and higher hole density (10−1 to 102 S/cm and 1018–1021 cm−3) than phase-pure CuSbS2 (10−3 to 10−2 S/cm and 1016–1017 cm−3). Thus, it should be possible to combine electric measurements with XRD and Raman spectroscopy characterizations to rule out the existence of the 12-4-13 phase.22,41 The phase purity is equally important but less difficult to determine in CuSbSe2 [Fig. 3(b)], where less stable Cu3SbS3 and Cu1.8Se secondary phases can be more easily identified by XRD, despite some overlap with the weaker CuSbSe2 (113)/(105)/(203) peaks.

The knowledge of the band structure and DOS helps us to understand the optoelectronic properties of CuSbCh2 compared with the more studied absorber materials such as CIS, as shown in Fig. 4. It has been predicted that chalcostibite CuSbS2 and CuSbSe2 have slightly indirect bandgaps [Egdir–Egind ≈ 0.1 eV, Figs. 4(a) and 4(b)] in contrast with the well-defined direct bandgap of chalcopyrite CuInSe2 [Fig. 4(c)]. Chalcostibites also present higher DOS compared to chalcopyrites. This band-structure feature not only results in a larger optical extinction coefficient (more effective absorption of photons) but also in higher effective masses (less efficient charge-carrier collection for the PV device).28 

FIG. 4.

Electronic band structure of chalcostibite-type (a) CuSbS2 and (b) CuSbSe2, in comparison with (c) the chalcopyrite-type CuInSe2 band structure. The band structure calculated with HSE06 and GGA-PBE functionals are shown in solid (red) and dashed lines (black), respectively, and the valence band maximum (VBM) was set to 0 eV. Figures were adapted from T. Maeda and T. Wada, Thin Solid Films 582, 401 (2015). Copyright 2015 Elsevier B.V.

FIG. 4.

Electronic band structure of chalcostibite-type (a) CuSbS2 and (b) CuSbSe2, in comparison with (c) the chalcopyrite-type CuInSe2 band structure. The band structure calculated with HSE06 and GGA-PBE functionals are shown in solid (red) and dashed lines (black), respectively, and the valence band maximum (VBM) was set to 0 eV. Figures were adapted from T. Maeda and T. Wada, Thin Solid Films 582, 401 (2015). Copyright 2015 Elsevier B.V.

Close modal

Theoretical calculations in combination with experimentally measured X-ray photoemission spectra show that the top of the valence band (VB) of CAS is mainly formed by strongly antibonding Cu 3d and S 3p states. In contrast, Sb 5p and S 3p states are main contributions for the bottom of the conduction band (CB).28,48 It is also interesting to note that Sb 5s electrons in CuSbCh2 are not fully inert and localized lone pairs. Instead, these Sb 5s states contribute somewhat to the formation of bonding states in the VB and antibonding states in the CB. These states are also partially associated with the rising of the energy levels of the bands,48 which has important implications for the selection of contacts to CuSbCh2 absorbers.

In agreement with the theoretical observations, the CuSbCh2 materials feature strong experimental optical absorption coefficients of >105 cm−1 for photon energies 0.5 eV higher than their bandgaps,19,21–23 as shown in Fig. 5(a), and have slightly indirect bandgaps of 1.1–1.2 eV19,21 (for Ch=Se) and 1.4–1.6 eV (for Ch=S)44 [Figs. 5(b) and 5(c)]. In Fig. 5(b), the presence of an electronic defect 0.22 eV below the bandgap energy can also be observed (i.e., 0.22 eV close to VB or CB); the different defects likely present in CuSbCh2 samples will be discussed in Sec. II D.

FIG. 5.

Graphs of (a) optical absorption coefficient (α) versus photon energy () and Tauc plots for the (b) indirect and (c) direct gaps for thin films of CuSbS2 and CuSbSe2. Data were adapted from A. W. Welch et al., Sol. Energy Mater. Sol. Cells 132, 499 (2014). Copyright 2014 Elsevier B.V. and A. W. Welch et al., Appl. Phys. Express 8, 82301 (2015). Copyright 2015 The Japan Society of Applied Physics.

FIG. 5.

Graphs of (a) optical absorption coefficient (α) versus photon energy () and Tauc plots for the (b) indirect and (c) direct gaps for thin films of CuSbS2 and CuSbSe2. Data were adapted from A. W. Welch et al., Sol. Energy Mater. Sol. Cells 132, 499 (2014). Copyright 2014 Elsevier B.V. and A. W. Welch et al., Appl. Phys. Express 8, 82301 (2015). Copyright 2015 The Japan Society of Applied Physics.

Close modal

CuSbCh2 possesses p-type conductivity and presents a tunable hole concentration in a range of 1015–1018 cm−3 suitable for PV application.22 The hole mobility (μh) has been measured by the Hall effect, showing values of 0.1–49 cm2 V−1 s−1 (for Ch=S)35 and 12 cm2 V−1 s−1 (for Ch=Se).21 The large variation in the values of μh reported in the literature may be related to differences in both the concentration of defects and anisotropic crystal structures in combination with different crystallographic textures of the films. Based on optical-pump THz probe (OPTP) transient reflection spectroscopy, it has been observed that the hole and electron mobilities in 2D-like CuSbS2 are similar to each other, in contrast to the significantly different hole and electron mobilities in 3D-bonded materials such as CIGS, CZTS, and CTS.44 

Some of the most important characteristics of the PV absorbers that should be considered for their device performance are defect properties and photoexcited charge-carrier dynamics. Thus, despite a relatively small number of CuSbCh2 publications on these topics, such studies are crucial for understanding CuSbCh2 thin-film solar cell performance and for enhancing their efficiencies. There are many methods that can be used to study defects, including first-principles calculations, photoluminescence (PL) spectroscopy, admittance spectroscopy (AS), and deep-level transient spectroscopy (DLTS). The photoexcited charge-carrier dynamics can be studied using transient OPTP reflection spectroscopy and time-resolved photoluminescence (TRPL). Some of these techniques have been previously reported for studying defects and recombination in CuSbCh2 materials.

First-principles calculations of native point defects in CuSbS249 have been performed for different chemical potentials, such as the experimentally used Sb2S3-rich conditions [Fig. 6(a)]. The results show that the most plentiful acceptor is the copper vacancy (VCu), with a very shallow transition level (0/1−) at ∼0.03 eV above the valence band maximum (VBM). Another active acceptor defect is the copper on antimony antisite (CuSb), showing deeper transition levels of 0.1 and 0.2 eV. The likely dominant donor defects are Cu interstitials (within the layer—Cui-in and between the layers—Cui-out), which are very shallow (transition levels close to the conduction band minimum, CBM). In addition, the sulfur vacancies (VS-in and VS-out) also have relatively low formation enthalpy (ΔHf). The (0/1+) transition level is around 0.15 eV and 0.35 eV for VS-out and VS-in, respectively, and the (1+/2+) transition level for VS-out is around 0.10 eV. Hence, these vacancies are amphoteric defects that may act as donors or as acceptors, depending on the position of the Fermi level (EF). These data are in agreement with the previously reported properties of defects in CuSbS2.35 Similar conclusions for CuSbSe219,26 can be made, as can be seen in Fig. 6(b).

FIG. 6.

Calculated transition energy levels, and formation enthalpy (ΔHf), at simulated equilibrium Fermi level (EF), for acceptor and donor intrinsic defects in (a) CuSbS2 and (b) CuSbSe2 at Sb2Ch3-rich condition. Data for figures (a) and (b) were adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 The Journal of Materials Chemistry A and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, respectively.

FIG. 6.

Calculated transition energy levels, and formation enthalpy (ΔHf), at simulated equilibrium Fermi level (EF), for acceptor and donor intrinsic defects in (a) CuSbS2 and (b) CuSbSe2 at Sb2Ch3-rich condition. Data for figures (a) and (b) were adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 The Journal of Materials Chemistry A and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, respectively.

Close modal

From an experimental point of view, defects in different CuSbS2 samples (as-deposited, TT, and micro-crystals) have been studied by AS, PL, and DLTS.49 Three acceptor defects were found at about 0.08, 0.17, and 0.24 eV above the VBM by the capacitance-based methods, in agreement with free-to-bound radiative photoluminescent transitions observed in these thin films. Comparing these results to the theoretical calculation [Fig. 6(a)], the shallower defect was associated with VCu, whereas the two deeper ones were attributed to CuSb and/or VS, as can be seen in the diagram of Fig. 7(a). No investigation of the experimental behavior of the CuSbSe2 defects can be found in the literature.

FIG. 7.

(a) Simplistic diagram for experimental photoluminescent transitions observed in CuSbS2 thin films. Figure adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 The Journal of Materials Chemistry A. (b) Time-domain optical pump-terahertz probe (OPTP) differential reflectance for CuSbCh2 thin films, showing the tri-exponential fit to determine the carrier lifetime (τ). Data for this figure were adapted from F. W. de Souza Lucas et al., J. Phys. Chem. C 120, 18377 (2016). Copyright 2016 American Chemical Society and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

FIG. 7.

(a) Simplistic diagram for experimental photoluminescent transitions observed in CuSbS2 thin films. Figure adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 The Journal of Materials Chemistry A. (b) Time-domain optical pump-terahertz probe (OPTP) differential reflectance for CuSbCh2 thin films, showing the tri-exponential fit to determine the carrier lifetime (τ). Data for this figure were adapted from F. W. de Souza Lucas et al., J. Phys. Chem. C 120, 18377 (2016). Copyright 2016 American Chemical Society and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

Close modal

To understand the effects of point defects on CuSbCh2 photoexcited charge carriers, transient OPTP reflection spectroscopy [a.k.a time-resolved terahertz spectroscopy (TRTS)] has been performed. An experimental study of the carrier dynamics in the solution-processed CuSbS2 nanoplates33 indicates a minority lifetime (τ) of up to ∼1 ns and a mobility (μ) of ∼1 cm2 V−1 s−1, which result in a diffusion length (Ld) estimate of ∼60 nm [Ld = (kTμτ/e)1/2 Ref. 21]. Similar results have been shown using the OPTP for the thermally treated CuSbS2 films (τ of 0.7 ns, μ of ∼4 cm2 V−1 s−1, and estimated Ld of ∼90 nm).44 The photoexcited charge-carrier dynamics in CuSbSe2 thin films studied using the OPTP indicate higher μ (∼12 cm2 V−1 s−1), but smaller τ (0.2 ns), and hence similar Ld (∼80 nm). Figure 7 shows the time-domain OPTP spectra for CuSbS2 and CuSbSe2 thin films.

Finally, a summary of the important optoelectronic properties of the chalcostibite CuSbCh2 is shown in Table I, in comparison with the chalcopyrite Cu(In,Ga)Se2 and kesterite Cu2ZnSn(S,Se)4. Overall, both chalcostibites show similar properties to one another but somewhat different if compared to chalcopyrites and kesterites. Despite the higher absorption coefficient shown by the 2D chalcostibite materials, they show larger effective masses of charge carriers and lower photoexcited charge-carrier lifetime. However, the carrier concentrations and the bandgaps of chalcostibites and chalcopyrites/kesterites are quite similar.

TABLE I.

Summary of optoelectronic properties of CuSbCh2 (CAS and CASe), Cu2ZnSn(S,Se)4 (CZTSSe), and Cu(In,Ga)Se2 (CIGS). p = Hole concentration, and μh = majority-carrier mobility measured by the Hall effect. m* = “e” or “h” effective masses, m0 = electron rest mass (9.1 × 10−31 kg). τ = minority-carrier lifetime measured by #transient optical-pump THz probe (OPTP) reflection spectroscopy or $time-resolved photoluminescence (TRPL). α = absorption coefficient for photon energy 0.5 eV higher than the bandgap energy (Eg).

Material p (cm−3) m* (m0) μh (cm2 V−1 s−1) τ (ns) α (cm−1) Eg (eV)
CAS  1016–1020  m*e = 2.9  0.1–49  #0.5–1.3  > 105  1.4–1.6 
  (Ref. 41 m*h = 3.7 (Ref. 41 (Ref. 35 (Ref. 33 (Ref. 44 (Ref. 44
CASe  1016–1018  m*e = 2.5  12  #0.2–1.3  > 105  1.1–1.2 
  (Ref. 21 (Ref. 21 (Ref. 21 (Ref. 21 (Refs. 19 and 21 (Refs. 19 and 21
CZTSSe  1015–1016  m*e = 0.07–0.18  $1–10  > 104  1.0–1.6 
  (Ref. 50 m*h = 0.2–2 (Ref. 51 (Ref. 50 (Ref. 50 (Ref. 51 (Ref. 51
CIGS  1014–1017  m*e = 0.26  1–100  $64  > 104  1.0−1.6 
  (Ref. 52 (Ref. 53 (Ref. 53 (Ref. 54 (Ref. 52 (Ref. 52
Material p (cm−3) m* (m0) μh (cm2 V−1 s−1) τ (ns) α (cm−1) Eg (eV)
CAS  1016–1020  m*e = 2.9  0.1–49  #0.5–1.3  > 105  1.4–1.6 
  (Ref. 41 m*h = 3.7 (Ref. 41 (Ref. 35 (Ref. 33 (Ref. 44 (Ref. 44
CASe  1016–1018  m*e = 2.5  12  #0.2–1.3  > 105  1.1–1.2 
  (Ref. 21 (Ref. 21 (Ref. 21 (Ref. 21 (Refs. 19 and 21 (Refs. 19 and 21
CZTSSe  1015–1016  m*e = 0.07–0.18  $1–10  > 104  1.0–1.6 
  (Ref. 50 m*h = 0.2–2 (Ref. 51 (Ref. 50 (Ref. 50 (Ref. 51 (Ref. 51
CIGS  1014–1017  m*e = 0.26  1–100  $64  > 104  1.0−1.6 
  (Ref. 52 (Ref. 53 (Ref. 53 (Ref. 54 (Ref. 52 (Ref. 52

Advancing from intrinsic properties of the CuSbCh2 material to CuSbCh2 device integration, the development of suitable front contacts (also known as “buffers”) and back contacts is an important direction of research. This is related to the anisotropic properties of the layered CuSbCh2 structure, which makes the surface energies, densities of surface states, and energy band positions [ionization potentials (IP) and electron affinities (χ)] different than in CIGS and related absorbers. First-principles calculations indicate that CuSbS2 has two lower-energy surfaces: the (001)-plane with a surface energy of 12.4 meV Å−1 [(a) and (b) planes in Fig. 1(c)] and the (010)-plane with 14.6 meV Å−1. As shown in Fig. 8(a), the (001) surface has a lower density of surface states than the (010) surface, which is consistent with the weak bonding of the (001) planes. Figure 8(b) shows the calculated energy band positions for [001]- and [010]-oriented CuSbS2 absorber surfaces compared to the commonly used CdS contact layer. These results suggest a cliff-type CuSbS2/CdS band offset of 0.85–1.43 eV, depending on the surface orientation.41 

FIG. 8.

First-principles calculated (a) density of states (DOS) of CuSbS2 surface slabs, normalized by the surface area and offset such that the Fermi level is at zero, and (b) expected band offsets for the lowest-energy [00l]- and [010]-CuSbS2 surfaces with various CdS surfaces. The yellow areas in (a) are the surface states, whereas the red and blue lines are projections onto the surface and bulk, respectively. These figures were adapted from A. W. Welch et al., Prog. Photovoltaics Res. Appl. 24, 929 (2015). Copyright 2015 John Wiley & Sons, Ltd.

FIG. 8.

First-principles calculated (a) density of states (DOS) of CuSbS2 surface slabs, normalized by the surface area and offset such that the Fermi level is at zero, and (b) expected band offsets for the lowest-energy [00l]- and [010]-CuSbS2 surfaces with various CdS surfaces. The yellow areas in (a) are the surface states, whereas the red and blue lines are projections onto the surface and bulk, respectively. These figures were adapted from A. W. Welch et al., Prog. Photovoltaics Res. Appl. 24, 929 (2015). Copyright 2015 John Wiley & Sons, Ltd.

Close modal

Experimental measurements of IP and work function (Φ) on CAS thin films have been performed by X-ray photoemission spectroscopy (XPS)48 and ultraviolet photoemission spectroscopy (UPS).35 XPS for the as-deposited and thermally treated CuSbS2 is shown in Fig. 9(a). The IP of 4.98/5.25 eV (XPS/UPS) and the Φ of 4.73/4.86 eV (XPS/UPS) were found. Combining these results with the experimentally measured direct/indirect bandgap energy (1.58 eV/1.48 eV41) of the CuSbS2 and the estimated band position of the CdS buffer (IP of 7.1 eV vs. vacuum level and bandgap of 2.6 eV55), the CBM of this absorber is expected to be 0.7–1.1 eV higher than that of CdS. These experimental results (Fig. 9) are consistent with the theoretical results (Fig. 8) and suggest the presence of interfacial recombination centers.56 Temperature-dependent J–V measurements (JV–T) performed on CuSbS2/CdS devices [Fig. 9(b)] showed an open-circuit voltage (VOC) extrapolated to 0 K of ∼0.7 V, also indicative of a 0.7–0.9 eV cliff-type band offset at the CuSbS2/CdS interface.44 In contrast, the Fermi level and VBM of CuSbSe2 measured by UPS are −4.63 eV and −4.88 eV with respect to vacuum level (0 eV), respectively.26 Thus, CuSbSe2 may have a less detrimental CB offset with CdS, compared to CuSbS2.

FIG. 9.

(a) Fittings of the valence band maximum (VBM) and secondary electron cutoff (SEC) from XPS spectra for the as-deposited and thermally treated CuSbS2 films, before and after surface cleaning. The Fermi level is at 0 eV. (b) Short-circuit current density (JSC) and open-circuit voltage (VOC) data from temperature-dependent J-V measurements on a CuSbS2/CdS device. Figures (a) and (b) were adapted from T. J. Whittles et al., ACS Appl. Mater. Interfaces 9, 41916 (2017). Copyright 2017 American Chemical Society and F.W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 American Michaelis Society, respectively.

FIG. 9.

(a) Fittings of the valence band maximum (VBM) and secondary electron cutoff (SEC) from XPS spectra for the as-deposited and thermally treated CuSbS2 films, before and after surface cleaning. The Fermi level is at 0 eV. (b) Short-circuit current density (JSC) and open-circuit voltage (VOC) data from temperature-dependent J-V measurements on a CuSbS2/CdS device. Figures (a) and (b) were adapted from T. J. Whittles et al., ACS Appl. Mater. Interfaces 9, 41916 (2017). Copyright 2017 American Chemical Society and F.W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 American Michaelis Society, respectively.

Close modal

Motivated by the high CB offset between the CuSbCh2 absorber and the commonly used CdS contact, some research has been directed at alternative buffer layers, such as undoped and Ga-doped Cd1−xZnxS deposited by atomic layer deposition.57 The CuSbS2 devices made with Zn-rich undoped buffers (Cd0.61Zn0.39S and Cd0.14Zn0.86S) showed high series resistance and very low photoresponse, but 2.4%–5.5% Ga doping in Cd0.6Zn0.4S partially addressed these problems. Comparing the performance of the standard chemical-bath-deposited CdS, VOC improved from 211 mV to 449 mV, the short-circuit current density (JSC) improved from 3.82 to 6.24 mA cm−2, and the efficiency increased from ∼0.3% to 1%. Addressing the issue of Cd toxicity, the CdS/ZnO layers were replaced with GaN/In0.15Ga0.85N in CAS devices, achieving the efficiency of ∼3% and the highest value of JSC (∼34 mA cm−2) reported in the literature,31 which is above the theoretical limit for this 1.5 eV bandgap material.

Different back-contact materials have been also evaluated for CAS PV devices, such as Au, W, Ni, Pd, Pt, FTO, and Mo.41 Many of these back contacts did not result in functioning PV devices due to delamination, pinholes, interface reaction, or other reasons. Among these back contacts, Mo provided the best current collection, even though its work function (4.35–4.90 eV58) was not favorable for charge extraction from the CuSbS2 absorber. Thus, for improving the back contact, the effect of the addition of charge-selective layers (CSLs) on Mo was also studied.41 The MoOx CSL-based device showed a significant increase in efficiency (from 0.49% to 0.86%) because of the improvement in JSC (from 3.53 to 8.91 mA cm−2). This effect on JSC was explained by the deeper (6.6 eV59) work function of the MoOx layer, which promotes an upward band bending in the absorber, reflecting the photogenerated electrons.41 A similar device with a sulfurized hybrid ink/spin-coated CuSbS2 thin-film absorber has shown a record efficiency of 3.22% and JSC of 15 mA cm−2 (see Table II), albeit without statistical histograms or certification results reported in the paper.20 This high performance may be related to the carbon-containing layer at the Mo/CuSbS2 interface measured by Auger electron spectroscopy (AES) depth profiling, which may also aid charge selection.

TABLE II.

Summary of CuSbS2(CAS) and CuSbSe2(CASe) photovoltaic devices with different architectures and for different methods of absorber growth.

No Device architecture VOC (mV) JSC (mA cm−2) FF Efficiency (%)
Mo/C-rich CAS/£,  $bCAS/CdS/ZnO/ZnO:Al/Al20   470  15.64  0.44  3.22a 
Mo/#CAS/CdS/ZnO/ZnO:Al/Al41   330  3.53  0.41  0.49 
Mo/MoOx/#CAS/CdS/ZnO/ZnO:Al/Al41   309  8.91  0.31  0.86 
Mo/#,  $cCAS/CdS/ZnO/ZnO:Al/Al44   350  5.20  0.55  1.0 
FTO/£,  $aCAS/CdS/ZnO/ZnO:Al/Au35   440  3.65  0.31  0.5 
hv→FTO/CdS/i-Sb2S3/¢,  $aCAS/C/Ag24   405  7.54  0.32  1.0 
Mo/@,  $bCAS/CdS/ZnO:Al  490  14.73  0.44  3.13b 
Mo/#,  $aCAS/CdS/GaN/ITO30   104  1.29  0.26  0.04 
Mo/#,  $cCAS/CdS/In0.3Ga0.7 N/ITO30   251  8.58  0.31  0.76 
10  Mo/TiN/#,  $cCAS/GaN/In0.15Ga0.85 N/ITO31   295  33.78c  0.30  2.99 
11  Mo/£,  $aCA(S0.08,Se1.92)/CdS/ZnO/ZnO:Al/Au42   360  20.52  0.37  2.70 
12  Mo/#CASe/CdS/ZnO/ZnO:Al/Ni-Al21   336  26.0  0.53  4.70 
13  Mo/Na-doped #CASe/CdS/ZnO/ZnO:Al/Ni-Al21   394  19.0  0.57  4.28 
14  Mo/#CASe/CdS/ZnO/ZnO:Al/Ni-Al21   346  20.5  0.44  3.5 
15  FTO/£,  $aCASe/CdS/ZnO/ITO/Al26   274  11.84  0.40  1.32 
No Device architecture VOC (mV) JSC (mA cm−2) FF Efficiency (%)
Mo/C-rich CAS/£,  $bCAS/CdS/ZnO/ZnO:Al/Al20   470  15.64  0.44  3.22a 
Mo/#CAS/CdS/ZnO/ZnO:Al/Al41   330  3.53  0.41  0.49 
Mo/MoOx/#CAS/CdS/ZnO/ZnO:Al/Al41   309  8.91  0.31  0.86 
Mo/#,  $cCAS/CdS/ZnO/ZnO:Al/Al44   350  5.20  0.55  1.0 
FTO/£,  $aCAS/CdS/ZnO/ZnO:Al/Au35   440  3.65  0.31  0.5 
hv→FTO/CdS/i-Sb2S3/¢,  $aCAS/C/Ag24   405  7.54  0.32  1.0 
Mo/@,  $bCAS/CdS/ZnO:Al  490  14.73  0.44  3.13b 
Mo/#,  $aCAS/CdS/GaN/ITO30   104  1.29  0.26  0.04 
Mo/#,  $cCAS/CdS/In0.3Ga0.7 N/ITO30   251  8.58  0.31  0.76 
10  Mo/TiN/#,  $cCAS/GaN/In0.15Ga0.85 N/ITO31   295  33.78c  0.30  2.99 
11  Mo/£,  $aCA(S0.08,Se1.92)/CdS/ZnO/ZnO:Al/Au42   360  20.52  0.37  2.70 
12  Mo/#CASe/CdS/ZnO/ZnO:Al/Ni-Al21   336  26.0  0.53  4.70 
13  Mo/Na-doped #CASe/CdS/ZnO/ZnO:Al/Ni-Al21   394  19.0  0.57  4.28 
14  Mo/#CASe/CdS/ZnO/ZnO:Al/Ni-Al21   346  20.5  0.44  3.5 
15  FTO/£,  $aCASe/CdS/ZnO/ITO/Al26   274  11.84  0.40  1.32 
a

Efficiency without statistical histograms.

b

Details of the device characterization are still lacking.

c

Value above the theoretical limit for this 1.5 eV bandgap material. The absorbers were fabricated by (#) co-sputtering, (£) solution-processed/spin-coating, or (@) electrodeposition of metallic stack. $. The absorbers were thermally treated under (a) low vacuum, (b) H2S(g)/S(vapor), or (c) Sb2Se3(vapor) atmosphere. ¢. CBD of Sb2S3 and thermal evaporation of Cu. The “hv→” symbol indicated the direction of incident light in the superstrate configuration.

For the most part, CuSbCh2 research has so far focused on the intrinsic optoelectronic properties of the materials (e.g., absorption coefficient, bandgap, carrier concentration). However, significant challenges arise on integrating absorber thin films into PV device prototypes,19 such as the choice of adequate back/front contacts, growth of pinhole-free absorber layers, optimization of absorber thickness, and possible chemical reactions or inter-diffusion of the absorber with contacts. All of these parameters are very time-consuming to evaluate using the traditional experimental approach. To address this challenge, high-throughput experimental (HTE) combinatorial research methods60,61 can be extended from materials studies62,63 to accelerate PV device research.64,65

The first attempt to apply the HTE approach to CuSbCh2 devices has focused on studying the effects of the crystallographic orientation, phase purity, composition, morphology, and thickness of the CuSbS2 absorber on the PV performance of devices with the SiO2/Mo/CuSbS2/CdS/i-ZnO/Al-ZnO/Al architecture and on evaluating different back contacts.41 In this particular research, a <1%-efficient device was obtained with the optimal absorber thickness of 0.8 μm and MoOX CSL (device #3 in Table II).41 Further, an improvement in CuSbS2 PV device performance (1%) and reproducibility was achieved after absorber thermochemical treatment (TT) under Sb2S3 vapor followed by selective KOH chemical etch of a likely Sb2S3 surface layer, prior to the deposition of front contacts (device #4 in Table II).44 

A similar HTE approach was also applied in the accelerated development of 3%–5%-efficient CuSbSe2 devices (device #12 in Table II) and to understand the tradeoffs of light absorption and charge transport in these layered materials.19,21 In an effort to suppress the tradeoff between JSC and VOC observed in this drift PV device, the Na incorporation in the absorber was investigated. The Na incorporation has enhanced VOC by ∼50 mV (comparing device #13 with #12 in Table II), but it also caused a decrease in JSC and efficiency.21 Several other adjustments in the device fabrication process were attempted, compared to the baseline device (#14 in Table II).19 Some improvement of the efficiency resulted from using an MgF anti-reflection coating and from better design of the front-contact collector grids.21 

Comparing the performances of the CuSbS2 and CuSbSe2 devices with similar structure (e.g., devices #4 and #12 in Fig. 10 and Table II), it is seen that VOC for both materials is 300–400 mV [Fig. 10(a)], which is very low compared to their bandgaps (1.1–1.5 eV). This effect may be directly related to the low quasi Fermi-level splitting due to insufficient absorber quality, and to large CB offset at the interface between these absorbers and the commonly used CdS front contacts. However, the CuSbSe2 devices have shown higher efficiencies, mainly caused by higher JSC values. The higher JSC may be related to the better overlap of the CuSbSe2 absorption with the solar spectrum indicated by better external quantum efficiency (EQE) [Fig. 10(b)] and by lower recombination of the photogenerated charge carriers at the front interface. To represent this comparison, Fig. 10 shows comparative graphs of current density-voltage (J-V) and EQE for these CuSbCh2 PV devices. The poor carrier collection for photon energies near the band edge can be observed in both CuSbS2 and CuSbSe2 EQE graphs [Fig. 10(b)], which is mainly indicative of the short transport lengths of the photogenerated charge carriers.

FIG. 10.

(a) J–V curves under simulated AM1.5G illumination (100 mW cm2) at 25 °C and (b) external quantum efficiency (EQE) for the CuSbCh2 PV device. Data for figures (a) and (b) were adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 American Chemical Society (CuSbS2) and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (CuSbSe2).

FIG. 10.

(a) J–V curves under simulated AM1.5G illumination (100 mW cm2) at 25 °C and (b) external quantum efficiency (EQE) for the CuSbCh2 PV device. Data for figures (a) and (b) were adapted from F. W. de Souza Lucas et al., J. Mater. Chem. A 5, 21986 (2017). Copyright 2017 American Chemical Society (CuSbS2) and A. W. Welch et al., Adv. Energy Mater. 7, 1601935 (2017). Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (CuSbSe2).

Close modal

In addition to the traditional substrate architecture (CIGS-like) of the PV devices described earlier, several other devices in substrate architectures and one in superstrate architectures (CdTe-like) have been attempted. These diverse CuSbCh2 PV device configurations and performances are summarized in Table II. Most of the CuSbCh2 PV device prototypes fabricated in the substrate configuration have the glass/back-contact/CuSbCh2/buffer/TCO/front-contact architecture, where TCO is a transparent conductive oxide such as Al:ZnO or Sn:In2O3. Only one of the PV devices has a superstrate configuration with the glass/TCO/CuSbCh2/buffer/back-contact architecture (i.e., device #6 in Table II), where TCO is F:SnO2. So far, substrate devices had higher efficiencies than this superstrate device, further supporting the analogy of CuSbCh2 materials with CIGS. However, in both cases, more research and development toward higher-quality absorbers and suitable contacts is needed to improve the efficiencies of CuSbCh2 PV devices.

This article provides a research update on the emerging chalcostibite family of absorber materials for thin-film photovoltaic solar cells, with particular focus on the most recent research efforts (past 3 years). The CuSbCh2 (Ch=S, Se) features layered the 2D-like chalcostibite crystallographic structure, in contrast to the 3D-like chalcopyrite structure of CuInSe2 and CuGaSe2. Because of the narrow composition phase width of CuSbCh2, these absorber materials require a self-regulated adsorption-limited synthesis method to avoid detrimental Cu-rich competing phases such as Cu12Sb4S3 or Cu2Se. The band structure and optoelectronic properties of CuSbCh2 feature higher density of states and optical absorption, but heavier effective masses and worse charge transport, compared to CIGS. The moderate hole density is set by compensation between copper vacancies and interstitials, whereas the charge recombination and carrier dynamics are determined by chalcogen vacancies and Cu-on-Sb antisites. The PV-relevant optoelectronic properties of CuSbS2 and CuSbSe2 are summarized in Table I, where they are also compared with the more studied CIGS and CZTS absorbers.

Both CuSbS2 and CuSbSe2 have large cliff-like interface band offsets with CdS, calling for further development of the chalcostibite contacts. The chalcostibite photovoltaic device efficiencies are currently limited to 1%–3% for CuSbS2 and 3%–5% for CuSbSe2 by short transport lengths of the photogenerated charge carriers and by detrimental voltage and current effects of the contact band offset. Hence, more research is needed on the chalcostibite absorber quality and contact selection. Table II summarizes device architectures and the resulting device performances of different CuSbCh2 solar cells reported in the literature. In summary, chalcostibite CuSbCh2 materials are a promising non-toxic and Earth-abundant family of absorbers for applications in thin-film photovoltaic solar cell devices.

The CuSbCh2 work at the National Renewable Energy Laboratory (NREL) managed and operated by Alliance for Sustainable Energy, LLC, was supported by the U.S. Department of Energy, Office of Energy Efficiency and Renewable Energy, Solar Energy Technologies Office, under Contract No. DE-AC36-08GO28308 to NREL. The views expressed in the article do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes. F.W.S.L. was funded by the São Paulo Research Foundation (FAPESP), Grant Nos. 2016/10513-3 and 2014/12166-3.

1.
See https://www.nrel.gov/pv/assets/images/efficiency-chart.png for Best research-cell efficiencies (last accessed
January 2018
).
2.
C.
Wadia
,
A. P.
Alivisatos
, and
D. M.
Kammen
,
Environ. Sci. Technol.
43
,
2072
(
2009
).
3.
M.
Yang
,
A.
Zakutayev
,
J.
Vidal
,
X.
Zhang
,
D. S.
Ginley
, and
F. J.
DiSalvo
,
Energy Environ. Sci.
6
,
2994
(
2013
).
4.
T.
Minami
,
Y.
Nishi
, and
T.
Miyata
,
Appl. Phys. Express
8
,
22301
(
2015
).
5.
Y.
Zhou
,
L.
Wang
,
S.
Chen
,
S.
Qin
,
X.
Liu
,
J.
Chen
,
D. J.
Xue
,
M.
Luo
,
Y.
Cao
,
Y.
Cheng
,
E. H.
Sargent
, and
J.
Tang
,
Nat. Photonics
9
,
409
(
2015
).
6.
C. M.
Caskey
,
R. M.
Richards
,
D. S.
Ginley
, and
A.
Zakutayev
,
Mater. Horiz.
1
,
424
(
2014
).
7.
A.
Zakutayev
,
C. M.
Caskey
,
A. N.
Fioretti
,
D. S.
Ginley
,
J.
Vidal
,
V.
Stevanovic
,
E.
Tea
, and
S.
Lany
,
J. Phys. Chem. Lett.
5
,
1117
(
2014
).
8.
V.
Steinmann
,
R.
Jaramillo
,
K.
Hartman
,
R.
Chakraborty
,
R. E.
Brandt
,
J. R.
Poindexter
,
Y. S.
Lee
,
L.
Sun
,
A.
Polizzotti
,
H. H.
Park
,
R. G.
Gordon
, and
T.
Buonassisi
,
Adv. Mater.
26
,
7488
(
2014
).
9.
A.
Sánchez-Juárez
,
A.
Tiburcio-Silver
, and
A.
Ortiz
,
Thin Solid Films
480-481
,
452
(
2005
).
10.
L.
Lahourcade
,
N. C.
Coronel
,
K. T.
Delaney
,
S. K.
Shukla
,
N. A.
Spaldin
, and
H. A.
Atwater
,
Adv. Mater.
25
,
2562
(
2013
).
11.
T.
Gershon
,
B.
Shin
,
N.
Bojarczuk
,
M.
Hopstaken
,
D. B.
Mitzi
, and
S.
Guha
,
Adv. Energy Mater.
5
,
1400849
(
2015
).
12.
K.
Ramasamy
,
M. A.
Malik
, and
P.
O’Brien
,
Chem. Sci.
2
,
1170
(
2011
).
13.
R.
Baier
,
C.
Leendertz
,
D.
Abou-Ras
,
M. C.
Lux-Steiner
, and
S.
Sadewasser
,
Sol. Energy Mater. Sol. Cells
130
,
124
(
2014
).
14.
M. S.
Leite
,
M.
Abashin
,
H. J.
Lezec
,
A.
Gianfrancesco
,
A. A.
Talin
, and
N. B.
Zhitenev
,
ACS Nano
8
,
11883
(
2014
).
15.
M. B.
Costa
,
F. W.
de Souza Lucas
, and
L. H.
Mascaro
,
ChemElectroChem
4
,
2507
(
2017
).
16.
C.
Tablero
,
J. Alloys Compd.
678
,
18
(
2016
).
17.
A.
Zakutayev
,
L. L.
Baranowski
,
A. W.
Welch
,
C. A.
Wolden
, and
E. S.
Toberer
, in
2014 IEEE 40th Photovoltaic Specialist Conference (PVSC)
(
IEEE
,
2014
), pp.
2436
2438
.
18.
A.
Gassoumi
,
M.
Musa Saad H.-E.
,
S.
Alfaify
,
T.
Ben Nasr
, and
N.
Bouarissa
,
J. Alloys Compd.
725
,
181
(
2017
).
19.
A. W.
Welch
,
L. L.
Baranowski
,
P.
Zawadzki
,
S.
Lany
,
C. A.
Wolden
, and
A.
Zakutayev
,
Appl. Phys. Express
8
,
82301
(
2015
).
20.
S.
Banu
,
S. J.
Ahn
,
S. K.
Ahn
,
K.
Yoon
, and
A.
Cho
,
Sol. Energy Mater. Sol. Cells
151
,
14
(
2016
).
21.
A. W.
Welch
,
L. L.
Baranowski
,
H.
Peng
,
H.
Hempel
,
R.
Eichberger
,
T.
Unold
,
S.
Lany
,
C.
Wolden
, and
A.
Zakutayev
,
Adv. Energy Mater.
7
,
1601935
(
2017
).
22.
A. W.
Welch
,
P. P.
Zawadzki
,
S.
Lany
,
C. A.
Wolden
, and
A.
Zakutayev
,
Sol. Energy Mater. Sol. Cells
132
,
499
(
2014
).
23.
L.
Shi
,
Y.
Li
,
C.
Wu
, and
Y.
Dai
,
J. Alloys Compd.
648
,
507
(
2015
).
24.
B.
Krishnan
,
S.
Shaji
, and
R.
Ernesto Ornelas
,
J. Mater. Sci.: Mater. Electron.
26
,
4770
(
2015
).
25.
L. L.
Baranowski
,
P.
Zawadzki
,
S.
Lany
,
E. S.
Toberer
, and
A.
Zakutayev
,
Semicond. Sci. Technol.
31
,
123004
(
2016
).
26.
D. J.
Xue
,
B.
Yang
,
Z. K.
Yuan
,
G.
Wang
,
X.
Liu
,
Y.
Zhou
,
L.
Hu
,
D.
Pan
,
S.
Chen
, and
J.
Tang
,
Adv. Energy Mater.
5
,
1501203
(
2015
).
27.
L.
Yu
,
R. S.
Kokenyesi
,
D. A.
Keszler
, and
A.
Zunger
,
Adv. Energy Mater.
3
,
43
(
2013
).
28.
T.
Maeda
and
T.
Wada
,
Thin Solid Films
582
,
401
(
2015
).
29.
A.
Rabhi
,
M.
Kanzari
, and
B.
Rezig
,
Thin Solid Films
517
,
2477
(
2009
).
30.
W. L.
Chen
,
D. H.
Kuo
, and
T. T. A.
Tuan
,
J. Electron. Mater.
45
,
688
(
2015
).
31.
A. D.
Saragih
,
D. H.
Kuo
, and
T. T. A.
Tuan
,
J. Mater. Sci.: Mater. Electron.
28
,
2996
(
2017
).
32.
C.
Macías
,
S.
Lugo
,
Á.
Benítez
,
I.
López
,
B.
Kharissov
,
A.
Vázquez
, and
Y.
Peña
,
Mater. Res. Bull.
87
,
161
(
2017
).
33.
M. E.
Edley
,
B.
Opasanont
,
J. T.
Conley
,
H.
Tran
,
S. Y.
Smolin
,
S.
Li
,
A. D.
Dillon
,
A. T.
Fafarman
, and
J. B.
Baxter
,
Thin Solid Films
646
,
180
(
2018
).
34.
J. A.
Ramos Aquino
,
D. L.
Rodriguez Vela
,
S.
Shaji
,
D. A.
Avellaneda
, and
B.
Krishnan
,
Phys. Status Solidi C
13
,
24
(
2016
).
35.
B.
Yang
,
L.
Wang
,
J.
Han
,
Y.
Zhou
,
H.
Song
,
S.
Chen
,
J.
Zhong
,
L.
Lv
,
D.
Niu
, and
J.
Tang
,
Chem. Mater.
26
,
3135
(
2014
).
36.
S.
Suehiro
,
K.
Horita
,
M.
Yuasa
,
T.
Tanaka
,
K.
Fujita
,
Y.
Ishiwata
,
K.
Shimanoe
, and
T.
Kida
,
Inorg. Chem.
54
,
7840
(
2015
).
37.
L.
Zhang
,
Y.
Li
,
X.
Li
,
C.
Li
,
R.
Zhang
,
J. J.
Delaunay
, and
H.
Zhu
,
Nano Energy
28
,
135
(
2016
).
38.
D.
Tang
,
J.
Yang
,
F.
Liu
,
Y.
Lai
,
J.
Li
, and
Y.
Liu
,
Electrochim. Acta
76
,
480
(
2012
).
39.
H.
Su
,
Solid State Ionics
123
,
319
(
1999
).
40.
C.
An
,
Q.
Liu
,
K.
Tang
,
Q.
Yang
,
X.
Chen
,
J.
Liu
, and
Y.
Qian
,
J. Cryst. Growth
256
,
128
(
2003
).
41.
A. W.
Welch
,
L. L.
Baranowski
,
P.
Zawadzki
,
C.
DeHart
,
S.
Johnston
,
S.
Lany
,
C. A.
Wolden
, and
A.
Zakutayev
,
Prog. Photovoltaics Res. Appl.
24
,
929
(
2016
).
42.
B.
Yang
,
C.
Wang
,
Z.
Yuan
,
S.
Chen
,
Y.
He
,
H.
Song
,
R.
Ding
,
Y.
Zhao
, and
J.
Tang
,
Sol. Energy Mater. Sol. Cells
168
,
112
(
2017
).
43.
F. W.
De Souza Lucas
,
A. W.
Welch
,
L. L.
Baranowski
,
P. C.
Dippo
,
L. H.
Mascaro
, and
A.
Zakutayev
, in
2015 IEEE 42nd Photovoltaic Specialist Conference (PVSC)
(
IEEE
,
2015
), pp.
1
5
.
44.
F. W.
de Souza Lucas
,
A. W.
Welch
,
L. L.
Baranowski
,
P. C.
Dippo
,
H.
Hempel
,
T.
Unold
,
R.
Eichberger
,
B.
Blank
,
U.
Rau
,
L. H.
Mascaro
, and
A.
Zakutayev
,
J. Phys. Chem. C
120
,
18377
(
2016
).
45.
X.
Song
,
X.
Ji
,
M.
Li
,
W.
Lin
,
X.
Luo
, and
H.
Zhang
,
Int. J. Photoenergy
2014
,
1
.
46.
See http://www.fiz-karlsruhe.de/icsd.html for Inorganic Crystal Structure Database (last accessed
January 2018
).
47.
T.
Rath
,
A. J.
MacLachlan
,
M. D.
Brown
, and
S. A.
Haque
,
J. Mater. Chem. A
3
,
24155
(
2015
).
48.
T. J.
Whittles
,
T. D.
Veal
,
C. N.
Savory
,
A. W.
Welch
,
F. W.
De Souza Lucas
,
J. T.
Gibbon
,
M.
Birkett
,
R. J.
Potter
,
D. O.
Scanlon
,
A.
Zakutayev
, and
V. R.
Dhanak
,
ACS Appl. Mater. Interfaces
9
,
41916
(
2017
).
49.
F. W.
de Souza Lucas
,
H.
Peng
,
S.
Johnston
,
P. C.
Dippo
,
S.
Lany
,
L. H.
Mascaro
, and
A.
Zakutayev
,
J. Mater. Chem. A
5
,
21986
(
2017
).
50.
T.
Gokmen
,
O.
Gunawan
, and
D. B.
Mitzi
,
Appl. Phys. Lett.
105
,
33903
(
2014
).
51.
U.
Saha
and
M. K.
Alam
,
RSC Adv.
7
,
4806
(
2017
).
52.
U.
Rau
, “
Electronic properties of Cu(In,Ga)Se2 thin-film solar cells: An update
,” in
Advances in Solid State Physics
, edited by
B.
Kramer
(
Springer
,
Berlin, Heidelberg
,
2004
), Vol. 44.
53.
See www.materials.nrel.gov for Effective masses and other electronic structure parameters (last accessed
December 2017
).
54.
J. V.
Li
,
S.
Grover
,
M. A.
Contreras
,
K.
Ramanathan
,
D.
Kuciauskas
, and
R.
Noufi
,
Sol. Energy Mater. Sol. Cells
124
,
143
(
2014
).
55.
S. M.
Park
,
T. G.
Kim
,
Y. D.
Chung
,
D.-H.
Cho
,
J.
Kim
,
K. J.
Kim
,
Y.
Yi
, and
J. W.
Kim
,
J. Phys. D: Appl. Phys.
47
,
345302
(
2014
).
56.
T.
Minemoto
,
T.
Matsui
,
H.
Takakura
,
Y.
Hamakawa
,
T.
Negami
,
Y.
Hashimoto
,
T.
Uenoyama
, and
M.
Kitagawa
,
Sol. Energy Mater. Sol. Cells
67
,
83
(
2001
).
57.
L. L.
Baranowski
,
S.
Christensen
,
A. W.
Welch
,
S.
Lany
,
M.
Young
,
E. S.
Toberer
, and
A.
Zakutayev
,
Mater. Chem. Front.
1
,
1342
(
2017
).
58.
D. L.
Jacobson
and
A. E.
Campbell
,
Metall. Mater. Trans. B
2
,
3063
(
1971
).
59.
C.
Battaglia
,
X.
Yin
,
M.
Zheng
,
I. D.
Sharp
,
T.
Chen
,
S.
McDonnell
,
A.
Azcatl
,
C.
Carraro
,
B.
Ma
,
R.
Maboudian
,
R. M.
Wallace
, and
A.
Javey
,
Nano Lett.
14
,
967
(
2014
).
60.
M. L.
Green
,
I.
Takeuchi
, and
J. R.
Hattrick-Simpers
,
J. Appl. Phys.
113
,
231101
(
2013
).
61.
M. L.
Green
,
C. L.
Choi
,
J. R.
Hattrick-Simpers
,
A. M.
Joshi
,
I.
Takeuchi
,
S. C.
Barron
,
E.
Campo
,
T.
Chiang
,
S.
Empedocles
,
J. M.
Gregoire
,
A. G.
Kusne
,
J.
Martin
,
A.
Mehta
,
K.
Persson
,
Z.
Trautt
,
J.
Van Duren
, and
A.
Zakutayev
,
Appl. Phys. Rev.
4
,
11105
(
2017
).
62.
L. L.
Baranowski
,
P.
Zawadzki
,
S.
Christensen
,
D.
Nordlund
,
S.
Lany
,
A. C.
Tamboli
,
L.
Gedvilas
,
D. S.
Ginley
,
W.
Tumas
,
E. S.
Toberer
,
A.
Zakutayev
,
S.
David
,
W.
Tumas
,
E. S.
Toberer
, and
A.
Zakutayev
,
Chem. Mater.
26
,
4951
(
2014
).
63.
P. P.
Rajbhandari
,
A.
Bikowski
,
J. D.
Perkins
,
T. P.
Dhakal
, and
A.
Zakutayev
,
Sol. Energy Mater. Sol. Cells
159
,
219
(
2017
).
64.
S.
Siol
,
T. P.
Dhakal
,
G. S.
Gudavalli
,
P. P.
Rajbhandari
,
C.
Dehart
,
L. L.
Baranowski
, and
A.
Zakutayev
,
ACS Appl. Mater. Interfaces
8
,
14004
(
2016
).
65.
K.
Mokurala
,
L. L.
Baranowski
,
F. W.
De Souza Lucas
,
S.
Siol
,
M. F. A. M.
Van Hest
,
S.
Mallick
,
P.
Bhargava
, and
A.
Zakutayev
,
ACS Comb. Sci.
18
,
583
(
2016
).