Open Access Article
Koichiro Suekuni
*ab,
Mei Yamamotoa,
Susumu Fujii
c,
Pierric Lemoine
d,
Philipp Sauerschnig
e,
Michihiro Ohta
e,
Emmanuel Guilmeau
f and
Michitaka Ohtaki
ab
aInterdisciplinary Graduate School of Engineering Sciences, Kyushu University, Kasuga, Fukuoka 816-8580, Japan. E-mail: suekuni.koichiro.063@m.kyushu-u.ac.jp
bTransdisciplinary Research and Education Center for Green Technologies, Kyushu University, Kasuga, Fukuoka 816-8580, Japan
cDepartment of Materials, Faculty of Engineering, Kyushu University, Motooka, Fukuoka 819-0395, Japan
dUniversité de Lorraine, CNRS, IJL, Nancy, F-54000, France
eGlobal Zero Emission Research Center, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Ibaraki 305-8569, Japan
fCRISMAT, CNRS, Normandie Université, ENSICAEN, UNICAEN, Caen 14000, France
First published on 20th January 2026
A copper-based multicomponent sulphide, Cu26Ti2(Sb,Ge)6S32 colusite, is a promising thermoelectric material. We investigated the effects of sulphur deficiency on the crystal structure, electronic structure, and thermoelectric properties in the series Cu26Ti2Sb4Ge2S32−x. By combining experiments and ab initio calculations, we found that sulphur deficiency induced the formation of interstitial Cu atoms in the sphalerite-like framework of Cu26Ti2Sb4Ge2S32. This resulted in a decrease in the hole carrier concentration and a susbtantial enhancement of ZT up to unity at 673 K. We also fabricated a power generation device composed of the sulphur-deficient colusite, Ni–Sb-based compounds (interface material), and Ni. The maximum conversion efficiency of the device reached 3.2% with a temperature difference of 266 K.
To achieve large-scale, cost-effective, and environmentally friendly TE applications, materials must not only be high-performing but also composed of constituent elements that are non-toxic, environmentally benign, and low-cost. Examples include Mg-based compounds (Mg3(Bi,Sb)2,2–4 MgAgSb,5–7 and Mg2(Si,Sn)8–10), Half-Heusler compounds (MNiSn and MCoSb, with M = Ti, Hf, Zr, and NbFeSb),11–13 and sulphides (Cu-based,14–16 Bi-based,17,18 and Ti-based compounds19–22). Cu-based sulphides have emerged as promising p-type TE materials, and their TE properties have been studied extensively since ∼2010. The worldwide research led to significant advances in TE Cu-based sulphides (e.g., chalcosite (Cu2S),23 digenite (Cu1.8S),24 tetrahedrite (Cu12Sb4S13),25,26 and colusite (Cu26T2M6S32 (T = Ti, V, Nb, Ta, Cr, Mo, W; M = Ge, Sn, Sb))27–33 with ZT reaching ∼0.5–1 at 673 K. However, less effort has been devoted to the fabrication of TE devices/modules34–36 unlike the Mg-based compounds37–39 and Half-Heusler compounds.11,13,40 To maximize the conversion efficiency of a TE module, it is crucial to minimize the electrical and thermal contact resistance between the TE material and the electrodes that connect the devices in series. It is often necessary to insert interface materials between the TE material and the electrode, and identification of the more suitable materials is critically important.
We have recently discovered a colusite, Cu26Ti2Sb6S32, which showed semiconducting properties and low κlat.32 The substitution of Ge4+ for Sb5+ increased the hole carrier concentration, n, leading to the enhancement of S2ρ−1. The combination of large S2ρ−1 and low κlat resulted in a ZT value of 0.9 at 673 K. From our subsequent investigations on Cu26Ti2Sb6−xGexS32, we found that the previously studied samples32 most probably present sulphur deficiency due to inadequate recovery of the sulphur residuals produced during the synthesis (reaction) process. In this study, we therefore investigated how the sulphur deficiency affects the crystal structure, electronic structure, and TE properties through experiments and ab initio calculations.
We then fabricated TE devices using the Cu26Ti2Sb6−xGexS32 colusites. In our previous study,34 we explored diffusion barrier materials from pure metals and reported that a single-leg device of Cu26Nb2Ge6S32 with Au layers (diffusion barrier layers) showed low contact resistance at the Au/colusite interface and a TE conversion efficiency of 3.3% at a temperature difference of ΔT∼270 K. However, an issue arose from the macroscopic diffusion of Au into the colusite matrix. Specifically, the ΔV generated from the device was lower than the ΔV predicted based on the material's properties. Continued efforts are required to address this issue, while the exploration of diffusion barrier/interface materials holds comparable importance. In this study, we selected Ni as a diffusion barrier material. Ni is known to be reluctant to interdiffuse with Ag,41 which is often used as a paste/electrode material. However, direct hot-press bonding between Cu26Ti2Sb6−xGexS32 colusites and Ni was unsuccessful due to a reaction between the materials. Consequently, we explored interface materials to be placed between the colusite and Ni. As effective interface materials often share elements with TE materials (e.g., MgCuSb for MgAgSb, NiTe2 for Bi0.5Sb1.5Te3, and CoAl for CoSb3),42 we selected Sb-based compounds, more specifically the Ni–Sb system (NiSb and NiSb2) with metallic properties,43 as potential candidates of interface materials for the colusite containing Sb.
Ni–Sb compounds (NiSb, NiSb2, and Cu/Co-substituted NiSb: Ni0.9Cu0.1Sb and Ni0.9Co0.1Sb) were synthesized by directly reacting the elements (Ni, 99.9%, powder; Co, 99.9%, powder; Cu, 99.9%, powder; Sb, 99.9999%, powder). The elements were mixed and then molded into a pellet, which was sealed in an evacuated quartz tube. The pellet was heated to 1323 K, maintained at this temperature for 24 h, cooled to 873 K, maintained at this temperature for 100 h, and then cooled to RT.
The powders of Cu26Ti2Sb4Ge2S31.5 (ball-milled), Ni–Sb based compounds, and Ni were placed in a WC die to form 5 layers in the order Ni/Ni–Sb/Cu26Ti2Sb4Ge2S31.5/Ni–Sb/Ni and hot-pressed under the aforementioned conditions to fabricate the TE devices. The obtained pellet was cut and polished into bars for the measurement of power generation properties.
The surface morphologies and chemical compositions of the sintered samples were investigated by scanning electron microscopy (SEM) and energy dispersive X-ray spectroscopy (EDS). SEM and EDS were performed using a microscope (JCM-6000Plus NeoScope, JEOL).
The resistance scanning measurement was performed for the device composed of Cu26Ti2Sb4Ge2S31.5, Ni–Sb-based compounds, and Ni by a four probe AC method using a laboratory-built system with a movable voltage probe. The TE conversion efficiency η of the device was evaluated in a vacuum using a measurement system (Mini-PEM, ADVANCE RIKO). η was calculated as P/(P + Qout), where P is the output power and Qout is the heat released into the low-temperature heat bath through the sample. In this measurement, the hot-side temperature of the device was set at TH = 321–566 K, while keeping the cold-side temperature at TC = 295–300 K.
We performed the calculations for pristine Cu26Ti2Sb6S32, Cu26Ti2Sb6S31□, and Cu27Ti2Sb6S32. The cubic structure (space group P
3n, no. 218) of Cu26Ti2Sb6S32 includes non-equivalent crystallographic sites: three for Cu (6d, 8e, 12f), one for Ti (2a), one for Sb (6c), and two for S (8e, 24i).32,54 One S atom at 8e or 24i was removed from the crystal structure to construct two Cu26Ti2Sb6S31□ models, whereas one additional Cu atom was placed at unoccupied “interstitial” sites (6b, 24i) to construct two Cu27Ti2Sb6S32 models. The electronic band structures and density of states (DOS) were calculated for the five optimized structural models. For the band structure calculations, the reciprocal path for Cu26Ti2Sb6S32, as suggested by SeeK-path,55 was applied to all models. This approach enables direct comparison of the electronic structures of the pristine and defect-containing models, although the defects break the original symmetry of Cu26Ti2Sb6S32. For the DOS calculations, a Γ-centred k-point mesh of 9 × 9 × 9 was employed. Note that, in our calculations, two electrons per f.u. were intentionally removed to adjust the number of hole carriers in these non-Ge-substituted models to those for experimental Ge-substituted composition Cu26Ti2Sb4Ge2S32−x. This procedure is useful to exclude the influence of interactions between Ge and other defects and to facilitate a concise discussion of sulphur deficiency.
The formation energies of a Cu interstitial and an S vacancy were calculated from the energy differences between the defect models and the pristine model, where two electrons were removed from both models, while including the chemical potentials of Cu or S to account for the exchange of these atoms with their reservoirs (i.e., the competing phases). The chemical potentials of Cu and S were determined from the equilibrium conditions in the computed phase diagram.
3n), suggesting single phase samples, but the peaks were accompanied by a small shoulder at the lower angle side (Fig. S1). This result suggests inhomogeneous composition distribution. The peak asymmetry could be reproduced by assuming the existence of two colusite phases with and without Ge. For the x = 0 sample, the chemical compositions/fractions obtained by the Rietveld analysis were Cu26Ti2Sb3.4(1)Ge2.6(1)S32/∼94 wt% and Cu26Ti2Sb6.0(1)Ge0.0(1)S32/∼6 wt%. The former phase has a smaller lattice parameter, a, (10.7157(1) Å) than the latter phase (10.7536(3) Å), which probably results from the smaller ionic radius of Ge4+ compared to Sb5+. Another possibility is that the composition distribution is linked to the formation of sulphur-deficient (Cu-rich) colusite (see Section 3.3). In any case, the amount of the secondary phase is likely to be small. A similar result was obtained for the x = 0.5 sample. Such a shoulder, if it exists, could not be detected in the PXRD patterns for the x = 1.0 and x = 1.5 samples. The Rietveld refinements indicated that the x = 1.0 sample was composed of a single colusite phase, whereas the x = 1.5 sample is composed of colusite and small amount of tetrahedrite (Cu12Sb4S13, <2 wt%). The traces of secondary phases (Ge-poor or sulphur-deficient colusite for x = 0 and x = 0.5, and tetrahedrite for x = 1.5) in the samples should have minor effects on the TE properties.
The most prominent variation with x was an increase in a from 10.714 Å (x = 0, primary colusite phase) to 10.757 Å (x = 1.5). It should be noted that the sulphur deficiency was hard to be confirmed by the Rietveld analysis due to the strong interaction between the site occupation factors and the thermal parameters. Instead, EDS showed that the content of sulphur decreased with x, while those of cations (Cu, Ti, Sb, and Ge) were close to their nominal values (Table S1). It is noteworthy that the value of a for the previously synthesized “Cu26Ti2Sb4Ge2S32” sample32 was between those for the x = 1.0 and x = 1.5 samples of Cu26Ti2Sb4Ge2S32−x (inset of Fig. 1). This fact suggests sulphur deficiency in the previously synthesized sample. The defect species that are preferentially generated (sulphur vacancies or interstitial cations), their crystallographic sites, and the mechanism of lattice expansion were investigated by ab initio calculations (see Section 3.3).
![]() | ||
| Fig. 2 Secondary electron images for (a) polished surfaces of the Cu26Ti2Sb4Ge2S32−x (x = 0, 0.5, 1.0, 1.5) samples and (b) fractured surfaces of the x = 0 and x = 1 samples. | ||
Fig. 3 displays the TE properties for Cu26Ti2Sb4Ge2S32−x (x = 0, 0.5, 1.0, 1.5) and previously synthesized “Cu26Ti2Sb4Ge2S32”.32 The x = 0 sample exhibited metallic behaviour in S and ρ (Fig. 3a and b). The values of S and ρ increased with increasing x, indicating a decrease in n. Indeed, the value of n obtained from Hall effect measurements at RT decreased from 3.5 × 1021 cm−3 (x = 0) to 2.4 × 1021 cm−3 (x = 0.5), 1.8 × 1021 cm−3 (x = 1.0) and 0.9 × 1021 cm−3 (x = 1.5). A similar trend in the electronic properties associated with sulphur deficiency was observed for colusite Cu26Cr2Ge6S32−δ.31 The values of S2ρ−1 for Cu26Ti2Sb4Ge2S32−x were equal to 1.4 mWK−2 m−1 at 673 K for the x = 0–1.0 samples, whereas it decreased to 0.97 mWK−2 m−1 for the x = 1.5 sample (Fig. 3c). The decrease in n with x led to the reduction in the electronic thermal conductivity, κele (Table S2). As a result, the value of κ at 673 K decreased from 1.3 WK−1 m−1 (x = 0) to 0.62 WK−1 m−1 (x = 1.5) (Fig. 3d). The lattice thermal conductivity κlat was estimated by subtracting κele from κ. Here the values of κele were estimated from the Wiedemann–Franz law, κele = LTρ−1, where the Lorentz number L was calculated using an equation, L = 1.5 + exp(−|S|/116), based on a single parabolic band model with acoustic phonon scattering.56 In the lower temperature region, κlat values for the x ≥ 0.5 samples were equivalent and slightly lower than that for the x = 0 sample (Fig. 3e). Because the morphology of the sample (grain size) was similar between x = 0 and x = 1.0, as mentioned above (Fig. 2), the κlat reduction can be attributed to the structural modification due to sulphur deficiency (see Section 3.3). The combination of high S2ρ−1 and low κ led to relatively high ZT, which increased from 0.7 (x = 0) to 0.9 (x = 0.5) and 1.0 (x = 1.0, 1.5) at 673 K (Fig. 3f). The values of ρ, S, κ as well as ZT for the previously synthesized “Cu26Ti2Sb4Ge2S32”32 were between those for x = 1.0 and x = 1.5 (Fig. 3), consistent with the XRD results, as discussed above.
First, we compared the stability of two sulphur vacancies at two different sites. The formation energy of a sulphur vacancy at the 8e site was found to be higher (+1.8 eV) than that at the 24i site, indicating that the 8e-site vacancy is energetically less favourable. Because the 8e site locates in a rigid [TiS4]Cu6 tetrahedral–octahedral complex,30,54 vacancy formation at this site is not preferred. The electronic structure of Cu26Ti2Sb6S31□ with the 24i-site sulphur vacancy exhibits degenerate semiconducting characteristics similar to those of Cu26Ti2Sb6S32 (Fig. S2). This result was contrary to our initial anticipation that a neutral sulphur vacancy would lead to electron doping. The relaxed structure (Fig. S3) and the orbital projected DOS for Sb (Fig. S4) suggest that the neutral sulphur vacancy on 24i site induces the formation of a 5s2 lone pair in Sb, thereby reducing the valence state of Sb from +5 to +3. Considering that Sb atoms are in tetrahedral coordination of S(24i) in colusite, it is reasonable to consider that the lone pair of the Sb3+ cation compensate the sulphur vacancy to form an SbS(24i)3LP polyhedron. Furthermore, only a slight increase in the calculated lattice parameter a (∼0.002 Å) was observed after removing the sulphur atom from the 24i site, which disagrees with the increase of lattice parameter, determined from the XRD analyses (Fig. 1). In addition, the localization of two electrons from the sulphur vacancy prevents any decrease in hole concentration, as evidenced by the unchanged Fermi level relative to the valence band maximum (Fig. S2). This result is also inconsistent with the experimental values of n decreasing with x (Fig. 3), suggesting that sulphur vacancy formation is unlikely under our synthesis conditions.
According to our previous studies,57 sulphur deficiency (volatilization) is compensated by the incorporation of cations into the unoccupied interstitial 24i site of the sphalerite-like framework of colusite. Indeed, an interstitial Cu atom at the 24i site (Fig. 4a) is energetically more favourable than one at the 6b site, with a calculated formation energy lower by 0.17 eV. The 24i-site interstitial caused an increase in the calculated a of 0.030 Å, which is significantly larger than that of the sulphur vacancy model and agrees with the experimental observations (Fig. 1). A Cu-excess composition/structure, if present, explains the geometrical densities exceeding 100% of the theoretical ones (i.e. without interstitial cations, Section 3.2). In addition, the number of hole carriers decreased as the Fermi level shifts toward the band edge to maintain charge balance (Fig. 4b), which is consistent with the measured TE properties (Fig. 3). Therefore, structural modifications due to Cu interstitials probably occurred in the Cu26Ti2Sb4Ge2S32−x (x > 0) samples. Indeed, the formation energy of a Cu interstitial is lower than that of an S vacancy under all equilibrium conditions in the computed phase diagram. For example, under an S-poor condition in equilibrium with Cu12Sb4S13, Cu7S4, and Cu2S, the formation energies are 0.23 eV and 0.53 eV, respectively. This indicates that Cu interstitials are more readily formed than S vacancies in the colusite. It is noteworthy that the value of κlat near the room temperature was reduced at x > 0 (Fig. 3e). This result indicates that the interstitial Cu acts as a phonon scattering center, as reported previously.57,58
We then investigated how chemical substitution in NiSb influences Rc. The devices containing Cu- and Co-substituted NiSb showed no crack near the interfaces between the NiSb-based compounds and the colusite (Fig. S6). The devices with Ni0.9Cu0.1Sb and Ni0.9Co0.1Sb exhibited, respectively, higher Rc (∼30, ∼50 mΩmm2) and lower Rc (∼9 mΩmm2) at the NiSb-based compounds/colusite interfaces compared to the device with NiSb (Fig. 5d). The interfaces were analysed by SEM (Fig. S7). For the devices containing NiSb and Ni0.9Cu0.1Sb, thin NiSbS, Sb2S3, and multiphase layers were formed. Conversely, for the device with Ni0.9Co0.1Sb, only small amounts of (Ni,Co)SbS and Sb2S3 were detected. The NiSbS layer likely forms through the sulphurization of NiSb. Given that NiSbS itself possesses metallic characteristics,59 the formation of the Sb2S3 layer is primarily responsible for the increase in Rc (Fig. 5d). In devices containing Ni0.9Co0.1Sb, the presence of a small amount of Sb2S3 should be responsible for a non-negligible Rc.
The Rc value for the device with Ni0.9Co0.1Sb was ∼9 mΩ mm2 (Fig. 5d). The Rc remained nearly constant after annealing at 573 K for 50 h, but increased significantly as the annealing temperature was elevated to 623 K and 673 K (Fig. S8). After annealing, a thin (Ni,Co)SbS layer formed and its thickness increased with elevating the annealing temperature (Fig. S9). Its metallic characteristics would have limited impact on Rc. For the sample annealed at 673 K, Sb2S3 was clearly visible in a SEM image (Fig. S9). Therefore, the formation of Sb2S3 is primarily responsible for the observed increase in Rc, which is consistent with the claim made above.
For the unannealed device containing Ni0.9Co0.1Sb, the ρ value, estimated from the slope of the resistance scanning data, was 9.7 Ω m at RT. This value was slightly higher than that for x = 0.5 of Cu26Ti2Sb4Ge2S32−x (colusite used for the devise, Cu26Ti2Sb4Ge2S31.5) but was comparable to that for x = 1.0 (Fig. 3b). The increase in ρ for the colusite phase was probably attributed to a decrease in sulphur content due to the reaction between the colusite and Ni0.9Co0.1Sb during the sintering. We therefore compared the experimental power generation properties for the device to simulate results based on the TE properties of x = 1.0 (Fig. S10 and Tables S3, S4).
The power generation properties of the device with Ni0.9Co0.1Sb were investigated with TH reaching up to 573 K, while TC was maintained at ∼300 K (Fig. 5e and S11). The maximum output power, Pmax, obtained from the voltage–current plot increased with increasing temperature difference ΔT and reached 23 mW at ΔT = 266 K (Fig. S11a and b). It should be noted that the open circuit voltage Voc and internal resistance Rin were reversible between the heating and cooling processes (Fig. S11c and d), indicating the device's stability under the current measurement conditions. ηmax = P/(P + Qout) increased with increasing ΔT and reached 3.2% at ΔT = 266 K (Fig. 5e and S11e, f). This value is equivalent to ηmax measured for a Cu26Nb2Ge6S32-based device (3.3%),34 while is lower than that for Cu2ZnSnS4-based single crystals (4%).60 ηmax for an ideal device with the TE properties of x = 1.0, calculated using COMSOL Multiphysics® and a web simulator,61 was 5.7% when TL and TH were set at 300 K and 573 K, respectively (Fig. 5e and Table S4). Because the Voc values were comparable between the experiment and calculation, the reduced ηmax for the fabricated device can be mainly attributed to the non-negligible contact resistance between the colusite and Ni0.9Co0.1Sb (resulting in an increase in device resistance) and the consequent reduction in P (Fig. S11). The calculated ηmax for x = 1.0 reached 8.1% at ΔT = 373 K (TL = 300 K, TH = 673 K) (Fig. 5e), highlighting its high potential for TE applications. While this value is comparable to or still lower than other promising materials, e.g., Mg-based compounds and Half-Heusler compounds,11,13,37,38,40 Cu26Ti2Sb4Ge2S32−x could be a strong candidate for practical application if the interfacial material is optimized, given that its primary constituent elements (Cu and S) are low-toxicity, abundant elements.
| This journal is © The Royal Society of Chemistry 2026 |