Sayani
Barman
a,
Sweta
Yadav
a,
Akshay K.
Ray
a,
Swati
a,
M.
Deepa
a,
Manish K.
Niranjan
b and
Jai
Prakash
*a
aDepartment of Chemistry, Indian Institute of Technology Hyderabad, Kandi, Sangareddy, Telangana 502284, India. E-mail: jaiprakash@chy.iith.ac.in
bDepartment of Physics, Indian Institute of Technology Hyderabad, Kandi, Sangareddy, Telangana 502284, India
First published on 17th December 2024
Depending on their bandgaps, mixed metal layered chalcogenides are potential candidates for thermoelectric and photovoltaic applications. Herein, we reported the exploratory synthesis of Sr–Zr–Cu–Q (Q = S/Se) systems, resulting in the identification of two novel quaternary chalcogenides: Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9. These isoelectronic compounds (Sr3Zr2Cu4Q9) crystallized in two different structural types. The Sr3Zr2Cu4S9 structure (space group: P) adopted the Ba3Zr2Cu4S9 structure type with eighteen unique atomic sites: 3 × Sr, 2 × Zr, 4 × Cu, and 9 × S. In contrast, the Sr3Zr2Cu4Se9 structure (P
) represented a unique structure type with nineteen unique atomic positions including one additional Cu site compared to the Sr3Zr2Cu4S9 structure. The sulfide structure was stoichiometric, whereas the selenide structure was found to be non-stoichiometric with three partially occupied Cu positions. The Sr3Zr2Cu4Q9 structures consisted of
layers with the Sr2+ cations occupying the interstitial spaces. In both structures, the Zr atoms occupied distorted octahedral positions. A striking difference between the two structures resulted from the distinct bonding interactions between the Cu and Q atoms. The optical bandgap of polycrystalline Sr3Zr2Cu4S9 was 1.7(1) eV. Interestingly, resistivity measurements of polycrystalline Sr3Zr2Cu4Se9 revealed metallic/degenerate semiconducting behavior at low temperatures. The photovoltaic performance of semiconducting Sr3Zr2Cu4S9 demonstrated ∼24% increment in power conversion efficiency when incorporated into a TiO2/CdS photoanode due to its narrower bandgap, which increased the light-harvesting ability of the cell. We also explored the theoretical electronic structures, COHP, and Bader charges of the Sr3Zr2Cu4Q9 structures using DFT calculations.
Copper-containing chalcogenides, especially sulfides and selenides, are emerging materials for photovoltaic and thermoelectric (TE) applications due to their non-toxicity and earth abundance.4,17–20 The high mobility of Cu1+ ions in these multinary chalcogenide structures is well known in compounds such as Cu2−xSe,21 Cu7PSe6,22 CuCrSe2,23 and AgCuSe.24 However, some of these structures are unstable at higher temperatures due to the migration of Cu ions within their crystal lattices (e.g., Cu2−xSe).25,26 Unlike copper oxides, the Cu atoms in chalcogenides exist in the monovalent state and are mostly four-coordinated with a tetrahedral coordination geometry. Notable examples of such compounds include BaCu6−xSeTe6,27 BaLnCuS3 (Ln = Pr, Sm, Dy, Ho, Yb),28,29 and Cu4SnS4.30 In these structures, alkali (A) and alkaline-earth (Ak) metals function as electron donors and exhibit higher coordination numbers than the transition metals. The A1+ or Ak2+ ions in these compounds can be replaced with lower or higher valent chemical dopants to control the number of charge carriers. Also, many of these multinary chalcogenides offer low intrinsic thermal conductivity values due to their complex structures and chemical bonding. Therefore, the Cu-containing multinary chalcogenides with narrow bandgaps are actively pursued for their potential thermoelectric applications. Examples of such compounds are ACuZrQ331 and BaSc1−xGdxCuTe3.32
Moreover, the tunable bandgaps of the Cu/Ag-containing chalcogenides also make them potential candidates for solar cells, light-emitting diodes (LEDs), and photocatalysis applications. As discussed earlier, thanks to the high mobility of copper or silver atoms in these semiconducting chalcogenides, additional electronic levels close to the valence band are present in these compounds, making them efficient in electron transfer rather than hole recombination on sunlight absorption.33,34 BaCu2SnS4−xSex,35 CuGaSe2,36 CuSbS2,37 BaCeCuS3,38 Cu2ZnSnQ4 (Q = S, Se),39 and CuInxGa1−x(SxSe1−x)240 are a few examples of such chalcogenide materials with excellent photovoltaic power conversion efficiencies. CuSbS2
37 is one of the chalcogenides actively explored for photovoltaics and thermoelectric applications.
Therefore, many research groups worldwide have been actively synthesizing new Cu/Ag-containing multinary chalcogenides with semiconducting properties to identify new potential materials for solar cells and TE applications. We have produced several novel complex metal chalcogenides in our quest to develop new semiconducting compounds for TE and solar cell applications; these include 12-BaBi2S4,41 Ba8Zr2Se11(Se2),42 and BaCeCuS3.38 In this work, we report the synthesis of two novel compounds, namely, Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9, with layered structures. Sr3Zr2Cu4S9 is the first quaternary phase in the Sr–Zr–Cu–S system, while Sr3Zr2Cu4Se9 is the second member of the Sr–Zr–Cu–Se system for which only the quaternary phase, Sr0.5CuZrSe3,43 is reported. Sr3Zr2Cu4Se9 represents a new structure type. We have studied the structural aspects and physical properties of Sr3Zr2Cu4Q9 compounds. Thermoelectric properties of Sr3Zr2Cu4Se9, a semimetal/degenerate semiconductor, are explored in detail. The Sr3Zr2Cu4S9 sulfide, a semiconductor with an optical bandgap of 1.7(1) eV, is explored for photovoltaic applications. Incorporating the quaternary title sulfide phase in a TiO2/CdS photoanode shows a 24% increment in the power conversion efficiency of the fabricated solar cell. The electronic band structures of the Sr3Zr2Cu4Q9 phases are studied using the DFT method.
We used the SnO2:
F (FTO) coated glass substrates (Pilkington) with a sheet resistance of ∼25 Ω cm−2 to fabricate solar cells for photovoltaic studies. Other chemicals (reagent grade) used for preparing the solar cells and photovoltaic measurements are the following: TiO2 P25 powder, fumed silica and Ni-foam were procured from Evonik, Cabosil, and Gelon, respectively, and cadmium acetate, sodium sulfide, titanium tetrachloride, methanol, toluene, Triton X-100, acetylacetone, multi-walled carbon nanotubes (MWCNTs), polyvinylidene fluoride (PVdF), carbon black (CB), and N-methyl-2-pyrrolidone (NMP) were purchased from Merck and used directly.
Similarly, a vacuum-sealed tube containing the elements strontium (37.28 mg, 0.43 mmol), zirconium (25.88 mg, 0.28 mmol), copper (36.05 mg, 0.57 mmol), and selenium (100.79 mg, 1.28 mmol) was also heated in the furnace to prepare single crystals of Sr3Zr2Cu4Se9.
The tubes were heated in the furnace using the temperature profile described in section S1 of the ESI.† After completion of the reactions, dark reddish and black-colored melted lumps were found for the reactions targeting the sulfide and selenide, respectively. These lumps were further fractured, and a few crystals were chosen for their semiquantitative elemental energy-dispersive X-ray spectroscopic (EDX) analyses using a field-emission scanning electron microscope (FE-SEM) (Model: JSM 7610F) and an X-ray spectrometer (octane elite, EDAX Inc., USA).
Most of the red-colored crystals selected from the sulfide reaction product showed Sr:
Zr
:
Cu
:
S ≈ 3
:
2
:
4
:
9, consistent with the loaded Sr3Zr2Cu4S9 composition (Fig. 1a). A few crystals turned out to be binary SrS (1
:
1). The crystals of Sr3Zr2Cu4Se9 were black in color, and their average composition was also consistent with the loaded ratio of Sr
:
Zr
:
Cu
:
Se ≈ 3
:
2
:
4
:
9 (Fig. 1a inset).
A few of these EDX-analyzed red and black-colored crystals were further used for structural characterization at room temperature (RT) using the single-crystal X-ray diffraction (SCXRD) method.
The elements, strontium (266.03 mg, 3.04 mmol), zirconium (184.65 mg, 2.02 mmol), copper (257.25 mg, 4.05 mmol), and sulfur (292.07 mg, 9.11 mmol), were sealed in a fused silica tube under vacuum and heated inside the furnace to synthesize polycrystalline Sr3Zr2Cu4S9.
Similarly, a vacuum-sealed fused silica tube containing strontium (186.41 mg, 2.13 mmol), zirconium (129.38 mg, 1.41 mmol), copper (180.26 mg, 2.84 mmol), and selenium (503.95 mg, 6.38 mmol) was heated in the furnace to prepare the polycrystalline Sr3Zr2Cu4Se9 using the heating profile discussed in section S1 of the ESI.†
The products of the reactions were homogenized into fine powders and were compressed individually into cylindrical-shaped thin pellets. These pellets were again encapsulated inside evacuated fused silica tubes and annealed at 1023 K for two days. This step is required to make high-quality polycrystalline Sr3Zr2Cu4Q9 (Q = S, Se) samples. The heat-treated pellets were ground again, and the products were characterized using the PXRD studies.
We also attempted to synthesize polycrystalline Sr3Zr2Cu4S9 by heating stoichiometric amounts of metal oxides/carbonates under N2 and CS2 atmosphere at 1073 K for 9 h. The details of this CS2 method are given in section S2 of the ESI.† The ease of making bulk samples and the scalability of the method without the need for expensive fused silica tubes were the motivations behind attempting the CS2 method. However, we failed to synthesize the phase pure material via the CS2 method. Further optimization of the reaction temperature and time is required to find suitable reaction conditions to synthesize the Sr3Zr2Cu4S9 phase in the future.
The PXRD data sets of the powdered samples mentioned above were collected with the help of a PANalytical X'pert pro-X-ray diffractometer (Cu-Kα, λ = 1.5406 Å) at RT. The TOPAS V6 software program was used to refine the unit cell parameters of the polycrystalline samples, Sr3Zr2Cu4Q9, using the Le Bail refinement method (Fig. 1b).44
The crystal structures for the Sr3Zr2Cu4Q9 compounds were determined and refined in the centrosymmetric triclinic space group P with the help of the SHELXTL46 program suite. For the sulfide structure, all the atomic positions were easily assigned to respective elements, which resulted in the formula Sr3Zr2Cu4S9. However, the reliability factor values were poor: R(F) = 8.87%, Rw(Fo2) = 26.72%, and the maximum positive residual electron density was 13.3 e− Å−3. Later, the CELLNOW45 program was used to check the presence of any twinning in the dataset, which identified two twin domains for the crystal. Also, absorption correction was done afterward using the TWINABS.47 The second twin domain was found to be rotated from the first domain with an angle of about 180° along the [1 0 1] direction. After these steps, the refinement parameters were significantly improved with final R(F) = 4.74% and Rw(Fo2) = 12.38%. The Sr3Zr2Cu4S9 structure is stoichiometric and isostructural to the Ba3Zr2Cu4S9
48 (space group: P
) structure. The final structure solution identified a total of 18 independent sites, out of which three sites were identified as Sr, two as Zr, four sites as Cu, and the remaining nine as S sites. The STRUCTURE TIDY49 program was used to standardize the atomic positions of the Sr3Zr2Cu4Q9 structure.
We also found multiple twin domains for the Sr3Zr2Cu4Se9 dataset using the CELLNOW45 program. Similar to the sulfide structure, for this structure as well, the two domains are related with a rotation angle of about 180° with the [1 0 1] direction as the rotation axis. The APEX345 was used to integrate the SCRXD dataset using two twin domains in the triclinic setting. The TWINABS47 was used for the X-ray absorption corrections of the observed reflections. The final structure solution of the Sr3Zr2Cu4Se9 with the P space group was comprised of nineteen crystallographically independent sites. The structure was refined considering 100% occupancy for all the atoms. The reliability parameters R(F) and Rw(Fo2) were 6.68% and 17.56%, respectively. The anisotropic displacement parameters (ADPs) of the Cu atoms were abnormally large, indicating some vacancies at these sites. Subsequently, the fractional occupancy factor (SOF) values of these sites were allowed to be refined freely, resulting in an improvement of the reliability parameters: R(F) = 5.34% and Rw(Fo2) = 11.82%. The freely refined SOF values of the Cu1, Cu3, and Cu4 atomic sites are 0.598, 0.856, and 0.533. The final refined formula, Sr3Zr2Cu3.99(1)Se9, is close to the electrically neutral composition, Sr3Zr2Cu4Se9. Thus, we refer to the chemical formula of the selenide as Sr3Zr2Cu4Se9 in this study.
We finally validated the symmetry of the Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 structures using the ADDSYM program of the PLATON50 software. Further crystallographic details are given in Tables 1, 2 and Tables S1, S2, S3, and S4 in the ESI.†
a λ = 0.71073 Å, T = 298(2) K. b R(F) = ∑||Fo| − |Fc||/∑|Fo| for Fo2 > 2σ(Fo2). c R w(Fo2) = {∑[w(Fo2 − Fc2)2]/∑wFo4}1/2. For Fo2 < 0, w = 1/[σ2(Fo2) + (mP)2 + nP]; where P = (Fo2 + 2Fc2)/3. For the sulfide structure, m = 0.0574 and n = 6.3137, and for the selenide structure, m = 0 and n = 27.2292. | ||
---|---|---|
Sr3Zr2Cu4S9 | Sr3Zr2Cu4Se9 | |
Space group |
P![]() |
|
a (Å) | 6.6907(13) | 6.9779(14) |
b (Å) | 10.312(2) | 10.685(2) |
c (Å) | 11.929(2) | 12.302(3) |
α (°) | 109.49(3) | 109.18(3) |
β (°) | 104.54(3) | 105.38(3) |
γ (°) | 96.73(3) | 96.56(3) |
V (Å3) | 732.6(3) | 814.6(3) |
Z | 2 | |
ρ (g cm−3) | 4.479 | 5.745 |
μ (mm−1) | 19.12 | 36.13 |
R(F)b | 0.047 | 0.053 |
R w(Fo2)c | 0.124 | 0.118 |
Sr3Zr2Cu4S9 | Sr3Zr2Cu4Se9 | |
---|---|---|
Zr–Q (Å) | 2.490(2)–2.694(2) | 2.6342(19)–2.8083(18) |
Tetrahedral Cu–Q (Å) | 2.309(2)–2.372(2) | 2.386(4)–2.682(5) |
Trigonal Cu–Q (Å) | 2.234(2)–2.312(2) | — |
Sr–Q (Å) | 2.924(3)–3.628(2) | 3.0655(19)–3.4277(19) |
The thermal diffusivity (D) values of Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 were measured using a LINSEIS XFA 500 instrument. The details of the measurements are given elsewhere.41 The Dulong–Petit law53 was employed to estimate the heat capacity (Cp) values of the Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 samples. Finally, the total thermal conductivity (κtot = D × d × Cp) values were calculated. The final κtot value has an uncertainty of around ±10%.
The Sr3Zr2Cu4Se9 pellet was further used for the simultaneous determination of electrical resistivity and Seebeck coefficient (S) values at higher temperatures (323 K–573 K) with the help of a ULVAC-RIKO ZEM-3 equipment under a low-pressure He-atmosphere. No hysteresis between the heating and cooling cycles of the measured data was observed.
A counter electrode (CE) was fabricated by coating MWCNTs over a Ni foam substrate. CB, PVdF, and MWCNTs were mixed in a 1:
1
:
8 weight ratio, ground well, and converted into a slurry using NMP. The slurry was coated on the Ni-foam and dried. A polysulfide gel was prepared by adding 5 wt% of fumed silica in an aqueous suspension of (1 M Na2S + 1 M S) under constant stirring until it became a uniform gel.
Finally, we made the solar cells, FTO/TiO2/CdS/S2−, S/MWCNTs/Ni-foam and FTO/TiO2/CdS/Sr3Zr2Cu4S9/S2−, and S/MWCNTs/Ni-foam, by sandwiching the polysulfide gel in between the photoanodes and the CE using binder clips. These solar cells were subjected to I–V measurements with the help of a Newport Oriel 3A solar simulator coupled with a Keithley 2420 source meter. A 450 W Xenon arc lamp was used as the light source with a light intensity of 100 mW cm−2 and Air Mass (AM) 1.5G illumination; the spatial uniformity of irradiance was determined by calibrating with a 2 cm × 2 cm Si reference cell and confirmed using a Newport power meter.
Furthermore, we used two methods to prepare bulk polycrystalline samples of Sr3Zr2Cu4S9: (a) the direct reaction of elements at high temperatures and (b) the CS2 method. The CS2 method failed to produce the targeted Sr3Zr2Cu4S9 phase. Instead, the cubic SrS phase (space group: Fmm)56 was formed as the major phase according to the PXRD analysis (see Fig. S1 of the ESI†). Presumably, the phases containing Zr and Cu are amorphous in the powdered sample obtained by the CS2 reaction. On the other hand, the high-temperature reaction of elements with one intermittent grinding and reheating produced a monophasic Sr3Zr2Cu4S9 sample. The polycrystalline Sr3Zr2Cu4Se9 sample was prepared using heating conditions similar to those employed for the sulfide sample.
The PXRD patterns of the Sr3Zr2Cu4Q9 (Q = S, Se) phases are shown in Fig. 1b. The refined unit cell parameters of Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9, calculated from the PXRD data using the Le Bail refinement method, are a = 6.695(2) Å, b = 10.322(1) Å, c = 11.941(2) Å, α = 109.60(1)°, β = 104.54(1)°, γ = 96.67(1)°, and a = 6.986(2) Å, b = 10.695(2) Å, c = 12.272(1) Å, α = 109.33(1)°, β = 105.35(1)°, γ = 96.57(1)°, respectively. The cell parameters of Sr3Zr2Cu4Q9 (Q = S, Se) obtained from the powder and single crystal XRD studies are in good agreement.
The single-crystal XRD study showed that, at RT, Sr3Zr2Cu4S9 crystallizes in the triclinic Ba3Zr2Cu4S948 structure. As expected, the unit cell volume of the centrosymmetric triclinic Sr3Zr2Cu4S9 structure (space group: P
) with the refined cell parameters a = 6.6907(13) Å, b = 10.312(2) Å, c = 11.929(2) Å, α = 109.49(3)°, β = 104.54(3)°, and γ = 96.73(3)°, is smaller than the Ba3Zr2Cu4S9 cell. These experimental lattice parameters augur well with those obtained from theoretical calculations (see section 3.7). The asymmetric unit of the Sr3Zr2Cu4S9 structure contains nine sulfur and nine metal sites (3 × Sr sites, 2 × Zr sites, and 4 × Cu sites). As seen in Fig. 2a, the
layers of the Sr3Zr2Cu4S9 structure are separated by the Sr2+ cations. The anionic
layer features distorted ZrS6 octahedra, CuS4 tetrahedra (Cu2 and Cu4), and trigonal planar CuS3 units (Cu1 and Cu3), which serve as the main motifs of the structure.
The Sr3Zr2Cu4Se9 structure is isoelectronic to the related sulfide. Interestingly, its structure is different from the Sr3Zr2Cu4S9 structure. The selenide adopts the original structure type with triclinic symmetry (space group: P). The refined cell constants of the Sr3Zr2Cu4Se9 structure are a = 6.9779(14) Å, b = 10.685(2) Å, c = 12.302(3) Å, α = 109.18(3)°, β = 105.38(3)°, and γ = 96.56(3)°. In contrast to the sulfide structure, the Sr3Zr2Cu4Se9 phase is non-stoichiometric with partially vacant Cu sites. Sr3Zr2Cu4Se9 is best described as the pseudo-2D structure, like the sulfide structure, as shown in Fig. 2b.
The selenide structure comprises nineteen independent atoms (3 × Sr, 2 × Zr sites, 5 × Cu sites, and 9 × Se sites), i.e., one additional Cu site in reference to the sulfide structure. The three Cu sites, Cu1, Cu3, and Cu4, have free SOF values of 0.598(8), 0.856(7), and 0.532(8), respectively, in the Sr3Zr2Cu4Se9 structure. The composition of the infinite layers present in the selenide structure is identical to the sulfide structure, i.e., and their negative charges are counterbalanced by the Sr2+ ions. The Zr and Cu atoms in the selenide structure are octahedrally and tetrahedrally coordinated with the Se atoms, respectively. The short Cu1⋯Cu1 and Cu4⋯Cu4 distances (<2.2 Å) in the Sr3Zr2Cu4Se9 structure are the result of their occupancy disorder and imply that these Cu atoms do not coexist simultaneously close to each other in the structure.
Next, we compare the layers of the Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 structures in Fig. 3.
![]() | ||
Fig. 3 (a) The ![]() ![]() ![]() |
The layers of both the structures have edge and vertex-sharing ZrQ6 octahedra (see Fig. 3). The oppositely oriented Cu2S4 and Cu4S4 tetrahedra share edges and vertices with the ZrS6 octahedra to form chains of (see Fig. 3a). Similarly, the connectivity of ZrSe6 octahedra, Cu2Se4, and Cu5Se4 tetrahedra in the Sr3Zr2Cu4Se9 structure creates
chains (see Fig. 3b). Finally, the
layers in both structures are formed by filling of Cu atoms in the structures, which hold together the
chains, as shown in Fig. 3. The main difference between the layers of the title structures is the bonding environments of the additional Cu atoms, Cu1 and Cu3, of the sulfide structure and Cu1, Cu3, and Cu4 atoms in the selenide structure. The Cu atoms (Cu1 and Cu3) in the sulfide structure are only three-coordinated in contrast to the four-coordinated (tetrahedral) Cu atoms of the selenide structure.
The important interatomic distances of the Sr3Zr2Cu4Q9 structures are provided in Table 2. The Zr–S distances of Sr3Zr2Cu4S9 (see Table 2) match well with the corresponding distance of the Ba3Zr2Cu4S9 (2.525(1) Å–2.704(1) Å)48 structure. The Cu–S distances of the CuS3 unit (2.234(2) Å–2.312(2) Å) are shorter than the tetrahedral CuS4 units (2.309(2) Å–2.372(2) Å) in the title sulfide structure. These distances compare well with those found in Ba3Zr2Cu4S9 (2.241(2) Å–2.364(2) Å)48 and the BaCeCuS3 (2.351(1) Å–2.461(2) Å)38 structures.
The Cu–Se (2.386(4) Å–2.682(5) Å) distances of Sr3Zr2Cu4Se9 are consistent with those found in the BaCu2Se2 (2.456(1) Å to 2.618(3) Å)57 and BaNdCuSe3 (2.460(4) Å to 2.567(3) Å)58 structures. The Zr–Se bond distances of ZrSe6 octahedra vary from 2.6342(19) Å to 2.8083(18) Å in the Sr3Zr2Cu4Se9 structure. These distances are in good agreement with those reported for the Ba0.5ZrCuSe3 (2.661(7) Å to 2.707(5) Å)43 and Ba8Zr2Se11(Se2) (2.6337(7) Å to 2.7940(7) Å)42 structures.
Another interesting feature of Sr3Zr2Cu4Q9 (Q = S, Se) is the Cu⋯Cu interactions, which are intermediate to the single Cu–Cu bond (∼2.556 Å)59 and the van der Waals distance of 2.80 Å (ref. 60) between two Cu atoms. These interactions are shown in Fig. 4 for the Sr3Zr2Cu4Q9 (Q = S, Se) structures. The Cu⋯Cu interactions are shorter in the Sr3Zr2Cu4Se9 structure (2.621(5) Å to 2.711(4) Å) than in the Sr3Zr2Cu4S9 structure (2.753(3) Å to 2.845(3) Å). Similar Cu⋯Cu interactions are also known for the Cu5FeS4 (2.587 Å–2.989 Å),61 K2Cu3AlS4 (2.734(3) Å–2.736(3) Å),62 and Cu3SbSe3 (2.665(2) Å)63 structures. The chalcogen atoms of Sr3Zr2Cu4Q9 structures make distorted bicapped trigonal prism-like geometries around the Sr atoms, as shown in Fig. S2 in the ESI.†
![]() | ||
Fig. 4 A fragment of the ![]() |
To the best of our knowledge, only monovalent copper species are known for the chalcogenide structures. Also, the Sr3Zr2Cu4Q9 structures are traditional chalcogenides (not polychalcogenides) with more stable Q2− species. Therefore, these two structures can be charge balanced with the Sr2+, Zr4+, Cu1+, and Q2− species, i.e., the charge partitioned formula of the title phases is (Sr2+)3(Zr4+)2(Cu1+)4(Q2−)9.
The Sr3Zr2Cu4Se9 (P) type is closely related to Sr0.5ZrCuSe3
43 in terms of the coordination environments of the transition metals, i.e., Zr and Cu atoms, which occupy distorted octahedral and tetrahedral sites, respectively. The CuS3 units with a triangular planar geometry make the triclinic Ak3Zr2Cu4S9 (Ak = Ba and Sr) structure type (P
)48 unique to the other two structure types. No short Cu⋯Cu interactions exist in the Sr0.5ZrCuSe3
43 structure. In contrast, the Ak3Zr2Cu4S9 (Ak = Ba and Sr) and Sr3Zr2Cu4Se9 structure types feature a variety of short Cu⋯Cu interactions.
Although the Ak3Zr2Cu4S9 (Ak = Ba, Sr) and Sr3Zr2Cu4Se9 structures are isoelectronic and layered, the intralayer connections are entirely different in these two structure types, as described above (see Fig. 3). While the stoichiometric sulfur compound has two Cu sites as the bridging elements, the non-stoichiometric selenide compound has three Cu atoms to link the chains, as shown in Fig. 3 and 4. Thus, the flexibility of copper atoms to stabilize different polyhedra in the title sulfide structure can be considered as the main contrasting feature between the Ak3Zr2Cu4S9 and Sr3Zr2Cu4Se9 types.
![]() | ||
Fig. 6 (a) Tauc plots of the finely ground polycrystalline Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 samples. (b) The electrical resistivity plot of the Sr3Zr2Cu4Se9 compound below RT. |
A direct bandgap energy value of 1.7(1) eV is calculated for the Sr3Zr2Cu4S9 phase from the Tauc plot, which agrees with the reddish-brown color of the powder. The theoretical bandgap of 1.78 eV for the Sr3Zr2Cu4S9 structure is in good agreement with the experimental value (see section 3.7 for details). On the other hand, Sr3Zr2Cu4Se9, which is black in color, does not show any sharp transition in the Tauc plot, indicating that this selenide has a narrow bandgap energy (Eg < 0.6 eV), or it is a metal/semimetal.
The sulfide sample, Sr3Zr2Cu4S9, is electrically insulating, having resistance values of the order of MΩ at RT in contrast to the Sr3Zr2Cu4Se9, which has electrical resistivity (ρ) values of about 80 mΩ cm (at RT). The ρ-value further drops to ∼20 mΩ cm on cooling the Sr3Zr2Cu4Se9 sample to 3 K, as shown in Fig. 6b. A kink in the resistivity plot was observed at ∼225 K, and this effect was reproducible. The origin of the kink observed in the resistivity plot is not clear. The resistivity values of Sr3Zr2Cu4Se9 are several orders higher than coinage metals like Cu (1.7 × 10−3 mΩ cm at RT) and Ag (1.6 × 10−3 mΩ cm at RT). Thus, Sr3Zr2Cu4Se9 can be described as a poor metal or a degenerate semiconductor. The detailed theoretical electronic structure of Sr3Zr2Cu4Se9 is discussed in section 3.7.
The κtot value for each sample is the sum of the electronic (κele) and lattice thermal conductivities (κlat) of the crystals. Since Sr3Zr2Cu4Se9 is metallic, in contrast to the insulating Sr3Zr2Cu4S9, the free charge carriers in the former compound contribute to the total thermal conductivity. On the contrary, the κele value of the insulating Sr3Zr2Cu4S9 is negligible in comparison to the κlat values. We calculated the κele value of Sr3Zr2Cu4Se9 using the Wiedemann–Franz model65 (refer to the caption of Fig. S3 in the ESI†) to compare the κlat values of the two compounds. The κlat value is also higher for Sr3Zr2Cu4Se9 (3.08 W mK−1 to 2.1 W mK−1 in the range of 323 K to 573 K) than for the isoelectronic sulfide (see Fig. 7a). The higher thermal conductivity values for Sr3Zr2Cu4Q9 (Q = S, Se) make them unsuitable for thermoelectric applications.
Next, we discuss the trends of the Seebeck coefficient and resistivity data obtained for Sr3Zr2Cu4Se9 as a function of temperature. The electrical resistivity plot shown in Fig. 7b is consistent with metal-like behavior with a gradual increase in ρ-values on increasing the sample's temperature. The ρ-value of Sr3Zr2Cu4Se9 elevated from 190 mΩ cm at 323 K to 320 mΩ cm at 573 K. The p-type nature of Sr3Zr2Cu4Se9 is confirmed by the positive sign of the S-values, as shown in Fig. 7c. As expected, the S-value increases at higher temperatures as the sample becomes more electrically resistive. Unfortunately, the low S values (12.51 μV K−1 at 323 K to 24.60 μV K−1 at 573 K) combined with the relatively higher thermal conductivity values of the Sr3Zr2Cu4Se9 sample make it unsuitable for thermoelectric applications.
The maximum thermoelectric figure of merit, zT, is almost zero (16.5 × 10−7), as can be seen from the zT plot in Fig. S3 in the ESI.† Further improvement of the zT value by optimization of the charge carrier concentration seems unlikely due to the poor S-value of Sr3Zr2Cu4Se9.
![]() | ||
Fig. 8 The J–V characteristics of the solar cells based on different photoanodes, S2−, S gel electrolyte, and MWCNTs/Ni-foam as CE under 1 sun (AM 1.5G, 100 m cm−2). |
The J–V plot of the FTO/TiO2/CdS/Sr3Zr2Cu4S9/S2−, S/MWCNTs/Ni-foam shows: VOC = 868 mV, JSC = 13.28 mA cm−2, FF = 60.11%, and the efficiency = 6.93%, as seen in Table 3. From the Tauc plot, shown in Fig. 6a, the bandgap of Sr3Zr2Cu4S9 was estimated to be 1.7 eV, which is much lower than the bandgaps of TiO2 (3.2 eV) and CdS (2.3 eV) used in our solar cell. By the inclusion of Sr3Zr2Cu4S9 in the photoanode, the overall efficiency was increased by ∼24% because Sr3Zr2Cu4S9, a narrow band gap semiconductor, serves as a co-sensitizer. This implies that in the co-sensitized cell, upon illumination, in addition to the CdS, the Sr3Zr2Cu4S9 phase also undergoes an electron–hole separation. Hence, in the co-sensitized cell, additional charge carriers are available for circulation, improving the VOC as well as the JSC, thus increasing the efficiency of the cell. Furthermore, co-sensitization also reduces back electron transfer to the electrolyte during cell operation, which maximizes charge separation and enhances efficiency.
Photoanode | VOC (mV) | JSC (mA cm−2) | FF (%) | PCE (%) |
---|---|---|---|---|
FTO/TiO2/CdS | 723 | 13.00 | 58.83 | 5.59 |
FTO/TiO2/CdS/Sr3Zr2Cu4S9 | 868 | 13.28 | 60.11 | 6.93 |
a (Å) | b (Å) | c (Å) | α | β | γ | E g (eV) | |||
---|---|---|---|---|---|---|---|---|---|
Cal. | GGA | 6.719 | 10.321 | 11.937 | 109.41 | 104.79 | 96.80 | 1.37 | |
Sr3Zr2Cu4S9 | LDA | 6.539 | 9.953 | 11.713 | 109.97 | 104.21 | 96.99 | 1.29 | |
GGA-1/2 | — | — | — | — | — | — | 1.78 | ||
Exp. | 6.662 | 10.232 | 11.939 | 109.69 | 104.84 | 96.39 | 1.7(1) | ||
Cal. | GGA | 7.044 | 10.697 | 12.333 | 109.31 | 105.68 | 96.50 | 1.17 | |
Sr3Zr2Cu4Se9 | LDA | 6.830 | 10.402 | 12.006 | 109.33 | 105.60 | 96.56 | 1.09 | |
GGA-1/2 | — | — | — | — | — | — | 1.56 | ||
Exp | 6.9779 | 10.685 | 12.302 | 109.18 | 105.38 | 96.56 | 0 |
The GGA and GGA-1/2 computed bandgap values of Sr3Zr2Cu4Se9 are 1.17 eV and 1.56 eV, respectively. These computed bandgap energies disagree with the experimental result that suggests Sr3Zr2Cu4Se9 is metallic/degenerate semiconducting in nature. This disagreement may be partially attributed to the slightly different occupancy values of Cu1, Cu3, and Cu4 sites in the unit cell used to model the Sr3Zr2Cu4Se9 structure as compared to the experimental unit cell content obtained from the SCXRD study. In our theoretical model, the Cu1 and Cu4 sites are half occupied (50%), whereas the Cu3 site is fully occupied (100%). However, in the experimental unit cell, the Cu1, Cu3, and Cu4 sites have occupancies of 59.8%, 85.8%, and 53.3%, respectively.
Fig. 9 shows the density of states (DOS) and the band structure plots for Sr3Zr2Cu4S9. The energies in the DOS plot are scaled in such a way that the valence band maximum (VBM) lies at 0 (see Fig. 9a). As can be seen, the top of the valence band (VB) (∼−5.5 eV < E < 0) is primarily comprised of Cu-3d, Zr-4d, and S-3p states. The VB at ∼−12 eV is mainly contributed by the S-3s localized states, with some minor contributions from the Sr-4s states. The conduction band (CB) close to the Fermi energy (∼1.1 eV < E < 2.2 eV) is primarily made up of the antibonding Zr-4d states. The contributions to the CB in the higher energy range (2.9 eV < E < 4.5 eV) are from the Zr-4d and Sr-4d states. The band structure of Sr3Zr2Cu4S9, as shown in Fig. 9b, indicates that the bandgap is indirect. However, the difference between the indirect and direct bandgap is minute (∼0.07 eV). The DOS and band structure plots for Sr3Zr2Cu4Se9 are found to be qualitatively similar to that for Sr3Zr2Cu4S9, as shown in Fig. S4 of the ESI.†
The projected crystal orbital Hamilton populations (COHP)67 are computed for the title structures to estimate the bonding type between the constituent atoms and their strength. Fig. 10 shows the −COHP (E) and integrated-COHP plots for Sr3Zr2Cu4S9. The −ICOHP values suggest that the bond strength between atom pairs varies in the order Zr–S > Cu–S > Cu–Zr > Sr–S. The COHP and ICOHP plots for Sr3Zr2Cu4Se9 are presented in Fig. S5 in ESI† and indicate a similar order for bonding strengths.
![]() | ||
Fig. 10 (a) The crystal orbital Hamilton populations (COHP) and (b) the integrated COHP (ICOHP) values for the bonding atoms in Sr3Zr2Cu4S9. |
We next discuss the computed Bader charges (QB) of the atoms in the Sr3Zr2Cu4Q9 structures. In general, the magnitudes of the Bader charges are smaller than the formal oxidation states of atoms since the latter values are calculated by considering the full charge transfer from metals (cations) to non-metals (anions), i.e., considering the crystalline inorganic solids as perfect ionic compounds. The QB values provide estimates of the relative charge transfer between atom pairs and the degree of ionic/covalent character of the bonds.68 For Sr3Zr2Cu4S9, these charges are +1.54, +2.02, +0.39, and −1.16 for Sr, Zr, Cu, and S ions, respectively. Likewise, in Sr3Zr2Cu4Se9, QB values of +1.49, +1.87, +0.35, and −1.10 were found for the Sr, Zr, Cu, and Se ions, respectively. Indeed, the magnitudes of the computed QB values are smaller than the formal oxidation states +II, +IV, +I, and −II for the Sr, Zr, Cu, and S (Se) ions. The deviation of QB values of the atoms, especially Zr, Cu, and S, in the Sr3Zr2Cu4Q9 structures from their formal oxidation numbers indicate the partial covalent character of the metal–chalcogen bonds due to the overlapping of orbitals, as discussed earlier.
Footnote |
† Electronic supplementary information (ESI) available: Additional crystallographic details of the Sr3Zr2Cu4S9 and Sr3Zr2Cu4Se9 structures. CCDC 2390675 and 2390676. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt02928c |
This journal is © The Royal Society of Chemistry 2025 |