DOI:
10.1039/D2QI00698G
(Research Article)
Inorg. Chem. Front., 2022,
9, 3552-3558
Structural dimension modulation in a new oxysulfide system of Ae2Sb2O2S3 (Ae = Ca and Ba)†
Received
1st April 2022
, Accepted 27th May 2022
First published on 31st May 2022
Abstract
Inspired by the abundant structural diversity and potential applications of Sb-based oxysulfides, two new compounds with the same stoichiometry, Ae2Sb2O2S3 (Ae = Ca, Ba), were successfully synthesized via solid state reactions. Both crystallize in the C2/c space group (no. 15) of a monoclinic system, however, possess different structural dimensions induced by different sizes of template alkaline-earth ions. Ca2Sb2O2S3 is composed of one-dimensional anionic [Sb2O2S3]4− chains isolated by Ca2+ ions, while Ba2Sb2O2S3 consists of zero-dimensional anionic [Sb2O2S3]4− clusters separated by Ba2+ ions. Higher structural dimension endows the 1D Ca2Sb2O2S3 with a higher melting point (m.p. = 720 °C) and a narrower band gap (Eg = 2.36 eV), compared with the m.p. = 692 °C and Eg = 2.78 eV of Ba2Sb2O2S3. Both exhibit interesting photoluminescence properties with multiemission characteristics. DFT calculations reveal that the band edges of Ae2Sb2O2S3 are mainly composed of Sb−S bond orbitals, while O 2p orbitals also contribute to the valence band maximum of Ba2Sb2O2S3.
Introduction
Owing to their manifold structural chemistry and novel physical properties, oxychalcogenides containing the 15-group metal Sb or Bi have shown great potential in aspects of superconductivity,1–3 optoelectronics,4–8 thermoelectricity9–12 and nonlinear optics (NLO).13 Because of the existence of stereochemically active lone-pair (SALP) hybridized s/p states, Sb3+ ions possess various coordination environments with low symmetry, like a 3-fold triangle pyramid, a 5-fold square pyramid and a 6-fold distorted octahedron. Besides, the moderate bonding affinity of antimony towards S and O makes it possible to form mixed-anion coordination polyhedra, such as [SbOS2]3−,14–16 [SbO3S]5−,17 and [SbOQ4]7− (Q = S, Se) units,4,9,13,18 further increasing the structural diversity of Sb-based oxychalcogenides. Heteroanionic coordination geometries with low symmetry can induce high local polarization, which can not only facilitate carrier separation, benefiting optoelectronic and photocatalysis applications,5,19 but also achieve large second harmonic generation (SHG) responses for nonlinear optics.13,20–23 However, the number of Sb-based oxysulfides containing Sb/O/S mixed-anion units is still limited, as is the investigation towards their potential applications.
Highly ionic cations like alkali metal (A) and alkaline-earth metal (Ae) ions are widely used as structural templates for the construction of new chalcogenides. The spatial arrangement of building units and the structural dimensions of crystals are directly determined by the radii and coordination environments of template ions. Oxychalcogenides containing alkali ions as templates have rarely been reported because the hard-Lewis-acid alkali ions tend to bond with the highly electronegative O2− instead of the soft S2−, inducing phase separation. In contrast, relatively softer alkaline-earth metal ions are more suitable as templates for the construction of oxychalcogenides, and have been used in many material systems like AeZnOS (Ae = Ca, Ba),24,25 Sr2Cu2MO2S2 (M = Mn, Co, Zn)26,27 and AeGeOQ2 (Ae = Sr, Ba; Q = S, Se).20,22 For Sb-based oxychalcogenides, early exploration of the Ae–O–Sb-Q system started from the structural determination of CaSb10O10S6, a mineral named sarabauite in 1978.14–16 Their subsequent synthesis attempts also implied the existence of other new phases in this system. However, over the past four decades, only one oxyselenide, Sr2Sb2O2Se3, in this system was reported in 2017.18 By adding another metal element into the quaternary Ae–O–Sb–Q system, some oxychalcogenides with novel crystal structures have been discovered such as SrOCuSbQ2 (Q = S, Se),4,9 Sr6Cd2Sb6O7S10
13 and Sr3.5Pb2.5Sb6O5S10.7 These materials are promising candidates for photoelectric, thermoelectric or NLO frequency conversion, which urged us to explore new members in the Ae–O–Sb–Q system.
Herein, we used the alkaline-earth metal ions Ca2+ and Ba2+ as templates and successfully synthesized two novel oxysulfides Ae2Sb2O2S3 (A = Ca, Ba). Both compounds crystallize in the C2/c space group, while they possess different coordination chemistries and structural dimensions. Ca2Sb2O2S3 is composed of one-dimensional (1-D) anionic [Sb2O2S3]4− chains isolated by Ca2+ ions, while Ba2Sb2O2S3 consists of zero-dimensional (0-D) anionic [Sb2O2S3]4− clusters separated by Ba2+ ions. Due to their distinct structural dimensions, the two compounds showed different optical absorption properties and thermal stabilities. First-principles calculations were also performed to gain insight into the electronic structures of the two compounds.
Experimental section
Synthesis
Ca2Sb2O2S3.
All operations are performed in an Ar-protected glove box. 0.0721 g of CaS (1 mmol), 0.0972 g of Sb2O3 (1/3 mmol) and 0.0566 g of Sb2S3 (1/6 mmol) were weighted and ground. The mixture was loaded in a carbon-coated silica tube, and it was subsequently flame-sealed under vacuum below 10–3 mbar. The tube was heated to 650 °C in 10 h, held for 48 h, and then cooled to 350 °C in 72 h. Yellow rod-like single crystals were obtained after the above process.
Ba2Sb2O2S3.
0.1694 g of BaS (1 mmol), 0.0972 g of Sb2O3 (1/3 mmol) and 0.0566 g of Sb2S3 (1/6 mmol) were weighted and loaded in a carbon-coated silica tube. The tube was flame-sealed under vacuum, then heated to 600 °C in 10 h, held for 48 h, and then cooled to 300 °C in 72 h. Yellow bulk single crystals were obtained after the above process.
Single crystal X-ray crystallography
Single crystals of Ca2Sb2O2S3 and Ba2Sb2O2S3 were picked out, and data collection was performed under 180 K on a Rigaku XtaALB PRO 007HF single crystal X-ray diffractometer equipped with mirror-monochromated Mo-Kα radiation and an Oxford Cryo stream (80–500 K). The crystal structures were solved by direct methods and refined by full-matrix least-square on F2 using the SHELXTL program package.28 The crystallographic data and structure refinement details of A2Sb2O2S3 (A = Ca and Ba) are summarized in Table S1.† Selected bond lengths and angles are shown in Table S2.† Atomic coordinates, equivalent isotropic displacement parameters and anisotropic displacement parameters are shown in Tables S3–S6.†
Powder X-ray diffraction and scanning electron microscopy
The polycrystalline samples were fully ground for the collection of powder X-ray diffraction (PXRD) data. PXRD analysis was performed at 298 K in the range of 2θ from 5° to 80° at a scan rate of 1.2° min−1 on a Bruker D2 phaser diffractometer equipped with a monochromatized source of Cu Kα radiation (λ = 0.15406 nm) at 4 kW (40 kV, 100 mA). Images of the single crystals and semi-quantitative energy dispersive X-ray (EDX) spectra were obtained using a Phenom Pro scanning electron microscope (SEM) equipped with a Princeton Gamma Tech (PGT) energy-dispersive X-ray analyzer. Single crystals were placed on the surface of a double-sided carbon/aluminium tape attached on an aluminium SEM substrate. EDX data were collected at an accelerating voltage of 15 keV with a 60 s accumulation time.
UV-Vis diffuse reflectance spectroscopy
Optical diffuse-reflectance measurements were performed at room temperature using a UV-4100 spectrophotometer operating from 1000 nm to 250 nm. BaSO4 was used as a 100% reflectance standard. The powder samples were spread on a compacted base of BaSO4 powder. The generated reflectance versus wavelength data were used to calculate the band gaps of the compounds using the Kubelka–Munk equation.29
Photoluminescence (PL) spectroscopy
The PL emission properties of the two compounds were investigated using a steady-state/transient fluorescence spectrometer (Edinburgh FLS980). Emission spectra were collected at room temperature under excitation wavelengths of 270, 280, 300, 320 and 340 nm, respectively.
Thermal analysis
The thermostability of Ae2Sb2O2S3 (Ae = Ca and Ba) was studied by differential scanning calorimetry (DSC) and thermogravimetry analysis (TGA) using a Thermal Analysis SDT2960 thermal analyzer under N2 flow. A silica crucible containing 5.0 mg of the powder sample was placed on the sample side of the detector, with another empty crucible on the reference side. The sample was heated to 850 °C (Ca2Sb2O2S3) or 900 °C (Ba2Sb2O2S3) and cooled down at a rate of ±15 °C min−1 for two cycles.
Electronic structure calculations
First-principles calculations were performed using the projected augmented wave method (PAW)30 within density functional theory (DFT) as implemented in the Vienna Ab Initio Simulation Package (VASP).31–33 The exchange correlation functional was treated within the spin-polarized generalized gradient approximation (GGA) and parameterized by the Perdew–Burke–Ernzerhof (PBE) version.34 The cutoff energy of the plane wave basis was set to 560 eV. The Monkhorst–Pack k-point grids of 4 × 12 × 4 for Ca2Sb2O2S3 and 5 × 10 × 10 for Ba2Sb2O2S3 were used for Brillouin zone (BZ) sampling. The crystal structure and lattice parameters were fixed as the values observed in experiments during structural optimization, while the positions of atoms were relaxed until the atomic forces on each atom were less than 0.01 eV Å−1. The crystal orbital Hamiltonian population (COHP) curves were extracted using the Lobster program.35
Results and discussion
Synthesis
The two compounds were obtained by the solid-state reactions of CaS or BaS, Sb2O3 and Sb2S3. The phase purities of the as-synthesized products were validated by comparing the experimental PXRD patterns with the simulated ones, as shown in Fig. 1a and c. The yellow Ca2Sb2O2S3 crystals show preferred orientation growth with a rod-like form, while the pale yellow Ba2Sb2O2S3 crystals possess bulk shapes, as shown in the SEM images in Fig. 1b and d. All the four elements distributed uniformly in the crystals revealed by EDX mapping analysis. The semiquantitative EDX spectra give molar ratios of Ca
:
Sb
:
S = 1
:
1.4
:
1.6 and Ba
:
Sb
:
S = 1
:
1.2
:
1.3 in Ca2Sb2O2S3 and Ba2Sb2O2S3 crystals, respectively (Fig. S1†).
 |
| Fig. 1 Simulated (red) and experimental (black) PXRD patterns of (a) Ca2Sb2O2S3 and (c) Ba2Sb2O2S3. The SEM images and EDX mapping analysis of (b) Ca2Sb2O2S3 and (d) Ba2Sb2O2S3. | |
Crystal structures
The crystal structures of Ae2Sb2O2S3 (Ae = Ca, Ba) were determined by single crystal X-ray diffraction and structural refinements, and the crystallographic data and refinement details are shown in Table S1.† Both compounds crystallize in the C2/c space group (no. 15) of the monoclinic system, with one independent alkaline-earth Ca or Ba site (8f Wyckoff position), one independent Sb site (8f), one independent O site (8f) and two independent S sites (4e, 8f). They share the same stoichiometry, nevertheless, possess distinct crystal structures with different structural dimensions, as shown in Fig. 2.
 |
| Fig. 2 The basic structural units and crystal structures of (a) Ca2Sb2O2S3 and (b) Ba2Sb2O2S3 viewed down the b axis. | |
Ca2Sb2O2S3 is composed of 1-D anionic [Sb2O2S3]4− chains isolated with each other by Ca2+ ions, which shares a similar structure with recently reported oxychalcogenides A2Bi2O2Se3 (A = Sr and Ba) and Sr2Sb2O2Se3.18 The edge-sharing [SbOS4]7− square pyramids extend in the b direction to form the [Sb2O2S3]4− chains that are truncated to two pyramidal units in the transverse direction. Within a single chain, two neighboring rows of [SbOS4]7− point to the opposite orientations. The O atoms are surrounded by three Ca atoms and one Sb atom to form edge-sharing tetrahedral chains (Fig. S2a†), which is similar to the PbO-type coordination environments of O atoms in oxychalcogenides such as CeOSbS2, Sr6Cd2Sb6O7S10 and SrOCuSbQ2 (Q = S, Se).4,9,13,36 In comparison, Ba2Sb2O2S3 possesses a novel structure type which consists of 0-D anionic [Sb2O2S3]4− clusters separated by Ba2+ ions. The neighboring [SbOS2]3− trigonal pyramids connect each other by sharing the S atom to form the bi-trigonal pyramid [Sb2O2S3]4−. This kind of oxysulfide building unit in Ba2Sb2O2S3 is reported for the first time as far as we know. The tetrahedra of the centered O atom and the surrounding three Ba atoms and one Sb atom are corner-sharing or edge-sharing with each other to form an ionic layer in the bc plane (Fig. S2b†). According to our computational analysis discussed below, Ba2+ ions interact with the oxygen atoms through ionic bonding, and in doing so interrupt the connection of the [Sb2O2S3]3− units via the bridging of the oxygen atoms to form 1D chains.
Alkaline-earth metal ions are structure-directing agents to modulate the connection modes and spatial arrangements of covalent building units, as well as the dimension of the crystal structures. In the oxychalcogenide Ae2Sb2O2Q3 (Ae = Ca, Sr, Ba; Q = S, Se) system, the coordination environments of Sb atoms and the connected modes of Sb/O/Q groups are determined by Ae ions. Furthermore, the structural dimensions of compounds in this system can be adjusted by changing the Ae ions (Fig. 3). In the three compounds Ca2Sb2O2S3, Sr2Sb2O2Se3 and Ba2Sb2O2S3, the coordination numbers of Ae ions become higher along with the increasing ionic radii, from 6-fold of Ca2+ (99 pm), to 8-fold of Sr2+(113 pm) and 9-fold of Ba2+ (135 pm). Larger ionic cations have stronger capability to ‘cut’ the covalent framework into smaller pieces, resulting in lower structural dimension.
 |
| Fig. 3 Structure comparison, and coordination environments of alkaline-earth ions and Sb atoms of (a) Ca2Sb2O2S3, (b) Sr2Sb2O2Se3 and (c) Ba2Sb2O2S3. | |
The Sb atoms in Ca2Sb2O2S3 and Sr2Sb2O2Se3 possess similar coordination modes. However, the [SbOS4]7− square pyramids in Ca2Sb2O2S3 are highly distorted with two short Sb–S bonds (2.479 Å and 2.608 Å) and two longer ones (3.007 Å and 3.057 Å), compared with the more uniform Sb–Se bonds (2.788–3.041 Å) in Sr2Sb2O2Se3. This difference indicates that the 1-D structure containing Ca2+ template is loosely packed. The Sb–S bond lengths in Ca2Sb2O2S3 correspond to those in other oxysulfides containing [SbOS4]7− units like SrOCuSbS2 (2.67–2.95 Å) and Sr6Cd2Sb6O7S10 (2.435–2.951 Å). With the increase of template size, the Sb/O/S covalent framework is divided into isolated [Sb2O2S3]4− clusters in the 0-D Ba2Sb2O2S3. The Sb–S bond lengths of the [SbOS2]3− trigonal pyramids are 2.415 Å and 2.608 Å, which are comparable to those in compounds such as CaSb10O10S6 (2.468–2.485 Å) containing [SbOS2]3−,15 KCu2SbS3 (2.448 Å)37 and Ba2SbS3I38 (2.389–2.408 Å) containing [SbS3]3− units. Therefore, this work provides a simple strategy to adjust the spatial arrangement of building units, structural dimension, and further the dimension-related transport properties and electronic structures in the Sb-based oxychalcogenide system.
Thermal behavior
The thermal behavior of the two compounds was investigated using DSC-TGA analysis under N2 flow, and the results are shown in Fig. 4. For Ca2Sb2O2S3, a slight weight loss of 5% occurred in the range of 580–720 °C in the TGA curve. Subsequently, it showed a significant weight loss on heating above 720 °C, accompanied by two endothermic peaks at 721 °C and 777 °C in the first cycle of the DSC curve. In the second cycle, both endothermic peaks disappeared. No exothermic peak was observed in the cooling process. The results indicate that Ca2Sb2O2S3 melts incongruently around 720 °C. The TGA curve of Ba2Sb2O2S3 showed a slight weight loss of 2% around 550 °C. A single endothermic peak at 692 °C appeared during the heating process of the first cycle of the DSC curve with an obvious weight loss. No other peak was observed in the DSC curve, indicating an incongruent-melting behavior of Ba2Sb2O2S3. The melting point of Ba2Sb2O2S3 (692 °C) is lower than that of Ca2Sb2O2S3 (720 °C), which reflects weaker interactions between the Sb/O/S units in the 0-D structure than those in the 1-D structure.
 |
| Fig. 4 The TGA curves (blue) and DSC curves (first cycle: black and second cycle: red) of (a) Ca2Sb2O2S3 and (b) Ba2Sb2O2S3. | |
Optical properties
The UV-Vis optical absorption properties of Ca2Sb2O2S3 and Ba2Sb2O2S3 were investigated by UV-Vis diffuse reflectance spectroscopy, and the results are shown in Fig. 5a. Both spectra exhibit obvious adsorption edges, indicating the semiconducting nature of the two compounds. The band gap values were determined by the extrapolation method using the Kubelak–Munk equation (Fig. 5b).
 |
| Fig. 5 (a) UV-Vis absorption spectra of Ae2Sb2O2S3, and (b) a plot of (αhν)2vs. energy obtained using the Kubelka–Munk equation. PL emission spectra of (c) Ba2Sb2O2S3 and (d) Ca2Sb2O2S3 excited at 270, 280, 300, 320, 340 nm. | |
The (αhν)2versus hν curves (where α is the absorption coefficient, and h is Planck's constant, ν is the frequency) reveal the band gaps of 2.36 eV and 2.78 eV for Ca2Sb2O2S3 and Ba2Sb2O2S3, respectively. The 0.42 eV band gap difference reflects the different electronic structures of the two compositionally close compounds. The narrower band gap of Ca2Sb2O2S3 mainly origins from the extending bandwidth brought by the stronger orbital interactions in the 1-D structure. The different coordination environments around the Sb atoms in the two compounds also lead to different contributions of Sb orbitals near the band edges, which would also affect the band gap values. The band gap of Ba2Sb2O2S3 is slightly larger than that of Ba2SbS3I (2.64 eV) containing isolated [SbS3]6− units,38 which could be induced by the incorporation of highly electronegative O2−.
The PL emission properties of the two compounds were examined at various excitation wavelengths. As shown Fig. 5c and d, both compounds exhibit multi-emission characters. Two emission bands centered at 406 and 550 nm appeared in the spectrum of Ba2Sb2O2S3 excited at 270 nm, while the former showed higher intensity. Once the excitation wavelength increased to 280 nm, the emission intensity at 406 nm sharply decreased while the band at 550 nm showed relatively slight weakness. Both emissions became weaker along with the increasing excitation wavelengths, and a new emission band centered around 395 nm appeared under 320 nm and 340 nm excitation. The PL emission properties of Ca2Sb2O2S3 are similar to those of Ba2Sb2O2S3. It also revealed two emission bands centered at 406 nm and 550 nm once excited at 270 nm, while the latter showed higher intensity than that of Ba2Sb2O2S3. The above-band-gap PL emissions of the two compounds should originate from the processes involving defect levels below the valence band maximum or above the conduction band minimum. Such above-band-gap PL emissions were also observed in other compounds, such as CH3NH3PbI3,39 CsPbCl3, CsPbBr3,40 TlPbI3,41 GaAs1−xNx
42 and GaAs1−xPx.43 The similar PL emission behaviors indicate similar types of defects in the two closely-related compounds.
Electronic structures
In order to reveal the origin of different optical absorption properties of the two compounds, DFT calculations were performed. The band structures shown in Fig. 6a and b reveal that Ca2Sb2O2S3 possesses a direct band gap (Γ to Γ) of 1.42 eV, while the minimum band gap of Ba2Sb2O2S3 is indirect (C to Γ) with an energy of 2.05 eV. As the energy at Γ near the valence band edge is just 0.05 eV lower than that at C, Ba2Sb2O2S3 can be considered as a quasi-direct-band gap semiconductor. The underestimation of calculated band gaps compared with the experimental values originates from the limitations of DFT calculations we performed. It is worth noting that the valence band dispersion of Ca2Sb2O2S3 near the Fermi level is more obvious than that of Ba2Sb2O2S3, reflecting the stronger bond interactions in the 1-D [Sb2O2S3]4− chains of the former. It also indicates a better hole-transport capability in Ca2Sb2O2S3.
 |
| Fig. 6 Band structures of (a) Ca2Sb2O2S3 and (b) Ba2Sb2O2S3, and total DOS and partial DOS of Sb, O and S in (c) Ca2Sb2O2S3 and (d) Ba2Sb2O2S3. | |
Total and partial density of states (DOS) reveals that the valence band maximum (VBM) and the conduction band minimum (CBM) of Ca2Sb2O2S3 are mainly contributed by the S and Sb orbitals (Fig. 6c). The VBM is composed of S 3p states, Sb 5s and 5p states, while the CBM is composed of Sb 5p and S 3p states. The localized Sb 5s states near the VBM indicate the lone-pair character of Sb 5s2 electrons. The main part of O 2p states is located at −1 eV below the Fermi level, making little contribution to the band edge. For Ba2Sb2O2S3, the VBM is mainly composed of S 3p, Sb 5s and 5p states, and some contribution of O 2p states (Fig. 6d). The CBM consists of Sb 5p and S 3p states, which is similar to that of Ca2Sb2O2S3. The lone-pair character of Sb 5s2 in Ba2Sb2O2S3 is not that obvious, which is reflected by the almost equal contribution of Sb 5s and 5p states near the VBM. In addition, the main distribution of Sb 5p states appears above 2.5 eV in the CB of Ba2Sb2O2S3 along with the increase O 2p density in this region, while the density of Sb 5p states in Ca2Sb2O2S3 appears near the CBM. These differences could be induced by O orbital participation near the band edges in Ba2Sb2O2S3. Therefore, the wider band gap of Ba2Sb2O2S3 origins not only from the narrow bandwidth caused by suppressed band dispersion in the 0-D structure, but also from the participation of O states in the band edges.
As the soft Lewis acid Ba2+ in Ba2Sb2O2S3 can presumably facilitate the formation of isolated [SbOS2]3− trigonal pyramids by interacting with the soft base lone pair on Sb3+, we first look at the local environment of Sb. As shown in Fig. S3,† Sb has four closest Ba neighbors at distances less than 4 Å (3.777–3.924 Å), while none of them is in the direction of the Sb lone pair. The reason is that three Ba atoms are close to the O atom at distances around 2.7 Å. The remaining one is close to the S atom of another [Sb2O2S3]4− cluster. In this fashion, the ionic attraction is maximized at the expense of the Lewis acid–base interaction. The COHP curves for the Sb–O, Sb–S (Fig. S4a†), Ba–Sb, Ba–O and Ba–S bonds (Fig. S4b†) were also calculated to analyze the strength and nature of the bonds. The COHP bonding values for the Sb–O and Sb–S bonds are about 10 times larger than those of the Ba–Sb, Ba–O and Ba–S contacts, indicating the dominant ionic interaction among the latter contacts. Furthermore, the Ba–O interaction is much stronger than those of the Ba–S and Ba–Sb contacts. All these confirm that the interaction of Ba2+ with the host lattice is ionic with negligible contribution from the Ba–Sb interaction.
Conclusions
In summary, we have successfully synthesized two novel Sb-based oxysulfides Ae2Sb2O2S3 (Ae = Ca, Ba) via solid state reactions. Both crystallize in the C2/c space group (no. 15) with different structure types. Ca2Sb2O2S3 is composed of 1-D anionic [Sb2O2S3]4− chains isolated by Ca2+ ions. The [Sb2O2S3]4− chain is constructed by edge-sharing [SbOS4]7− square pyramids. Ba2Sb2O2S3 consists of 0-D anionic [Sb2O2S3]4− clusters separated by Ba2+ ions. The higher structural dimensions of Ca2Sb2O2S3 induce a higher m.p. (720 °C) and a narrower band gap (Eg = 2.36 eV), compared with those of Ba2Sb2O2S3 (m.p. = 692 °C and Eg = 2.78 eV). Both compounds show interesting multi-emission characters. The calculated electronic structures reveal that the band edges of both compounds mainly consist of Sb–S bond orbitals, while O 2p orbitals also make some contribution to the VBM of Ba2Sb2O2S3. We suppose that this work provides an example of the template effect on group arrangement and structural dimension, and can serve as a guide for the structural design of new functional Sb-based oxychalcogenides.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (Grant 21871008, 22005006, and 22001263) and the China Postdoctoral Science Foundation (Grant 2019M660298 and 2020T130009). We would thank Prof. Chong Zheng for his support in the theoretical calculations.
Notes and references
- R. Matsumoto, Y. Goto, S. Yamamoto, K. Sudo, H. Usui, A. Miura, C. Moriyoshi, Y. Kuroiwa, S. Adachi and T. Irifune, Pressure-induced superconductivity in the layered pnictogen diselenide NdO0.8F0.2Sb1−xBixSe2 (x = 0.3 and 0.7), Phys. Rev. B, 2019, 100, 094528 CrossRef CAS.
- Y. Mizuguchi, S. Demura, K. Deguchi, Y. Takano, H. Fujihisa, Y. Gotoh, H. Izawa and O. Miura, Superconductivity in Novel BiS2-Based Layered Superconductor LaO1-xFxBiS2, J. Phys. Soc. Jpn., 2012, 81, 114725 CrossRef.
- Y. Mizuguchi, H. Fujihisa, Y. Gotoh, K. Suzuki, H. Usui, K. Kuroki, S. Demura, Y. Takano, H. Izawa and O. Miura, BiS 2-based layered superconductor Bi4O4S3, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 220510 CrossRef.
- K. Bu, M. Luo, R. Wang, X. Zhang, J. He, D. Wang, W. Zhao and F. Huang, Enhanced Photoelectric SrOCuSbS2 of a [SrO]-Intercalated CuSbS2 Structure, Inorg. Chem., 2018, 58, 69–72 CrossRef PubMed.
- S. Meng, X. Zhang, G. Zhang, Y. Wang, H. Zhang and F. Huang, Synthesis, crystal structure, and photoelectric properties of a new layered bismuth oxysulfide, Inorg. Chem., 2015, 54, 5768–5773 CrossRef CAS PubMed.
- J. Wu, Y. Liu, Z. Tan, C. Tan, J. Yin, T. Li, T. Tu and H. Peng, Chemical Patterning of High–Mobility Semiconducting 2D Bi2O2Se Crystals for Integrated Optoelectronic Devices, Adv. Mater., 2017, 29, 1704060 CrossRef PubMed.
- R. Wang, K. Bu, X. Zhang, Y. Gu, Y. Xiao, Z. Zhan and F. Huang, A novel two-dimensional oxysulfide Sr3.5Pb2.5Sb6O5S10: synthesis, crystal structure, and photoelectric properties, J. Mater. Chem. C, 2020, 8, 11018–11021 RSC.
- M. Haj Lakhdar, T. Larbi, B. Ouni and M. Amlouk, AC conductivity, dielectric relaxation and modulus behavior of Sb2S2O new kermesite alloy for optoelectronic applications, Mater. Sci. Semicond. Process., 2015, 40, 596–601 CrossRef CAS.
- K. Bu, J. Huang, M. Luo, M. Guan, C. Zheng, J. Pan, X. Zhang, S. Wang, W. Zhao and X. Shi, Observation of High Seebeck Coefficient and Low Thermal Conductivity in [SrO]-Intercalated CuSbSe2 Compound, Chem. Mater., 2018, 30, 5539–5543 CrossRef CAS.
- Y. Goto, A. Miura, R. Sakagami, Y. Kamihara, C. Moriyoshi, Y. Kuroiwa and Y. Mizuguchi, Synthesis, crystal structure, and thermoelectric properties of layered antimony selenides ReOSbSe2 (RE = La, Ce), J. Phys. Soc. Jpn., 2018, 87, 074703 CrossRef.
- A. Nishida, O. Miura, C.-H. Lee and Y. Mizuguchi, High thermoelectric performance and low thermal conductivity of densified LaOBiSSe, Appl. Phys. Express, 2015, 8, 111801 CrossRef.
- S. Tippireddy, P. K. D S, S. Das and R. C. Mallik, Oxychalcogenides as Thermoelectric Materials: An Overview, ACS Appl. Energy Mater., 2021, 4, 2022–2040 CrossRef CAS.
- R. Wang, F. Liang, F. Wang, Y. Guo, X. Zhang, Y. Xiao, K. Bu, Z. Lin, J. Yao and T. Zhai, Sr6Cd2Sb6O7S10: Strong SHG Response Activated by Highly Polarizable Sb/O/S Groups, Angew. Chem., Int. Ed., 2019, 58, 8078–8081 CrossRef CAS PubMed.
- K. Nagashima, M. Ogino and I. Nakai, Synthesis of Oxide-Chalcogenides. I. Hydrothermal Synthesis of Sarabauite CaSb10O10S6 and Related Compounds, Bull. Chem. Soc. Jpn., 1978, 51, 1761–1763 CrossRef CAS.
- I. Nakai, H. Adachi, S. Matsubara, A. Kato, K. Masutomi, T. Fujiwara and K. Nagashima, Sarabauite, a new oxide sulfide mineral from the Sarabau mine, Sarawak, Malaysia, Am. Mineral., 1978, 63, 715–719 CAS.
- I. Nakai, K. Nagashima, K. Koto and N. Morimoto, Crystal chemistry of oxide–chalcogenide. I. The crystal structure of sarabauite CaSb10O10S6, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1978, 34, 3569–3572 CrossRef.
- J. Hybler and S. Ďurovič, Kermesite, Sb2S2O: crystal structure revision and order–disorder interpretation, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2013, 69, 570–583 CrossRef CAS PubMed.
- J. R. Panella, J. Chamorro and T. M. McQueen, Synthesis and Structure of Three New Oxychalcogenides: A2O2Bi2Se3 (A = Sr, Ba) and Sr2O2Sb2Se3, Chem. Mater., 2016, 28, 890–895 CrossRef CAS.
- R. Kuriki, T. Ichibha, K. Hongo, D. Lu, R. Maezono, H. Kageyama, O. Ishitani, K. Oka and K. Maeda, A stable, narrow-gap oxyfluoride photocatalyst for visible-light hydrogen evolution and carbon dioxide reduction, J. Am. Chem. Soc., 2018, 140, 6648–6655 CrossRef CAS PubMed.
- B.-W. Liu, X.-M. Jiang, G.-E. Wang, H.-Y. Zeng, M.-J. Zhang, S.-F. Li, W.-H. Guo and G.-C. Guo, Oxychalcogenide BaGeOSe2: highly distorted mixed-anion building units leading to a large second-harmonic generation response, Chem. Mater., 2015, 27, 8189–8192 CrossRef CAS.
- Y. Tsujimoto, C. A. Juillerat, W. Zhang, K. Fujii, M. Yashima, P. S. Halasyamani and H.-C. zur Loye, Function of Tetrahedral ZnS3O Building Blocks in the Formation of SrZn2S2O: A Phase Matchable Polar Oxysulfide with a Large Second Harmonic Generation Response, Chem. Mater., 2018, 30, 6486–6493 CrossRef CAS.
- X. Zhang, Y. Xiao, R. Wang, P. Fu, C. Zheng and F. Huang, Synthesis, crystal structures and optical properties of noncentrosymmetric oxysulfides AeGeS2O (Ae = Sr, Ba), Dalton Trans., 2019, 48, 14662–14668 RSC.
- R. Wang, Y. Guo, X. Zhang, Y. Xiao, J. Yao and F. Huang, Sr5Ga8O3S14: A Nonlinear Optical Oxysulfide with Melilite-Derived Structure and Wide Band Gap, Inorg. Chem., 2020, 59, 9944–9950 CrossRef CAS PubMed.
- S. Broadley, Z. A. Gál, F. Cora, C. F. Smura and S. J. Clarke, Vertex-linked ZnO2S2 tetrahedra in the oxysulfide BaZnOS: a new coordination environment for zinc in a condensed solid, Inorg. Chem., 2005, 44, 9092–9096 CrossRef CAS PubMed.
- T. Sambrook, C. F. Smura, S. J. Clarke, K. M. Ok and P. S. Halasyamani, Structure and physical properties of the polar oxysulfide CaZnOS, Inorg. Chem., 2007, 46, 2571–2574 CrossRef CAS PubMed.
- W. Zhu and P. Hor, Unusual layered transition-metal oxysulfides: Sr2Cu2MO2S2 (M = Mn, Zn), J. Solid State Chem., 1997, 130, 319–321 CrossRef CAS.
- W. Zhu, P. Hor, A. Jacobson, G. Crisci, T. Albright, S.-H. Wang and T. Vogt, A2Cu2CoO2S2 (A = Sr, Ba), a novel example of a square-planar CoO2 layer, J. Am. Chem. Soc., 1997, 119, 12398–12399 CrossRef CAS.
-
G. Sheldrick, SHELX-97, Program for the Solution and Refinement of Crystal Structures, University of Göttingen, Göttingen, Germany, 1997 Search PubMed.
- G. Kortüm, W. Braun and G. Herzog, Principles and techniques of diffuse-reflectance spectroscopy, Angew. Chem., Int. Ed. Engl., 1963, 2, 333–341 CrossRef.
- P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef PubMed.
- G. Kresse and J. Furthmüller, Efficiency of ab initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS PubMed.
- G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
- S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, LOBSTER: A tool to extract chemical bonding from plane-wave based DFT, J. Comput. Chem., 2016, 37, 1030–1035 CrossRef CAS PubMed.
- M. Nagao, M. Tanaka, R. Matsumoto, H. Tanaka, S. Watauchi, Y. Takano and I. Tanaka, Growth and Structure of Ce(O, F)SbS2 Single Crystals, Cryst. Growth Des., 2016, 16, 3037–3042 CrossRef CAS.
- R. Wang, X. Zhang, J. He, C. Zheng, J. Lin and F. Huang, Synthesis, crystal structure, electronic structure, and photoelectric response properties of KCu2SbS3, Dalton Trans., 2016, 45, 3473–3479 RSC.
- R. Wang, X. Zhang, J. He, K. Bu, C. Zheng, J. Lin and F. Huang, Synthesis, Structure, and Optical Properties of Antiperovskite-Derived Ba2MQ3X (M = As, Sb; Q = S, Se; X = Cl, Br, I) Chalcohalides, Inorg. Chem., 2018, 57, 1449–1454 CrossRef CAS PubMed.
- Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao and J. Huang, Electron-hole diffusion lengths> 175 μm in solution-grown CH3NH3PbI3 single crystals, Science, 2015, 347, 967–970 CrossRef CAS PubMed.
- M. Sebastian, J. Peters, C. Stoumpos, J. Im, S. Kostina, Z. Liu, M. Kanatzidis, A. Freeman and B. Wessels, Excitonic emissions and above-band-gap luminescence in the single-crystal perovskite semiconductors CsPbBr3 and CsPbCl3, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 92, 235210 CrossRef.
- W. Lin, J. He, K. M. McCall, C. C. Stoumpos, Z. Liu, I. Hadar, S. Das, H. H. Wang, B. X. Wang and D. Y. Chung, Inorganic halide perovskitoid TlPbI3 for ionizing radiation detection, Adv. Funct. Mater., 2021, 31, 2006635 CrossRef CAS.
- P. Tan, X. Luo, Z. Xu, Y. Zhang, A. Mascarenhas, H. Xin, C. Tu and W. Ge, Photoluminescence from the nitrogen-perturbed above-bandgap states in dilute GaAs1−xNx alloys: A microphotoluminescence study, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 205205 CrossRef.
- D. Scifres, N. Holonyak Jr., C. Duke, G. Kleiman, A. Kunz, M. Craford, W. Groves and A. Herzog, Photoluminescence Associated with Multivalley Resonant Impurity States above the Fundamental Band Edge: N Isoelectronic Traps in GaAs1−xPx, Phys. Rev. Lett., 1971, 27, 191 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: EDX spectra, oxygen coordination environments, crystallographic data and structure refinement details, selected bond lengths and angles, atomic coordinates, equivalent isotropic displacement parameters and anisotropic displacement parameters of Ca2Sb2O2S3 and Ba2Sb2O2S3. CCDC 1985461 and 1985463. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2qi00698g |
|
This journal is © the Partner Organisations 2022 |
Click here to see how this site uses Cookies. View our privacy policy here.