Alexander
Stepanjuga
ab,
Rajyavardhan
Ray
acd,
Manuel
Richter
ac,
Salvatore
Carrocci
a,
Silke
Hampel
a,
Lydia
Galle
e,
Hans-Joachim
Grafe
a and
Martin
Valldor
*af
aLeibniz IFW Dresden, Helmholtzstraße 20, D-01069 Dresden, Germany. E-mail: b.m.valldor@kjemi.uio.no
bHochschule für Technik und Wirtschaft Dresden, University of Applied Science, Friedrich-List-Platz 1, DE-01069 Dresden, Germany
cDresden Center for Computational Materials Science (DCMS), TU Dresden, D-01062, Dresden, Germany
dDepartment of Physics, Birla Institute of Technology Mesra, Ranchi, Jharkhand 835215, India
eInorganic Chemistry Department I, Dresden University of Technology, Bergstraβe 66, D-01062 Dresden, Germany
fDepartment of Chemistry, University of Oslo, P.O. Box 1033 Blindern, NO-0315 Oslo, Norway
First published on 7th February 2022
Pure powders of Na(LiZn)S2 can be obtained through a solid-state reaction, and A(LiZn)S2 (A = K, Rb, Cs) result from metathesis reactions between alkali-metal chlorides and the same constituents used to prepare the Na(LiZn)S2 powder. Hence, the metathesis reaction enables extended sulphide chemistry without the use of either H2S gas or very reactive starting materials. By the metathesis reaction it was possible to obtain relatively pure Cs(LiZn)S2. Trigonal Na(LiZn)S2 and tetragonal A(LiZn)S2 (A = K, Rb, Cs) exhibit significant structural similarities, having (LiZn)S2 layers that are separated by alkali-metals (Na–Cs). Against expectations, Cs(LiZn)S2 is orange red in colour, Rb(LiZn)S2 is strongly yellow, K(LiZn)S2 is pale yellow, and Na(LiZn)S2 is colourless. Ultraviolet-visible spectroscopy data on Cs(LiZn)S2 and Na(LiZn)S2 contain several shoulders apart from apparent band-edges close to 3.3 eV. In the former, it seems as if the optical excitations range all the way into green (∼600 nm), which concurs with the observed red colour. Nuclear magnetic resonance investigations on cores 133Cs, 23Na, and 7Li suggest that these ions are firmly held in the atomic lattice, as judged by the resonance frequency widths and relatively long nuclear spin relaxation times (T1), ranging from 10 to 200 seconds for 23Na and 7Li. So there should be only electronic excitations in these compounds. Band-structure calculations of Li–Zn ordered versions of the lattices suggest a direct band-gap in both compounds, corresponding to an excitation from sulphur to zinc. The theoretical band-gaps amount to 2.54 eV for CsLiZnS2 and 1.85 eV for NaLiZnS2, and the steep edges in the density of states are found at 3.3 eV for both cases. As no Li–Zn ordering is observed by X-ray diffraction, there must be an inherent atomic disorder. By theoretical simulations, local Li–Zn anti-site orderings were introduced and the resulting electronic structure was evaluated. However, the simulated optical behaviour could only tentatively explain the spectroscopic data of Na(LiZn)S2; the orange red colour of Cs(LiZn)S2 must be an even more complex phenomenon, as the Li–Zn simulations were insufficient to explain the relatively strong optical activity in the range between 400 and 600 nm.
Na(LiB)S2 (B = Zn or Cd) can be obtained by solid-state reactions,5 but sulfides with larger alkali-metal ions instead of Na have only been synthesized by reacting carbonates and metals in a stream of H2S,6 which is a hazardous route that should not be scaled up for industrial purposes. For applications, several of the abovementioned oxides and sulfides do raise questions about environmental issues: In, Sn, Te, and Cd are generally unwanted for future applications. Therefore, we employ an alternative metathesis route to synthesize A(LiZn)S2 (A = alkali-metal) to minimize synthesis hazards and environmental problems as well as to characterize these layered sulfides further.
Na2S + Li2S + 2Zn + 2S + 2ACl → 2A(LiZn)S2 + 2NaCl (A = K, Rb, or Cs), |
For phase identification by X-ray diffraction, a Huber G670 Guinier camera with Co-Kα radiation (λ = 1.78897 Å) was employed. High resolution X-ray diffraction data were extracted with a STOE STADI_P using a Mo-Kα1 X-ray source (λ = 0.70926 Å) or Co-Kα radiation and Dectris Mythen 1K detectors. The WinXPOW software package8 was used to evaluate and compare the obtained diffraction data. A Rietveld simulation with the software JANA20069 was used to confirm the crystal structure of Na(LiZn)S2.
Scanning electron microscopy studies and basic elemental analyses were performed using a NOVA NanoSEM 200 (FEI), equipped with an EDX detector (AMETEK, Germany).
The melting behavior of Na(LiZn)S2 was investigated with an STD Q600 simultaneous thermal analyzer from TA Instruments using a corundum crucible for the sample as well as an empty twin crucible as the reference. During heating and cooling (10° min−1), the sample space was flushed with pure nitrogen (10 ml min−1).
The combination of inductively coupled plasma – optical emission spectroscopy (ICP-OES, ARCOS MV from SPECTRO) was used to determine the composition of selected bulk samples. A well-defined amount of the polycrystalline sample was dissolved in a strong alkali solution during microwave heating. After the solution was cooled again, H2O2 was added to oxidize S2− to SO42− for 24 h before performing the analyses.
Spectra in the ultraviolet-visible (UV-Vis) range were collected in a transmission mode using a Cary 4000 UV-Vis spectrophotometer from VARIAN Inc. Each sample was homogenized with optically pure BaSO4. An optimum signal was achieved using 10 mass% of the sample and 90 mass% BaSO4.
Nuclear magnetic resonance spectroscopy data were obtained using an Apollo spectrometer from TECMAG using a Magnex scientific magnet with a magnetic flux density of 7.0471 T. The cores in focus were 7Li (I = 3/2), 23Na (I = 3/2), and 133Cs (I = 7/2). The method for determining the spin relaxation (T1) was the free-induction-decay (FID) in the case of Li and Hahn spin-echo for the Na and Cs cores. To estimate the temperature dependence of the relaxation, spectra were obtained at room temperature (292 K) and 380 K.
All calculations were carried out within the PBE implementation10 of generalized gradient approximation (GGA) using the Full-Potential Local-Orbital (FPLO) code,11 version 18.00-52. Self-consistent calculations were carried out using scalar relativistic approximation.
To model the effects of disordered Li/Zn distribution on the electronic and optical properties, structural configurations with pseudo-random Li/Zn distribution as described in Appendix C were considered on a 2 × 2 × 2 supercell for Na(LiZn)S2 and on a 2 × 2 × 1 supercell for Cs(LiZn)S2. The supercell structures contain 40 atoms per unit cell for both the compounds and are considered in space group 1. These structural models also contain structures with different Li:
Zn, e.g. 3
:
5, ratios in each quasi-two-dimensional (LiZn)S2 layer. However, the stoichiometry of the compound was maintained, implying that deviations from the ordered distribution were compensated by the neighboring layer. In total, 15 configurations were obtained (labelled “Rn”) for each compound. For all the structural models, the internal parameters (atomic positions) were optimized such that the residual forces were smaller than 10 meV Å−1, using a k-mesh with 12 × 12 × 6 intervals. Atomic positions are only marginally shifted with respect to the initial positions, taken from the ordered structures, due to the very similar covalent radii of Li and Zn, 0.123 and 0.125 nm, respectively. We find that the minimum Li–S distances of all pseudo-random structures differ by only up to 2.5% and the minimum Zn–S distances by up to 3.5%. The DOS and the optical properties of the pseudo-random structures were obtained on k-meshes with 18 × 18 × 9 and 17 × 17 × 10 intervals, respectively, for Na(LiZn)S2 and Cs(LiZn)S2.
The optical properties were studied using the complex finite-frequency dielectric function ε(ω), where ω is the angular frequency of the incident photon. The imaginary part of the dielectric function, ε2(ω), was calculated up to ℏω = 10 eV for all the structural models using the “foptics” module of the FPLO code (plots not shown). The real part of the dielectric function, ε1(ω), was obtained by the Kramers–Kronig transformation. The absorption coefficient was subsequently obtained using the well-known relation:12
α(ω) = 21/2(ω/c){[(Reε(ω))2 + (Imε(ω))2]1/2 − Reε(ω)}1/2. |
For supercell structures with pseudo-random Li/Zn distribution, the average dielectric function and absorption coefficient were obtained by averaging over different considered structural configurations. Ensemble averaged 〈α(ω)〉 was obtained by averaging over the ordered model structure as well as the selected supercell structures. For this, all the structural configurations were considered equally likely. Further, orientational disorder of the crystallites was accounted for by directional averaging of the absorption coefficient. It was explicitly checked that the final results for ensemble averages 〈ε(ω)〉 and 〈α(ω)〉 do not depend on the specific sequence of the ensemble averaging: alternatively, the ensemble averaging was carried out by first obtaining the ensemble and then directionally averaging 〈ε2(ω)〉 for the ordered model structure as well as the selected supercell structures.13 Subsequently, 〈ε1(ω)〉 and 〈α(ω)〉 were obtained, which were found to be numerically identical to the results obtained using the other procedure.
The compounds Na(LiZn)S2 and Cs(LiZn)S2 were relatively phase pure, as estimated from X-ray diffraction data (Fig. 3), so these compounds were used for further analyses. The lack of super-structure reflections suggests that there are no long-range atomic ordering phenomena between Li and Zn in the (LiZn)S2 layer, despite the possibility of formation of stripe orderings, or similar. This means, in reciprocity, that there must be random domain formation (or clustering) of Li and Zn in the (LiZn)S2 layer, which will be very important for later discussions here.
Naturally, it can be argued that the metathesis reactions might be only partial for Cs(LiZn)S2. However, elemental analyses on individual crystallites, by EDX analyses during scanning electron microscopy studies, and the bulk sample, done by ICP-OES analyses, support a complete metathesis reaction in forming Cs(LiZn)S2: EDX data suggest Na0.88(LixZn1.00)S2.12 and Cs0.92(LixZn1.08)S2.00, when scaling the relative amounts up to a sum of 4, not counting Li as it cannot be reliably quantified by this technique. ICP-OES data, containing signals from Li, suggest Na0.99(3)(Li1.0(1)Zn1.04(1))S1.96(8) and Cs0.95(7)(Li1.0(1)Zn1.06(6))S1.9(1), when scaling to a sum of ∼5. All the observed compositions are convincingly close to the intended ones.
![]() | ||
Fig. 4 Differential scanning calorimetry–thermal gravimetric analysis combined data of an X-ray pure sample of Na(LiZn)S2 (Fig. 3a). Arrows along the curves indicate how the measurement progressed. * denotes an instrumental error. |
Despite the fact that the two optically measured compounds have slight differences in crystal structures, it is not expected that the Cs-compound has stronger absorption. In contrast, any ionic Cs compound is expected to have more ionicity than the corresponding Na compound because of the higher electropositivity of Cs compared to Na. Following this argument, it could be expected that we have a larger energy gap between the valence and conduction bands in the Cs compound. However, the strong optical activity of the Cs compound indicates that this simple deduction is misleading. Naturally, minor colouring effects can arise from F-centres (trapped electrons in lattice defects), but those effects are of a much lower magnitude. To understand the situation better in these two compounds, local probe NMR spectroscopy was employed.
The very broad satellite intensities in the 7Li (I = 3/2) spectrum of Na(LiZn)S2 (Fig. 6a) indicate that there is some disorder in the system, which might agree with the suggested atomic disorder at the Li–Zn positions. The line-width of the main intensity is 5.4 kHz (292 K) and it does not change at a higher temperature (380 K), which is in line with a rigid lattice. However, in Cs(LiZn)S2, the central line-width is 4.3 kHz but there are clear satellites that appear with a splitting of 12 kHz (Fig. 6b). This is probably due to a second hyperfine field, indicative of more prominent ordering phenomena, although short-ranged; there might be a high abundance of next-neighboring Li–Li in the (LiZn)S2 layer.
According to 23Na NMR data (Fig. 7a), Na is moving more than Li as T1 is significantly smaller. Even the two expected satellites from this I = 3/2 species are too broad to be detected. Yet, the relaxation is still too long to be comparable with typical Na–ion conductors. For example, 23Na in β-alumina has a T1 value of about 0.01 s.16 In contrast, 133Cs has I = 7/2 and should exhibit six satellites, which are all present (Fig. 7b) although slightly asymmetric in their intensity distribution. The reasons for this asymmetry and that the main intensity is found significantly shifted from the expected frequency are not understood.
As both NMR spectra and X-ray diffraction data suggest disorder in the (LiZn)S2 layers, a theoretical approach was tried to understand the strange color changes when inserting different alkali-metals in position A in A(LiZn)S2 layers.
![]() | ||
Fig. 8 The structures and Brillouin zones for NaLiZnS2 (a) and (c) and CsLiZnS2 (b) and (d). Band-structures along high symmetry lines for the ordered structures of (e) NaLiZnS2 and (f) CsLiZnS2. |
Each simulation cell contains 40 atoms, i.e., close in size to the structures shown in Fig. 8, where two (LiZn)S2 layers are involved. Crystallographically, this means that a supercell of NaLiZnS2 contains eight unit cells (2 × 2 × 2) and a supercell of CsLiZnS2 four unit cells (2 × 2 × 1). In addition, the simulation allowed the layers to contain different ratios of Li:
Zn, like 3
:
5, but the total composition was always maintained, meaning that the neighbouring layer has a complementary composition (5
:
3). Note that these calculations are supposed to represent local domains and not complete structures. Hence, only the “local” direct band-gap is estimated in relation to the change in total energy, as compared to the ordered structures (1 × 1 × 1) shown in Fig. 8. The total energies, band-gaps, band-edge positions (the “edge position” is related to a remarkable upturn of the DOS in the lower conduction band; see Fig. 9 and the related discussion below), and Li:Zn distributions of these structure models are shown in Tables 1 and 2.
![]() | ||
Fig. 9 Density-of-states (DOS) plots of the cation ordered compounds (a and c) and of three exemplary structures with pseudo-random Li–Zn distributions (b and d) to be compared with Tables 1 and 2. CBB indicates the ranges of band-gaps within all the considered pseudo-random structures and “edge” indicates the ranges of a prominent upturn of the DOS in the lower conduction band. The atom-projected partial DOS values are also shown for the ordered compounds. |
Model | ΔE (meV fu−1) | Gap (eV) | Edge (eV) | Li![]() ![]() |
---|---|---|---|---|
Colour coding agrees with the curves in Fig. 9b. “fu” is the formula unit. “Rn” denotes 15 different pseudo-random 2 × 2 × 2 structures, ordered according to their total energies. | ||||
1 × 1 × 1 | 0 | 1.85 | 3.3 | 4![]() ![]() ![]() ![]() |
R01 | −27 | 2.39 | 3.7 | 4![]() ![]() ![]() ![]() |
R02 | −9 | 2.38 | 3.4 | 4![]() ![]() ![]() ![]() |
R03 | 1 | 1.97 | 3.2 | 4![]() ![]() ![]() ![]() |
R04 | 14 | 1.76 | 2.9 | 3![]() ![]() ![]() ![]() |
R05![]() |
30 | 1.78 | 2.9 | 3![]() ![]() ![]() ![]() |
R06![]() |
33 | 1.77 | 2.9 | 3![]() ![]() ![]() ![]() |
R07 | 38 | 1.71 | 2.8 | 3![]() ![]() ![]() ![]() |
R08 | 40 | 1.62 | 2.8 | 3![]() ![]() ![]() ![]() |
R09 | 44 | 1.61 | 2.7 | 3![]() ![]() ![]() ![]() |
R10 | 53 | 1.73 | 2.8 | 3![]() ![]() ![]() ![]() |
R11 | 54 | 1.71 | 2.8 | 3![]() ![]() ![]() ![]() |
R12 | 61 | 1.61 | 2.7 | 3![]() ![]() ![]() ![]() |
R13 | 213 | 0.84 | 1.8 | 2![]() ![]() ![]() ![]() |
R14![]() |
222 | 0.91 | 1.9 | 2![]() ![]() ![]() ![]() |
R15 | 521 | Metal | — | 1![]() ![]() ![]() ![]() |
Model | ΔE (meV fu−1) | Gap (eV) | Edge (eV) | Li![]() ![]() |
---|---|---|---|---|
Colour coding agrees with the curves in Fig. 9d. “fu” is the formula unit. “Rn” denotes 15 different pseudo-random 2 × 2 × 1 structures, ordered according to their total energies. | ||||
1 × 1 × 1 | 0 | 2.54 | 3.3 | 4![]() ![]() ![]() ![]() |
R01 | 11 | 2.79 | 3.6 | 4![]() ![]() ![]() ![]() |
R02![]() |
11 | 2.81 | 3.7 | 4![]() ![]() ![]() ![]() |
R03 | 30 | 2.73 | 3.7 | 4![]() ![]() ![]() ![]() |
R04![]() |
37 | 2.84 | 3.6 | 4![]() ![]() ![]() ![]() |
R05 | 49 | 2.70 | 3.8 | 4![]() ![]() ![]() ![]() |
R06![]() |
49 | 2.80 | 3.7 | 4![]() ![]() ![]() ![]() |
R07 | 55 | 2.65 | 3.6 | 4![]() ![]() ![]() ![]() |
R08 | 96 | 2.03 | 2.9 | 3![]() ![]() ![]() ![]() |
R09 | 105 | 1.85 | 2.4 | 3![]() ![]() ![]() ![]() |
R10 | 105 | 2.04 | 2.9 | 3![]() ![]() ![]() ![]() |
R11 | 121 | 1.98 | 3.0 | 3![]() ![]() ![]() ![]() |
R12 | 121 | 1.99 | 2.8 | 3![]() ![]() ![]() ![]() |
R13 | 122 | 1.97 | 3.0 | 3![]() ![]() ![]() ![]() |
R14 | 122 | 1.98 | 2.8 | 3![]() ![]() ![]() ![]() |
R15 | 323 | 0.82 | ? | 2![]() ![]() ![]() ![]() |
The total energies Etot of the pseudo-random structures deviate from those of the ordered structure by −5 up to +100 meV atom−1. These energy differences ΔE = Etot [Rn] − Etot[1 × 1 × 1] clearly correlate with the random distribution of the Li and Zn atoms (Tables 1 and 2). Consider first the Na-homologue (Table 1): if equal numbers of Li and Zn atoms are situated in the upper and in the lower half of the supercell (4:
4), ΔE = −5 to 0 meV atom−1; for distributions 3
:
5(5
:
3), ΔE = 3 to 12 meV atom−1; for distributions 2
:
6 (6
:
2), ΔE = 40 meV atom−1; and for 1
:
7(7
:
1), ΔE is about 100 meV atom−1. In particular, the total energies of R13, R14, and R15 are significantly higher than those of other structures. This is probably related to a less favourable distribution of charge (Madelung energy) and/or atomic volume (strain energy). Thus, we consider R13, R14, and R15 as less probable realizations and will disregard these structures in the further analysis. Note that also the gap width and the edge position of these structures deviate essentially from those of the other structures. In particular, R15 with Li
:
Zn distribution of 1
:
7(7
:
1) is metallic.
On the other hand, for the Cs-homologue, the total energies of the pseudo-random structures are higher than that of the ordered structure by 2 to 60 meV per atom. The latter value is only obtained for the structure R15 with Li:Zn distribution of 2:
6(6
:
2), which is considered as less probable realization, due to the much higher energy than those of the other structures. Hence, R15 will be disregarded in further analyses. Note that the distinction between a small DOS at the conduction band bottom and a steep edge in the Cs system is less clear than that in the Na system.
The densities of states for the ordered structural models and for several pseudo-random structural configurations are shown in Fig. 9. In all cases, the top of the valence band is dominated by sulphur states and the bottom of the conduction band by zinc states. This would mean that the electron transfer requiring the least energy is the ligand-to-metal charge transfer from sulphur to zinc. Note that in both ordered structure models the lower conduction band DOS starts with a weak tail and shows a strong upturn at 3.3 eV. The position of these upturns is close to the edge positions shown in the optical spectra (Fig. 5). In the pseudo-random structures, the edge positions are distributed (indicated by a line with arrows), as are the band-gaps.
We now turn to the consideration of the optical properties. For the supercell structures, the dielectric functions are considered excluding the three (one) members for the Na (Cs) system that were considered unlikely, as discussed above.
The fully ordered Li–Zn lattices result in rather weak absorptions that stretch into the visible range, especially far for the Na-homologue (Fig. 10a). As the crystal structures of these sulphides are anisotropic, with one unique crystallographic axis in both cases, the optical properties are displayed with xx and zz components (Fig. 10b and c): at high photon energies, the optical anisotropy is large but the effect is faint in the visible range. As the optical response depends on the joint density of states between the initial and final states, the optical absorption at higher wavelengths (lower energies) is likely proportional to the DOS at the conduction band edges for the considered ensembles since the DOS at the valence band edges is large in all cases. Further, the edge positions and the band-gap values for both the compounds, listed in Tables 1 and 2, are highly correlated. Consequently, when examining the individual optical absorptions for the different Li–Zn orderings as individual curves and as a mean (Fig. 10d and e), it is obvious that the absorption in the visible range is only slightly affected. Comparing the two homologues, and especially their ensembles and individual orderings (Fig. 10f), the strongest effect is observed in the Na-homologues which might very well explain the two shoulders in Fig. 5. On the other hand, none of the theoretically calculated optical properties, based on local Li–Zn ordering phenomena, agrees with the very strong orange red colour of Cs(LiZn)S2. Interestingly, however, the ordered CsLiZnS2 structure displays relatively intense features in the 250–350 nm range (Fig. 10a and c) which eventually subside due to the ensemble averaging procedure. The electronic and optical properties, while being sensitive to the choice of the functional, shows approximately a constant shift when the local density approximation (LDA) within the Perdew–Wang implementation17 is employed instead (see Appendix D for details). It is, therefore, likely that the characteristic features of the electronic and optical properties of the A(LiZn)S2 compounds are not influenced drastically between LDA and GGA. On the other hand, it is plausible that the sample contains nano-domains with Li–Zn ordering, e.g., as phase separated regions, which would be consistent with the presence of satellite peaks in the NMR data discussed above (see Fig. 6b). A detailed characterization of the samples to rule out these possibilities is, however, beyond the scope of the present work. Another interesting possibility could be that the relative Li–Zn (anti-site) defect fraction in the Cs compound is smaller, to account for which, however, requires a weighted averaging over the considered ensembles. It should also be noted that effects beyond the Li–Zn disorder cannot be ruled out. For example, phonon-assisted indirect transitions might account for the measured edges at 3.3 eV, which coincide with the edges in the related DOS of the ordered structures.
Band-gap changes due to chemical disorder, also called band-bowing, have been observed in several other systems: ZnO–GaN,18 (1−x)ZnGeN2−2xGaN,19 and NixMg1−xO.20 In contrast, the atomic disorder in the A(LiZn)S2 phases is not inherited by a solid solution and does not directly result in composition dependent band-bowing. Instead, the Li–Zn shared position is an intrinsic property due to the stoichiometry and the oxidation state stability of the involved elements. Hence, the amount of disorder is probably due to subtle effects like lattice strains and synthesis conditions. In the series A(LiZn)S2, the band-gap seems to widen as the ion size on the A-site is increased based on DFT calculations and partly on measurements. The fact that the distance between (LiZn) sites increases with the size of A, should lower the Coulomb forces between the (LiZn) sites. Lower electrostatic forces allow for more Li–Zn disorder, including nano-domain formation, which also causes a stronger optical band-bowing effect. This agrees with the calculations for pseudo-random structures where the gap width shows a broader distribution in the case of the Cs compound (1.0 eV) than in the case of the Na compound (0.8 eV). A similar scenario, with anti-site defects, was suggested for the case of (ZnGe)N2, where the band-gap was narrowed by about 0.5 eV.21 In theory, the band-gap can be completely closed in A(LiZn)S2 (see for example R15 in Table 1) but the local energy increase (strain and Coulomb repulsions) for such an extreme case might be too high to be statistically relevant.
A. NaLiZnS2 (SG 156, P3m1) | |
---|---|
Param. | Value |
a, b (Å) | 3.9711 |
c (Å) | 6.7186 |
Atom | Wyckoff position | Fractional coordinates |
---|---|---|
Na | 1a | (0, 0, 0) |
Li | 1b | (1/3, 2/3, 0.3298) |
Zn | 1c | (2/3, 1/3, −0.4108) |
S | 1b | (1/3, 2/3, −0.2863) |
S | 1c | (2/3, 1/3, 0.2439) |
B. CsLiZnS2 (SG 119, I![]() |
|
---|---|
Param. | Value |
a, b (Å) | 4.092 |
c (Å) | 13.974 |
Atom | Wyckoff position | Fractional coordinates |
---|---|---|
Cs | 2a | (0, 0, 0) |
Li | 2c | (1/2, 0, 3/4) |
Zn | 2d | (1/2, 0, 1/4) |
S | 4e | (0, 0, 0.3432) |
![]() | ||
Fig. 11 The band-structures of (a) KLiZnS2 and (b) RbLiZnS2, featuring indirect band-gaps of 2.25 eV and 2.42 eV, respectively. The Brillouin zone is the same as that for CsLiZnS2 (see Fig. 8). The insets show a zoomed view of the valence band maxima close to the Γ point. |
System | GGA | LDA |
---|---|---|
NaLiZnS2 (SG 156) | 1.85 | 1.60 |
KLiZnS2 (SG 119) | 2.25 | 2.01 |
RbLiZnS2 (SG 119) | 2.42 | 2.15 |
CsLiZnS2 (SG 119) | 2.54 | 2.29 |
![]() | ||
Fig. 12 Absorption coefficient for the ordered structures of Na and Cs compounds within LDA and GGA. |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2022 |