Next Article in Journal
Novel, Simple and Low-Cost Preparation of Ba-Modified TiO2 Nanotubes for Diclofenac Degradation under UV/Vis Radiation
Previous Article in Journal
All-in-One Self-Powered Human–Machine Interaction System for Wireless Remote Telemetry and Control of Intelligent Cars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dimension-Dependent Bandgap Narrowing and Metallization in Lead-Free Halide Perovskite Cs3Bi2X9 (X = I, Br, and Cl) under High Pressure

1
Shandong Provincial Key Laboratory of Optics and Photonic Device, Collaborative Innovation Center of Light Manipulations and Applications, School of Physics and Electronics, Shandong Normal University, Jinan 250014, China
2
School of Electronic and Information Engineering (Department of Physics), Qilu University of Technology (Shandong Academy of Sciences), Jinan 250353, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(10), 2712; https://doi.org/10.3390/nano11102712
Submission received: 26 September 2021 / Revised: 9 October 2021 / Accepted: 12 October 2021 / Published: 14 October 2021
(This article belongs to the Special Issue Nanostructured Perovskites—Material, Physics and Applications)

Abstract

:
Low-toxicity, air-stable cesium bismuth iodide Cs3Bi2X9 (X = I, Br, and Cl) perovskites are gaining substantial attention owing to their excellent potential in photoelectric and photovoltaic applications. In this work, the lattice constants, band structures, density of states, and optical properties of the Cs3Bi2X9 under high pressure perovskites are theoretically studied using the density functional theory. The calculated results show that the changes in the bandgap of the zero-dimensional Cs3Bi2I9, one-dimensional Cs3Bi2Cl9, and two-dimensional Cs3Bi2Br9 perovskites are 3.05, 1.95, and 2.39 eV under a pressure change from 0 to 40 GPa, respectively. Furthermore, it was found that the optimal bandgaps of the Shockley–Queisser theory for the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites can be reached at 2–3, 21–26, and 25–29 GPa, respectively. The Cs3Bi2I9 perovskite was found to transform from a semiconductor into a metal at a pressure of 17.3 GPa. The lattice constants, unit-cell volume, and bandgaps of the Cs3Bi2X9 perovskites exhibit a strong dependence on dimension. Additionally, the Cs3Bi2X9 perovskites have large absorption coefficients in the visible region, and their absorption coefficients undergo a redshift with increasing pressure. The theoretical calculation results obtained in this work strengthen the fundamental understanding of the structures and bandgaps of Cs3Bi2X9 perovskites at high pressures, providing a theoretical support for the design of materials under high pressure.

1. Introduction

Organic–inorganic hybrid perovskites have attracted much attention from the researcher community because of their remarkable photoelectric properties, including high absorption coefficients in the visible light region, tunable bandgaps, high quantum yields, high carrier mobilities, and low effective carrier quality [1,2,3,4,5,6]. In addition, their processing is economical and utilizes simple solution treatments [7,8]. The power conversion efficiency (PCE) of perovskite solar cells (PSCs) has increased dramatically from 3.8% in 2009 to 25.5% [9,10]. Lead hybrid perovskites have the outstanding photovoltaic properties and a high PCE. However, the high toxicity of lead is a major challenge for the large-scale fabrication and commercialization of lead-based PSCs. Recently, nontoxic, all-inorganic, lead-free Bi-based perovskites have attracted significant attention owing to their stability, high photoluminescence, high quantum yield, and tunable bandgaps [11,12,13,14,15,16,17]. They have been widely used in applications, such as memory devices, photodetectors, solar cells, and X-ray detectors [18,19,20,21,22,23,24,25,26]. However, many studies have shown that there are some challenges to overcome for the use of Cs3Bi2X9 (X = I, Br, and Cl) perovskites in photovoltaic devices. The large bandgap (>2 eV) is the most important factor causing a low PCE, which limits their absorption efficiency and carrier transport performance. According to the Shockley–Queisser theory, a semiconductor with a bandgap in the range of 1.3–1.5 eV is an ideal material for solar cells. Although the PCE of Bi-based PSCs has improved slightly with the improvement of the thin-film technology, it still lags far behind that of lead-based perovskites. The traditional chemical modification cannot overcome this inherent limit. Therefore, tuning the bandgap with the aim of improving the photovoltaic performance of the Cs3Bi2X9 perovskites has become a key challenge.
High pressure (HP), which is a non-polluting tuning method, is widely used to modulate the physical and chemical properties of materials without changing their chemical composition [27,28]. In recent years, the properties of halide perovskites have been widely studied under HP; the properties studied include piezochromism, bandgap engineering, structural phase transitions and optical properties, [29,30,31,32]. In addition, various novel physical phenomena have been observed under HP conditions. For example, the organic–inorganic hybrid perovskite nanocrystals present the comminution and recrystallization under HP and exhibit higher photoluminescence quantum yield and a shorter carrier lifetime [29]. A reversible amorphization has been observed for the CH3NH3PbBr3 perovskite under a pressure of approximately 2 GPa; during this transition, the resistance increased by five orders of magnitude, and the material still retained its response to the visible light and semiconductor characteristics up to a pressure of 25 GPa [33]. Notably, pressure-induced structural changes and optical properties are reversible upon decompression, and a semiconductor–metal transition can be observed at 28 GPa [34]. It is notable that there are few studies on Cs3Bi2X9 perovskite systems of different dimensions at HP [34,35], and no reports exist on the one-dimensional halide perovskite Cs3Bi2Cl9.
In this work, we calculated the lattice constants, band structures, density of states (DOS), and optical properties of the one-dimensional perovskite Cs3Bi2Cl9 via the density functional theory (DFT) under HP for the first time and compared with the zero-dimensional perovskite Cs3Bi2I9 and the two-dimensional perovskite Cs3Bi2Br9. We discussed the relationship between the bandgap of the Cs3Bi2X9 perovskites and the HP and focused on the optimal bandgap of the Shockley–Queisser theory of the Cs3Bi2X9 perovskites. The Cs3Bi2I9 perovskite completed the transition from semiconductor to metal at 17.3 GPa, this finding indicated that HP is an effective means to induce the semiconductor–metal transition. Moreover, it was found that the lattice constants and bandgaps of the Cs3Bi2X9 perovskites are dependent upon dimension, that is, the changes in the lattice constants and bandgap gradually decrease as the dimension increases from zero to two under the same pressure. Our calculated results obtained in this work strengthen the basic understanding of different structures of the Cs3Bi2X9 (X = I, Br and Cl), providing theoretical guidance for the structure and bandgap regulation of Bi-based perovskites under HP.

2. Computational Model and Method

The DFT was performed in the Vienna Ab-initio Simulation Package (VASP) using the projected augmented wave (PAW) framework [36,37]. DFT is derived from the Schrodinger equation under the Born–Oppenheimer approximation, described by the Hohenberg–Kokn theorem and the Kohn–Sham equation. The pseudopotential is a hypothetical potential energy function used in place of the inner electron wave function to reduce the computation. The electron exchange–correction function was obtained via the generalized gradient approximation (GGA) parameterized using the Perdew–Burke–Ernzerhof (PBE) formalism [38]. The cut-off energy of the plane wave was set to 500 eV [34]. The convergence criteria for the energy and force were set to 10−5 eV and 0.01 eV/Å, respectively. The Brillouin zone integration were sampled with 4 × 4 × 4, 4 × 4 × 4, and 4 × 4 × 2 Gamma-pack k-point meshes during the structure optimization of Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9, respectively. The entire optimization of the structures was completely relaxed [39]. The valence electronic configurations of the Cs, Bi, I, Br, and Cl atoms are 5s25p66s1, 5d106s26p3, 5s25p5, 4s24p5, and 3s23p5, respectively. The Γ Brillouin zone center has a highly symmetric path, with coordinates Γ (0, 0, 0) to the M (0.5, 0, 0), K (0.333, 0.333, 0), Γ (0, 0, 0), A (0, 0, 0.5), L (0.5, 0, 0.5), and H (0.333, 0.333, 0.5). The spin–orbit coupling is not considered due to the high computational cost. Previous studies have shown that higher levels of calculation (including the spin–orbit coupling, the GW method, or hybrid functionals) obtain more accurate bandgaps; However, they induce little change in the band structure of the heavy-metal halide perovskites [40].

3. Results and Discussion

The structures of the three halogenated perovskite crystals Cs3Bi2X9 (X = I, Br, and Cl) are shown in Figure 1. The one-dimensional Cs3Bi2Cl9 perovskite has an orthogonal-crystal structure; the two-dimensional Cs3Bi2Br9 perovskite and the zero-dimensional Cs3Bi2I9 perovskite have hexagonal cells [41,42]. The numbers of atoms in the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites in the primitive cell are 28, 14, and 56, respectively. The Bi atom is located at the center of the octahedron in the Cs3Bi2X9 perovskites and is surrounded by six halogen atoms. Unlike in lead-based perovskites, in the Cs3Bi2Br9 and Cs3Bi2Cl9 perovskites, two octahedrons share one X atom (X = Br or Cl), whereas two octahedrons share three X atoms in the Cs3Bi2I9 perovskite. X atoms with different sizes and the Bi atoms form a novel double-perovskite structure. Figure 2 shows the changes in the lattice parameters and volume of the Cs3Bi2X9 perovskites under HP. In the primary cell of the Cs3Bi2I9 and Cs3Bi2Br9 perovskites, the lattice constants a and b are equal. The lattice constant and volume of Cs3Bi2X9 clearly decrease with an increase in pressure, and the slope also decreases gradually. As is well established, pressure induces a reduction in the lattice constant. From a microscopic point of view, the pressure shrinks the distance between two atoms. The strong Coulomb force makes it increasingly difficult to further compress the material as the pressure increases.
Based on the semiconductor theory, analyzing the band structure and the DOS very closely, the Fermi level is important for establishing the possible applications of a material in the photoelectric and photovoltaic fields. We therefore calculated the band structure near the Fermi level (from −5 eV to +5 eV) under different pressures. The calculated bandgaps at a pressure of 0 GPa of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 are 2.38, 2.60, and 3.08 eV, respectively. The scissor value approach was used to establish the band structure of the Cs3Bi2X9 perovskites in order to obtain the accurate value for the Cs3Bi2X9 perovskites when they reached the optimal bandgap of the Shockley–Queisser theory and the transition from semiconductor to metal; this methodology overcame the limitations of the GGA–PBE calculation method and has already been applied in the study of the photoelectric properties of the CsSnCl3 perovskite at HPs [40]. Many experimental studies have been conducted regarding the bandgaps of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites. Among these, Daniel R. Gamelin et al. synthesized Cs3Bi2X9 perovskites nanocrystals via the thermal injection method, and the bandgaps of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites were measured as 2.07, 2.76, and 3.26 eV, respectively [43]. Therefore, the scissors values of −0.31eV, 0.16eV, and 0.18eV were used for the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites, respectively.
Figure 3 shows the change in the band structure of the Cs3Bi2I9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d). Without external pressure, the conduction band minimum (CBM) and the valence band maximum (VBM) of Cs3Bi2I9 perovskite are located at the Γ- and M-points, respectively. We found the Cs3Bi2I9 perovskite to be an indirect bandgap material, which is consistent with the existing literature [44]. The semiconductor with the optimized band gap energy of 1.34 eV is critical to achieve the efficiency limit of 33.7% based on the Shockley–Queisser theory [34]. In Figure 3, it can be seen that the bandgap of the Cs3Bi2I9 perovskite decreases sharply with the increase in pressure and reaches its optimal bandgap value given by the Shockley–Queisser theory at 2–3 GPa. In addition, with the high pressure further increasing, the CBM continues to decrease and the VBM moves from the M-point to near the K-point. The Cs3Bi2I9 perovskite completed the transition from semiconductor to metal at 17.3 GPa, and this finding indicated that HP is an effective means to induce the semiconductor–metal transition. The band structure of the Cs3Bi2Br9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d) is presented in Figure 4. As can be seen from Figure 4, when the pressure is 0 GPa (a), the calculated band structure of the Cs3Bi2Br9 perovskite is consistent with the results reported by Brent C. Melot et al. [45] They found a low-lying 2.52 eV indirect transition as well as a slightly larger direct gap of 2.64 eV, which are essentially in agreement with our calculated results. This structural characteristic becomes increasingly evident with the increase in pressure. The Cs3Bi2Br9 perovskite reaches the optimal bandgap given by the Shockley–Queisser theory at 21–26 GPa, which means that the electron transitions from the valence band to conduction band become easier. Figure 5a–d shows the change of the band structure of Cs3Bi2Cl9 perovskite with HP. The Cs3Bi2Cl9 perovskite reaches the optimal bandgap in the range of 25–29 GPa; it is an indirect bandgap material. The CBM is at the Γ- point, and the VBM moves from the Γ-point to the Y-point, which is consistent with the results reported previously [41]. However, it was found that the Cs3Bi2Br9 and Cs3Bi2Cl9 perovskites did not metallize under HP despite the pressure reaching 40 GPa in both cases, which may be related to their unique structure. The bandgap changes in the Cs3Bi2X9 perovskites under a range of HP (0, 2, 4, 6, 8, 10, 20, 30, and 40 GPa) are shown in Supplementary Figures S1–S3.
Figure 6 shows the changes in the bandgaps of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites under HP. It can be seen that the bandgap of the Cs3Bi2X9 perovskites decreases with the increase in pressure (Figure 6a); the bandgap of the Cs3Bi2I9 perovskite takes a negative value, which is a typical indicator of metallic behavior. Figure 6b shows the change in the bandgap of the Cs3Bi2X9 perovskites after using the scissor values under HP. When the pressure is varied from 0 to 40 GPa, the bandgap differences of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites are 3.05, 1.95, and 2.39 eV, respectively. It was found that with the increase in dimension, the bandgap differences of the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites decrease in turn, which indicates that the bandgap of the Cs3Bi2X9 perovskites depends on dimension.
To explain the dimension-dependent bandgap of the Cs3Bi2X9, we investigate the construction of these perovskites. Based upon the primitive cell of Cs3Bi2I9, Cs3Bi2Br9 and the Cs3Bi2Cl9 perovskites, the 2 × 2 × 1, 2 × 2 × 2, and 3 × 1 × 1 supercells are shown in Figure 7a–c, respectively. In Figure 7d, ΔL represents the difference in the lattice constants of the Cs3Bi2X9 perovskites between 0 and 40 GPa along the a, b, and c coordinate axes. In general, for centrosymmetric perovskites, the ΔL values along the a- (ΔLa), b- (ΔLb), and c- (ΔLc) axes are equal under HP. However, we found that ΔLa, ΔLb, and ΔLc for the double perovskites were not equal; in other words, ΔL is anisotropic along the a-, b-, and c-axes. The ΔLa, ΔLb, and ΔLc values for the Cs3Bi2I9 perovskite (zero-dimensional) are 1.57, 1.57, and 5.2. It can be seen that ΔLc is much larger than both ΔLa and ΔLb. The reason for this difference is that the Bi2I93− frame of the double-perovskite is continuous along the a- and b-axes but not along the c-axis. Along the c-axis, only the Cs+ atoms are above or below the Bi2I93− frame, which indicates that the change in the lattice constant along the c-axis is bigger than that observed along the a- and b-axes when the zero-dimensional Cs3Bi2I9 perovskite is placed under HP.
The previous analysis shows that if the double-perovskite frame expands regularly in one direction, the ΔL in this direction will be smaller than in the other directions under HP. This indicates that the lattice constants of the Cs3Bi2X9 perovskites are dependent on dimension. In general, the lattice constant and the bandgap decrease as the pressure increases [34]. It has also been found that the changes in the bandgap of the zero-dimensional Cs3Bi2I9, one-dimensional Cs3Bi2Cl9, and two-dimensional Cs3Bi2Br9 perovskites between 40 and 0 GPa are 3.05, 2.39, and 1.95 eV, respectively. These results illustrate that the bandgaps of the Cs3Bi2X9 perovskites are also dependent upon dimension; that is, the changes in the bandgap decrease gradually as the dimension increased from zero to two under the same pressure.
The partial density of states (PDOS) of the zero-dimensional Cs3Bi2I9 (a), one-dimensional Cs3Bi2Cl9 (b), and two-dimensional Cs3Bi2Br9 (c) perovskites were shown in Figure 8. It can be seen that the VBM of the Cs3Bi2X9 perovskites is dominated by p-X states, whereas the CBM is dominated by the p-Bi and p-X states (see Figure 8a–c). The changes in the DOS under HP are shown in Figure 8d–f. It is clear that many valence bands in the Cs3Bi2X9 perovskites move to a deep level, and the conduction bands approach to the FE with an increase in pressure for Cs3Bi2X9 perovskites; the shift of the conduction bands under the HP will induce the changes in the bandgap. In contrast to those of the Cs3Bi2Br9 and Cs3Bi2Cl9 perovskites, the forbidden band width of the Cs3Bi2I9 gradually decreases and subsequently disappears, which indicates that the Cs3Bi2I9 perovskite is no longer a semiconductor; it becomes a metal at 17.3 GPa. This conclusion is agreement with the calculated results for the band structure.
A large absorption coefficient is of great significance in photoelectric and photovoltaic applications. Such a property improves the PCE of solar cells and the luminous efficiency. The absorption coefficient is usually described by the dielectric function according to the following expression [46]:
α = 2 ω [ ( ε 1 2 ( ω ) + ε 2 2 ( ω ) ) 1 / 2 ε 1 ( ω ) 2 ] 1 / 2
where the ω is the frequency of light, and ε 1 and ε 2 are the real and imaginary parts of the dielectric function, respectively. The calculated ε 1 and ε 2 values of the Cs3Bi2X9 perovskites are presented in Supplementary Figures S4–S6 along the a-, b-, and c-axes. The static dielectric function of the Cs3Bi2X9 perovskites increases gradually with increasing pressure from 0 to 40 GPa. The ε 2 value is closely related to the optical absorption and usually used to describe the absorption behavior of materials.
The calculated absorption coefficients of the Cs3Bi2X9 perovskites are shown in Figure 9 and Figure 10 along the a-, b-, and c-axes. It was found that the Cs3Bi2I9 perovskite has the same absorption coefficient along the a- and b-axes with the increase in pressure, as is the case for the Cs3Bi2Br9 perovskite. The absorption coefficients of the Cs3Bi2Cl9 perovskite are unequable along the a-, b-, and c-axes under HP. The ΔL value of the Cs3Bi2X9 perovskites in Figure 7d is consistent with these behaviors, which suggests that the changes in the structure of the Cs3Bi2X9 perovskites under HP will affect the optical properties considerably. The Cs3Bi2X9 perovskites exhibit a redshift with the increase in pressure, which indicates that the Cs3Bi2X9 can absorb the low-energy photons. Moreover, the absorption coefficients of the Cs3Bi2X9 perovskites increase gradually in the ultraviolet region as the pressure increases from 0 to 40 GPa. It was also found that the Cs3Bi2X9 perovskites have a large absorption coefficient in the visible region (on the order of 105 cm−1). Therefore, the Cs3Bi2X9 perovskites are an attractive candidate in applications of photoelectric and photovoltaic devices.

4. Conclusions

In summary, we investigated the lattice constants, band structure, DOS, and optical absorption of the cesium bismuth iodide Cs3Bi2X9 (X = I, Br and Cl) perovskites under HP by using the DFT. It was found that the optimal bandgap of the Shockley–Queisser theory for the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites can be obtained at 2–3 GPa, 21–26 GPa, and 25–29 GPa, respectively. The changes in the bandgap of Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites are 3.05, 1.95, and 2.39 eV under a pressure of 40 GPa, respectively. The Cs3Bi2I9 perovskite was found to transform from a semiconductor into a metal at 17.3 GPa. Furthermore, the dimension-dependent lattice constants, unit-cell volumes, and bandgaps of the Cs3Bi2X9 perovskites were studied. Our calculations show that HP is an effective way to tune the photovoltaic and optoelectronic properties of the Cs3Bi2X9 perovskites by modifying the crystal structure, which provides a promising method for material design and applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11102712/s1. Table S1: Summarized bandgap values of Cs3Bi2X9 perovskites under high pressure; Figure S1: Calculated band structures of the Cs3Bi2I9 perovskite under the different pressures (ai); Figure S2: Calculated band structures of the Cs3Bi2Br9 perovskite under the different pressures (ai); Figure S3: Calculated band structures of the Cs3Bi2Cl9 perovskite under the different pressures (ai); Figure S4: Real part of the dielectric function of the Cs3Bi2I9 perovskite along the a- and b-axes (a), and the c-axis (b) as a function of pressure (from 0 to 40 GPa). Imaginary part of the dielectric function of the Cs3Bi2I9 perovskite along the a- and b-axes (c), and along the c-axis) (d) as a function of pressure (from 0 to 40 GPa); Figure S5: Real part of the dielectric function of the Cs3Bi2Br9 perovskite along the a- and b-axes (a), and the c-axis (b) as a function of pressure (from 0 to 40 GPa). Imaginary part of the dielectric function of the Cs3Bi2Br9 perovskite along the a- and b-axes (c), and along the c-axis) (d) as a function of pressure (from 0 to 40 GPa); Figure S6: Real part of the dielectric function of the Cs3Bi2Cl9 perovskite along the a-axis (a), along the b-axis (b), and along the c-axis (c) as a function of pressure (from 0 to 40 GPa). Imaginary part of the dielectric function of the Cs3Bi2Cl9 perovskite along the a-axis (d), along the b-axis (e), and along the c-axis (f) as a function of pressure (from 0 to 40 GPa).

Author Contributions

Conceptualization, G.X., Y.W. and H.M.; theoretical calculation, G.X., Y.W., M.Z. and C.C.; writing—original draft preparation, G.X., M.Z. and H.M.; writing—review and editing, H.M. and J.L.; supervision, funding acquisition, J.L., C.C. and H.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Natural Science Foundation of Shandong Province (Nos. ZR2020MA081, ZR2019MA037, and ZR2018BA031). This work also received the financial support from by the National Natural Science Foundation of China (NFSC) (No. 11904212) and the Research Leader Program of Jinan Science and Technology Bureau (No. 2019GXRC061).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, X.; Xu, W.; Bai, S.; Jin, Y.; Wang, J.; Friend, R.H.; Gao, F. Metal halide perovskites for light-emitting diodes. Nat. Mater. 2021, 20, 10–21. [Google Scholar] [CrossRef]
  2. Jia, Y.; Kerner, R.A.; Grede, A.J.; Rand, B.P.; Giebink, N.C. Continuous-wave lasing in an organic-inorganic lead halide semiconductor. Nat. Photonics 2017, 11, 784–788. [Google Scholar] [CrossRef]
  3. Lin, K.; Xing, J.; Quan, L.; Arquer, F.P.G.; Gong, X.; Lu, J.; Xie, L.; Zhao, W.; Zhang, D.; Yan, C.; et al. Perovskite light-emitting diodes with external quantum efficiency exceeding 20 per cent. Nature 2018, 562, 245–248. [Google Scholar] [CrossRef]
  4. Chen, B.; Baek, S.W.; Hou, Y.; Aydin, E.; Bastiani, M.D.; Scheffel, B.; Proppe, A.; Huang, Z.; Wei, M.; Wang, Y.; et al. Enhanced optical path and electron diffusion length enable high-efficiency perovskite tandems. Nat. Commun. 2020, 11, 1–9. [Google Scholar] [CrossRef]
  5. Leng, K.; Abdelwahab, I.; Verzhbitskiy, I.; Telychko, M.; Chu, L.; Fu, W.; Chi, X.; Guo, N.; Chen, Z.; Chen, Z.; et al. Molecularly thin two-dimensional hybrid perovskites with tunable optoelectronic properties due to reversible surface relaxation. Nat. Mater. 2018, 17, 908–914. [Google Scholar] [CrossRef]
  6. Shao, H.; Bai, X.; Cui, H.; Pan, G.; Jing, P.; Qu, S.; Zhu, J.; Zhai, Y.; Dong, B.; Song, H. White light emission in Bi3+/Mn2+ ion co-doped CsPbCl3 perovskite nanocrystals. Nanoscale 2018, 10, 1023–1029. [Google Scholar] [CrossRef]
  7. Yang, S.; Chen, S.; Mosconi, E.; Fang, Y.; Xiao, X.; Wang, C.; Zhou, Y.; Yu, Z.; Zhao, J.; Gao, Y.; et al. Stabilizing halide perovskite surfaces for solar cell operation with wide-bandgap lead oxysalts. Science 2019, 365, 473–478. [Google Scholar] [CrossRef] [Green Version]
  8. Zhou, L.; Guo, X.; Lin, Z.; Ma, J.; Su, J.; Hu, Z.; Zhang, C.; Liu, S.; Chang, J.; Hao, Y. Interface engineering of low temperature processed all-inorganic CsPbI2Br perovskite solar cells toward PCE exceeding 14%. Nano Energy 2019, 60, 583–590. [Google Scholar] [CrossRef]
  9. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  10. Tang, S.; Bing, J.; Zheng, J.; Tang, J.; Li, Y.; Mayyas, M.; Cho, Y.; Jones, T.W.; Yang, T.C.J.; Yuan, L.; et al. Complementary bulk and surface passivations for highly efficient perovskite solar cells by gas quenching. Cell Rep. Phys. Sci. 2021, 2, 100511. [Google Scholar] [CrossRef]
  11. Yang, B.; Chen, J.; Hong, F.; Mao, X.; Zheng, K.; Yang, S.; Li, Y.; Pullerits, T.; Deng, W.; Han, K. Lead-Free, Air-Stable All-Inorganic Cesium Bismuth Halide Perovskite Nanocrystals. Angew. Chem. Int. Ed. 2017, 56, 12471–12475. [Google Scholar] [CrossRef]
  12. Leng, M.; Yang, Y.; Zeng, K.; Chen, Z.; Tan, Z.; Li, S.; Li, J.; Xu, B.; Li, D.; Hautzinger, M.P.; et al. All-Inorganic Bismuth-Based Perovskite Quantum Dots with Bright Blue Photoluminescence and Excellent Stability. Adv. Funct. Mater. 2017, 28, 1704446. [Google Scholar] [CrossRef]
  13. Lou, Y.; Fang, M.; Chen, J.; Zhao, Y. Formation of highly luminescent cesium bismuth halide perovskite quantum dots tuned by anion exchange. Chem. Commun. 2018, 54, 3779–3782. [Google Scholar] [CrossRef]
  14. Cao, Y.; Zhang, Z.; Li, L.; Zhang, J.; Zhu, J. An Improved Strategy for High-Quality Cesium Bismuth Bromine Perovskite Quantum Dots with Remarkable Electrochemiluminescence Activities. Anal. Chem. 2019, 91, 8607–8614. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, Z.; Shi, Z.; Wang, L.; Zhang, F.; Wu, D.; Yang, D.; Chen, X.; Zhang, Y.; Shan, C.; Li, X. Water-induced fluorescence enhancement of lead-free cesium bismuth halide quantum dots by 130% for stable white light-emitting devices. Nanoscale 2020, 12, 3637–3645. [Google Scholar] [CrossRef]
  16. Ji, Z.; Liu, Y.; Li, W.; Zhao, C.; Mai, W. Reducing current fluctuation of Cs3Bi2Br9 perovskite photodetectors for diffuse reflection imaging with wide dynamic range. Sci. Bull. 2020, 65, 1371–1379. [Google Scholar] [CrossRef]
  17. Gu, J.; Yan, G.; Lian, Y.; Mu, Q.; Jin, H.; Zhang, Z.; Deng, Z.; Peng, Y. Bandgap engineering of a lead-free defect perovskite Cs3Bi2I9 through trivalent doping of Ru3+. RSC Adv. 2018, 8, 25802–25807. [Google Scholar] [CrossRef] [Green Version]
  18. Han, J.; Le, Q.V.; Kim, H.; Lee, Y.J.; Lee, D.E.; Im, I.H.; Lee, M.K.; Kim, S.J.; Kim, J.; Kwak, K.J.; et al. Lead-Free Dual-Phase Halide Perovskites for Preconditioned Conducting-Bridge Memory. Small 2020, 16, 2003225. [Google Scholar] [CrossRef]
  19. Liu, D.; Yu, B.; Liao, M.; Jin, Z.; Zhou, L.; Zhang, X.; Wang, F.; He, H.; Gatti, T.; He, Z. Self-powered and broadband lead-free inorganic perovskite photodetector with high stability. ACS Appl. Mater. Interfaces 2020, 12, 30530–30537. [Google Scholar] [CrossRef]
  20. Li, W.; Wang, X.; Liao, J.; Jiang, Y.; Kuang, D. Enhanced On–Off Ratio Photodetectors Based on Lead-Free Cs3Bi2I9 Single Crystal Thin Films. Adv. Funct. Mater. 2020, 30, 1909701. [Google Scholar] [CrossRef]
  21. Khadka, D.B.; Shirai, Y.; Yanagida, M.; Miyano, K. Tailoring the film morphology and interface band offset of caesium bismuth iodide-based Pb-free perovskite solar cells. J. Mater. Chem. C 2019, 7, 8335–8343. [Google Scholar] [CrossRef]
  22. Johansson, M.B.; Philippe, B.; Banerjee, A.; Phuyal, D.; Mukherjee, S.; Chakraborty, S.; Cameau, M.; Zhu, H.; Ahuja, R.; Boschloo, G.; et al. Cesium Bismuth Iodide Solar Cells from Systematic Molar Ratio Variation of CsI and BiI3. Inorg. Chem. 2019, 58, 12040–12052. [Google Scholar] [CrossRef]
  23. Ge, S.; Guan, X.; Wang, Y.; Lin, C.; Cui, Y.; Huang, Y.; Zhang, X.; Zhang, R.; Yang, X.; Wu, T. Low-Dimensional Lead-Free Inorganic Perovskites for Resistive Switching with Ultralow Bias. Adv. Funct. Mater. 2020, 30, 2002110. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Liu, Y.; Xu, Z.; Ye, H.; Yang, Z.; You, J.; Liu, M.; He, Y.; Kanatzidis, M.G.; Liu, S. Nucleation-controlled growth of superior lead-free perovskite Cs3Bi2I9 single-crystals for high-performance X-ray detection. Nat. Commun. 2020, 11, 1–11. [Google Scholar]
  25. Hu, Y.; Zhang, S.; Miao, X.; Su, L.; Bai, F.; Qiu, T.; Liu, J.; Yuan, G. Ultrathin Cs3Bi2I9 Nanosheets as an Electronic Memory Material for Flexible Memristors. Adv. Mater. Interfaces 2017, 4, 1700131. [Google Scholar] [CrossRef]
  26. Liang, J.; Fang, Q.; Wang, H.; Xu, R.; Jia, S.; Guan, Y.; Ai, Q.; Gao, G.; Guo, H.; Shen, K.; et al. Perovskite-Derivative Valleytronics. Adv. Mater. 2020, 32, 2004111. [Google Scholar] [CrossRef]
  27. Samanta, D.; Saha, P.; Ghosh, B.; Chaudhary, S.P.; Bhattacharyya, S.; Chatterjee, S.; Mukherjee, G.D. Pressure-Induced Emergence of Visible Luminescence in Lead Free Halide Perovskite Cs3Bi2Br9: Effect of Structural Distortion. J. Phys. Chem. C 2021, 125, 3432–3440. [Google Scholar] [CrossRef]
  28. Nagaoka, Y.; Hills-Kimball, K.; Tan, R.; Li, R.; Wang, Z.; Chen, O. Nanocube Superlattices of Cesium Lead Bromide Perovskites and Pressure-Induced Phase Transformations at Atomic and Mesoscale Levels. Adv. Mater. 2017, 29, 1606666. [Google Scholar] [CrossRef]
  29. Lü, X.; Wang, Y.; Stoumpos, C.C.; Hu, Q.; Guo, X.; Chen, H.; Yang, L.; Smith, J.S.; Yang, W.; Zhao, Y.; et al. Enhanced Structural Stability and Photo Responsiveness of CH3NH3SnI3 Perovskite via Pressure-Induced Amorphization and Recrystallization. Adv. Mater. 2016, 28, 8663–8668. [Google Scholar] [CrossRef] [Green Version]
  30. Postorino, P.; Malavasi, L. Pressure-Induced Effects in Organic−Inorganic Hybrid Perovskites. J. Phys. Chem. Lett. 2017, 8, 2613–2622. [Google Scholar] [CrossRef]
  31. Szafranski, M.; Katrusiak, A. Mechanism of Pressure-Induced Phase Transitions, Amorphization, and Absorption-Edge Shift in Photovoltaic Methylammonium Lead Iodide. J. Phys. Chem. Lett. 2016, 7, 3458–3466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zhang, L.; Zeng, Q.; Wang, K. Pressure-Induced Structural and Optical Properties of Inorganic Halide Perovskite CsPbBr3. J. Phys. Chem. Lett. 2017, 8, 3752–3758. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.; Lü, X.; Yang, W.; Wen, T.; Yang, L.; Ren, X.; Wang, L.; Lin, Z.; Zhao, Y. Pressure-Induced Phase Transformation, Reversible Amorphization, and Anomalous Visible Light Response in Organolead Bromide Perovskite. J. Am. Chem. Soc. 2015, 137, 11144–11149. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, L.; Liu, C.; Wang, L.; Liu, C.; Wang, K.; Zou, B. Pressure-Induced Emission Enhancement, Band Gap Narrowing and Metallization of Halide Perovskite Cs3Bi2I9. Angew. Chem. Int. Ed. 2018, 57, 11213–11217. [Google Scholar] [CrossRef]
  35. Geng, T.; Wei, S.; Zhao, W.; Ma, Z.; Fu, R.; Xiao, G.; Zou, B. Insight into the structure–property relationship of two-dimensional lead-free halide perovskite Cs3Bi2Br9 nanocrystals under pressure. Inorg. Chem. Front. 2021, 8, 1410–1415. [Google Scholar] [CrossRef]
  36. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef] [PubMed]
  37. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  40. Islam, J.; Hossain, A. Semiconducting to metallic transition with outstanding optoelectronic properties of CsSnCl3 perovskite under pressure. Sci. Rep. 2020, 10, 1–11. [Google Scholar] [CrossRef]
  41. Morgan, E.E.; Mao, L.; Teicher, S.M.L.; Wu, G.; Seshadri, R. Tunable Perovskite-Derived Bismuth Halides: Cs3Bi2(Cl1−xIx)9. Inorg. Chem. 2020, 59, 3387–3393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Khoram, P.; Brittman, S.; Dzik, W.I.; Reek, J.N.H.; Garnett, E.C. Growth and Characterization of PDMS-Stamped Halide Perovskite Single Microcrystals. J. Phys. Chem. C 2016, 120, 6475–6481. [Google Scholar] [CrossRef]
  43. Creutz, S.E.; Liu, H.; Kaiser, M.E.; Li, X.; Gamelin, D.R. Structural Diversity in Cesium Bismuth Halide Nanocrystals. Chem. Mater. 2019, 31, 4685–4697. [Google Scholar] [CrossRef]
  44. Ghosh, B.; Chakraborty, S.; Wei, H.; Guet, C.; Li, S.; Mhaisalkar, S.; Mathews, N. Poor Photovoltaic Performance of Cs3Bi2I9: An Insight through First Principles Calculations. J. Phys. Chem. C 2017, 121, 17062–17067. [Google Scholar] [CrossRef]
  45. Bass, K.K.; Estergreen, L.; Savory, C.N.; Buckeridge, J.; Scanlon, D.O.; Djurovich, P.I.; Bradforth, S.E.; Thompson, M.E.; Melot, B.C. Vibronic Structure in Room Temperature Photoluminescence of the Halide Perovskite Cs3Bi2Br9. Inorg. Chem. 2017, 56, 42–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Xiang, G.; Wu, Y.; Li, Y.; Cheng, C.; Leng, J.; Ma, H. Structural and Optoelectronic Properties of Two-Dimensional Ruddlesden–Popper Hybrid Perovskite CsSnBr3. Nanomaterials 2021, 11, 2119. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Crystal structures of the zero-dimensional perovskite Cs3Bi2I9 (a), two-dimensional perovskite Cs3Bi2Br9 (b), and one-dimensional perovskite Cs3Bi2Cl9 (c).
Figure 1. Crystal structures of the zero-dimensional perovskite Cs3Bi2I9 (a), two-dimensional perovskite Cs3Bi2Br9 (b), and one-dimensional perovskite Cs3Bi2Cl9 (c).
Nanomaterials 11 02712 g001
Figure 2. Changes in the calculated lattice parameter as a function of pressure from 0 GPa to 40 GPa for the Cs3Bi2I9 perovskite (a), Cs3Bi2Br9 perovskite (b), and Cs3Bi2Cl9 perovskite (c). Changes in the unit cell volume of the Cs3Bi2X9 perovskites under the pressures from 0 GPa to 40 GPa (d).
Figure 2. Changes in the calculated lattice parameter as a function of pressure from 0 GPa to 40 GPa for the Cs3Bi2I9 perovskite (a), Cs3Bi2Br9 perovskite (b), and Cs3Bi2Cl9 perovskite (c). Changes in the unit cell volume of the Cs3Bi2X9 perovskites under the pressures from 0 GPa to 40 GPa (d).
Nanomaterials 11 02712 g002
Figure 3. Calculated band structure of the Cs3Bi2I9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Figure 3. Calculated band structure of the Cs3Bi2I9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Nanomaterials 11 02712 g003
Figure 4. Calculated band structure of the Cs3Bi2Br9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Figure 4. Calculated band structure of the Cs3Bi2Br9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Nanomaterials 11 02712 g004
Figure 5. Calculated band structure of the Cs3Bi2Cl9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Figure 5. Calculated band structure of the Cs3Bi2Cl9 perovskite under the pressures of 0 (a), 4 (b), 10 (c), and 40 GPa (d).
Nanomaterials 11 02712 g005
Figure 6. Calculated bandgap of the Cs3Bi2X9 perovskites as a function of pressure (from 0 GPa to 40 GPa) (a). Changes in the bandgap of the Cs3Bi2X9 perovskite after using the scissor values under HP (b). The black, red, and blue lines represent the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites, respectively.
Figure 6. Calculated bandgap of the Cs3Bi2X9 perovskites as a function of pressure (from 0 GPa to 40 GPa) (a). Changes in the bandgap of the Cs3Bi2X9 perovskite after using the scissor values under HP (b). The black, red, and blue lines represent the Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9 perovskites, respectively.
Nanomaterials 11 02712 g006
Figure 7. Crystal structure of the zero-dimensional Cs3Bi2I9 (a), one-dimensional Cs3Bi2Cl9 (b), and two-dimensional Cs3Bi2Br9 (c). Difference in the lattice constants of the Cs3Bi2X9 perovskites for the pressures between 0 and 40 GPa along the a, b, and c axes (d). The black, red, and blue lines represent Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9, respectively.
Figure 7. Crystal structure of the zero-dimensional Cs3Bi2I9 (a), one-dimensional Cs3Bi2Cl9 (b), and two-dimensional Cs3Bi2Br9 (c). Difference in the lattice constants of the Cs3Bi2X9 perovskites for the pressures between 0 and 40 GPa along the a, b, and c axes (d). The black, red, and blue lines represent Cs3Bi2I9, Cs3Bi2Br9, and Cs3Bi2Cl9, respectively.
Nanomaterials 11 02712 g007
Figure 8. Partial state of density (PDOS) of Cs3Bi2I9 (a), Cs3Bi2Br9 (b), and Cs3Bi2Cl9 (c). the VBM of the Cs3Bi2X9 perovskites (X = I, Br, and Cl) is dominated by the p-X states, and the CBM is dominated by the p-Bi and p-X states. Calculated density of states (DOS) of the Cs3Bi2I9 (d), Cs3Bi2Br9 (e), and Cs3Bi2Cl9 (f) under the pressures from 0 to 40 GPa.
Figure 8. Partial state of density (PDOS) of Cs3Bi2I9 (a), Cs3Bi2Br9 (b), and Cs3Bi2Cl9 (c). the VBM of the Cs3Bi2X9 perovskites (X = I, Br, and Cl) is dominated by the p-X states, and the CBM is dominated by the p-Bi and p-X states. Calculated density of states (DOS) of the Cs3Bi2I9 (d), Cs3Bi2Br9 (e), and Cs3Bi2Cl9 (f) under the pressures from 0 to 40 GPa.
Nanomaterials 11 02712 g008
Figure 9. Absorption coefficients of the Cs3Bi2I9 along the a- and b-axes (a), Cs3Bi2I9 along c-axis (b), Cs3Bi2Br9 along the a- and b-axes (c), and Cs3Bi2Br9 along the c-axis (d) as a function of the pressure (from 0 to 40 GPa).
Figure 9. Absorption coefficients of the Cs3Bi2I9 along the a- and b-axes (a), Cs3Bi2I9 along c-axis (b), Cs3Bi2Br9 along the a- and b-axes (c), and Cs3Bi2Br9 along the c-axis (d) as a function of the pressure (from 0 to 40 GPa).
Nanomaterials 11 02712 g009
Figure 10. Absorption coefficients of Cs3Bi2Cl9 along the a-axis (a), along the b-axis (b), and along the c-axis (c) as a function of the pressure (from 0 to 40 GPa).
Figure 10. Absorption coefficients of Cs3Bi2Cl9 along the a-axis (a), along the b-axis (b), and along the c-axis (c) as a function of the pressure (from 0 to 40 GPa).
Nanomaterials 11 02712 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiang, G.; Wu, Y.; Zhang, M.; Cheng, C.; Leng, J.; Ma, H. Dimension-Dependent Bandgap Narrowing and Metallization in Lead-Free Halide Perovskite Cs3Bi2X9 (X = I, Br, and Cl) under High Pressure. Nanomaterials 2021, 11, 2712. https://doi.org/10.3390/nano11102712

AMA Style

Xiang G, Wu Y, Zhang M, Cheng C, Leng J, Ma H. Dimension-Dependent Bandgap Narrowing and Metallization in Lead-Free Halide Perovskite Cs3Bi2X9 (X = I, Br, and Cl) under High Pressure. Nanomaterials. 2021; 11(10):2712. https://doi.org/10.3390/nano11102712

Chicago/Turabian Style

Xiang, Guangbiao, Yanwen Wu, Man Zhang, Chen Cheng, Jiancai Leng, and Hong Ma. 2021. "Dimension-Dependent Bandgap Narrowing and Metallization in Lead-Free Halide Perovskite Cs3Bi2X9 (X = I, Br, and Cl) under High Pressure" Nanomaterials 11, no. 10: 2712. https://doi.org/10.3390/nano11102712

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop