Next Article in Journal
Tailoring Al-Doped ZnO Nanoparticles via Scalable High-Energy Ball Milling–Solid-State Reaction: Structural, Optical, and Dielectric Insights for Light-Activated Antimicrobial Defense Against Medical Device Pathogens
Previous Article in Journal
Tannic Acid-Loaded Antibacterial Hydroxyapatite-Zirconia Composite for Dental Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unveiling the Impact of 0–20 Gpa Hydrostatic Pressure on the Physical Properties of (Cs2HfCl6) Double Perovskite

1
School of Physics and Electronic Engineering, Linyi University, Linyi 276000, China
2
Material Modeling Laboratory (MML), Islamia College Peshawar, Peshawer 25000, Pakistan
3
Department of Physics, College of Science, Qassim University, Buraydah 51452, Saudi Arabia
4
Department of Physics, College of Science, Northern Border University, Arar 73213, Saudi Arabia
5
Hubei Key Laboratory of Energy Storage and Power Battery, School of Optoelectronic Engineering, School of New Energy, Hubei University of Automotive Technology, Shiyan 442002, China
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(5), 395; https://doi.org/10.3390/cryst15050395
Submission received: 12 March 2025 / Revised: 13 April 2025 / Accepted: 16 April 2025 / Published: 24 April 2025

Abstract

:
The current work determines the physical properties of Cs2HfCl6 photovoltaic compounds including their structural, electronic, and optical behavior, utilizing the DFT approach. The simulated Cs2HfCl6 lattice constants, cell volumes, and bond lengths decrease as the pressure increases from 0 to 20 GPa. The band structure analysis reveals that the calculated under-pressure (0–20 GPa) of Cs2HfCl6 is semiconducting with a flexible indirect bandgap (5.44, 2.76, 2.02, 1.45, and 0.99) eV. The electronic bandgap diminishes (0–20 GPa), transitioning the compound from the UV to the visible spectra. This alteration improves the transition from the VBM to the CBM, hence augmenting the optical effectiveness. Concurrently, the dielectric function escalates, enhancing the absorption and conductivity, and causing a red shift in the optical spectra, while diminishing the reflection in the visible spectra. Our findings on the hydraulic pressure (0–20 GPa) and the electrical and optical properties indicate that Cs2HfCl6 may be utilized in the development of next-generation solar cells, LEDs, UV sensors, and high-pressure optical instruments.

1. Introduction

Organic–inorganic halide perovskites (OIHPs) are well-known for their outstanding optoelectronic properties and are used in solar cells, light emitting diodes (LEDs), X-ray detectors, and sensors [1,2,3]. Despite their success, OIHPs face significant hurdles in commercialization due to their lead (Pb) content and intrinsic instability. These challenges have encouraged scientists to develop stable lead-free all-inorganic perovskites [1,2,3,4,5,6]. The substituting of Pb2+ with group IVA elements like tin (Sn2+) or germanium (Ge2+) are ideal. However, these ions tend to oxidize in air, compromising their stability and their device performance [7]. To replace Pb, another approach focuses on double perovskites, where monovalent (e.g., Ag+) and trivalent (e.g., Bi3+, Sb3+, In3+) cations combine in structures like A2M+M3+X6. The best example is Cs2AgBiBr6; apart from this compound, many of these double perovskites exhibit a secondary phase, such as Cs2AgBil6 and Cs2AgSbI6, which form CsI, Cs3Bi2I9, and Cs3Sb2I9 structures, respectively [8].
A promising alternative lies in vacancy-ordered double perovskites (A2B4+X6), which replace Pb2+ with tetravalent metals (e.g., Zr4+, Hf4+, Pt4+) paired with ordered vacancies [9,10,11,12,13,14]. These materials maintain structural stability while offering tunable optoelectronic properties. For instance, compounds like Cs2ZrCl6, Cs2RuBr6, and Cs2PtI6 have demonstrated potential in light-emitting diode (LEDs) and photocatalytic applications due to their efficient charge-carrier dynamics and adaptable band structures [13,14]. By addressing both environmental concerns and stability issues, vacancy-ordered perovskites represent a critical pathway toward sustainable high-performance alternatives to conventional OIHPs. The research community has recently focused on the Cs2BX6 family. The cubic phase of Cs2SnI6, with a Sn4+ oxidation state, makes it a promising for optoelectronic applications [15,16,17,18]. Due to Sn’s tetravalent nature, Cs2SnI6 is more air-stable than CsSnI3 [16]. In addition, Cs2TiX6 and Cs2SnX6 PSCs have shown environmental promise [17]. Similarly, the wide-bandgap materials Cs2ZrCl6 and Cs2HfCl6 have shown promising results in luminescence [18,19,20]. These materials show potential for LEDs, photodetectors- and thermally induced delayed fluorescence due to their unique properties including high light yields, rapid decay kinetics, and tunable luminescent properties [21]. In recent years, Cs2HfX6 (X = Cl, Br, I) perovskite has been used as a scintillator in X-ray and gamma-radiation [22,23]. Jaroenjittichai et al. [24] theoretically investigated that Cs2HfCl6 is thermodynamically very stable and is not activated with H2O and O2. However, there has been little study on the structural stability and scintillation properties in an ambient environment. On the other hand, despite some experimental and theoretical study, the comprehensive properties for various applications have not been explored at high pressure. To understand the mechanism behind these materials, further study is required to deeply investigate these materials for optoelectronic applications [25].
Here, we used first-principles calculations based on density functional theory (DFT), which enables the study of material properties across sizes and structures, thus enabling the prediction and preemption of possible applications [26,27]. Lattice constant optimization with PBE and TB-mBJ accurately predicts the crystal bandgap, and hydraulic pressure affects the bandgap [28]. In this study, the structural, electrical, and optical properties of Cs2HfCl6 were investigated by varying the pressure (0–20) GPa. The aim of this work is to reduce the bandgap values with the applied hydraulic pressure, which enhances the optical performance of the compound making it suitable for various optoelectronics applications. We anticipate that our findings will serve as reference studies for preparing and applying the relevant materials in future research.

2. Results and Discussion

2.1. Structural Properties

The vacancy order (DP) Cs2HfCl6 has a cubic crystal structure with a space group of (225 Fm–3 m) as shown in Figure 1a,b [5]. The symmetrical structure comprises octahedra of Hf(Cl)6, with intervening regions occupied by Cs atoms. The octahedral structures in the material are defined by chlorine atoms, each of which exhibits a twelve-fold coordination. Figure 1 demonstrates that the Cs and Hf atoms are surrounded by 12 and 6 halogen ions, respectively. In a cubic structure, Cs, Hf, and Cl atoms are positioned at the 8c, 4a, and 24e Wyckoff sites with fractional coordinates (1/4, 1/4, 1/4), (0, 0, 0), and (0.2, 0, 0), respectively. The DP Cs2HfCl6 was optimized to ascertain the lattice constants. Table 1 shows the optimized lattice parameter using hydraulic pressure (0–20) GPa for Cs2HfCl6. The computed Cs2HfCl6 data were compared with the previous literature as shown in Table 1.
The present work utilizes the volume optimization approach to reduce the energy of the unit cell as a function of its volume, conforming to Murnaghan’s equation [29] for Cs2HfCl6, as illustrated in Figure 1. As the volume diminishes, the energy of the crystal correspondingly falls until it attains a minimum. Consequently, the minimal energy is referred to as the ground state energy, whereas the optimal volume denotes the volume of the system corresponding to the minimum energy. Raising the unit cell’s volume farther will elevate energy, resulting in system instability. The lattice constant, a o , the pressure derivative of the bulk modulus, B′, the bulk modulus, B o , and the ground state energy, E o , of these compounds are ascertained at the optimal volume and exhibited in Table 1. As the pressure of the computed material increases, it directly impacts the lattice parameter of the compounds. The lattice parameter would consistently decrease, as the hydraulic pressure increases from 0 to 20 GPa, as listed in Figure 2b. The structural stability was established by prior research, which confirms that these compounds are stable both experimentally and theoretically [24,30,31]. The optimized parameter of the current investigated Cs2HfCl6 is much closer to the previously calculated data displayed in Table 1.
Table 1. The optimized lattice parameter a o (Bohr), bulk, derivative of the bulk, volume, and ground state energy of Cs2HfCl6.
Table 1. The optimized lattice parameter a o (Bohr), bulk, derivative of the bulk, volume, and ground state energy of Cs2HfCl6.
Compound a o (bohar) B o B o ´ ( V o ) (a.u3) E o   ( R y ) Ref.
Cs2HfCl620.0933.144.452029.56−66,896.15Present work
10.7432.575.002091.68−66,896.16[5]
10.68----[31]
10.42----[24]

2.2. Electronic Properties

2.2.1. Electronic Band Structure (EBS)

Electronic band structures (EBSs) provide a fundamental framework for understanding the electronic distribution and electron behavior in materials [32]. By analyzing the EBS, the intrinsic properties of a material, such as its conductivity and optical response, can be investigated by varying the pressure from 0 GPa to 20 GPa [33]. The TB-mBJ potential was used to calculate the band structure, and the results are illustrated in Figure 3a,b. The band structures were plotted along the high-symmetry directions in the Brillouin zone (BZ), specifically W → L → Γ → X → W → K, within an energy range of −4 to +12 eV. At 0 GPa pressure, the valence band (VB) maximum and conduction band (CB) minimum of Cs2HfCl6 were located at the Γ point, indicating a direct bandgap semiconductor. However, as the pressure increased up to 20 GPa, the material underwent a transition from a direct to an indirect bandgap semiconductor. This shift is attributed to changes in the electron–ion potential and lattice constants under pressure, which lead to a reduction in the energy bandgap. The relationship between the hydrostatic pressure and bandgap was directly proportional, as shown in Figure 2a. The bandgap decreased consistently from 5.44 eV to 0.99 eV, as the pressure increased from 0 to 20 GPa, demonstrating a trend between the bandgap pressure as summarized in Table 1, which lists the computed bandgap values using the mBJ approach. The results confirm the direct bandgap nature of Cs2HfCl6 at ambient pressure and its cubic structural stability under varying pressures. The progressive decrease in the bandgap with increasing pressure, as depicted in Figure 3a,b, highlights the material sensitivity to external stress and its potential for tunable electronic properties. As the pressure increases, the nature of Cs2HfCl6 transforms from an insulator to a semiconductor. The present work is associated with the reported theoretical investigation and experimental work both having 5.67 eV [5] and 6.1 eV [34] bandgap values, which place these compounds in an insulating nature. Due to the larger bandgap value, the present work utilized hydraulic pressure to reduce the bandgap value to make Cs2HfCl6 effective for solar cells and other technological applications.

2.2.2. Density of States (DOS)

To study the electronic structure and energy distribution, the density of states (DOS) and partial density of states (PDOS) are essential tools. These metrics provide detailed insights into the contributions of different atomic orbitals (s, p, d, and f states) to the overall electronic states, enabling a deeper analysis of material electronic properties. The DOS and PDOS of Cs2HfCl6 were investigated within specific energy ranges: −6 to 6 eV at 0 GPa and −12 to 7 eV at 20 GPa, as shown in Figure 4a,b. Similarly, the contributions of individual atomic states (s, p, d, and f) to the DOS and PDOS are presented in Figure 5a,b. At 0 GPa, the Fermi level (EF), defined as 0 eV, aligns with the maximum of the VB. The electronic states near the EF were predominantly influenced by the Cs-p, Hf-d, and Cl-p states, which play a significant role in shaping the compound electronic characteristics. However, as the pressure increased to 20 GPa, significant changes occurred in the DOS and PDOS trends. The Hf-f states became more prominent in the VB, while the contributions of the p and d states diminished. This shift highlights the sensitivity of the compound electronic structure to external pressure, with the f states gaining importance under high pressure conditions. For this sequence of modeling, a range of pressures was utilized, and it was found that the pressure affected the states/eV relationship.

2.3. Electron Density

The bonding characteristics were determined using the charge density and its spatial distribution, as illustrated in Figure 5a,b. Unlike metallic and covalent bonds, which involve charge sharing between atoms, ionic bonds exhibit minimal charge overlap between their components [35]. In covalent binding, the charge density is localized, while the metallic bonds show a delocalized charge distribution. Figure 5a,b illustrates the spatial charge distribution of the vacancy-based DP along the 2D diagonal plane. The charge distributions of Cs atom do not interact with the Cl atom, which exhibits a completely spherical charge distribution. The formation of an ionic link between the Cl atom and the Cs atoms was demonstrated at (0 GPa). The nature of the compound would change due to a hydraulic pressure of (20 GPa). The charge distribution of Cs and Hf ranges from perfectly spherical to deformed, resulting in covalent interaction with the Cl atom in a dumbbell configuration. The investigated under-pressure (0–20 GPa) charge density is displayed in Figure 5a,b.

2.4. Optical Properties

The interaction of a material with electromagnetic radiation can be studied through its optical properties, which are largely influenced by two key factors: the rate of electron transitions and the rate of electron–hole (e-h) recombination. Electron transitions can be categorized into two types: intraband and interband shifts. Interband shifts, which involve electrons moving between different energy bands, are particularly significant in semiconductors and play a crucial role in determining their optical behavior [36].
The optical response of a material is often described using the dielectric function, denoted as ε ( ω ) = ε 1 ( ω ) + i ε 2 ( ω ) , where ω represents the frequency of the incident radiation [37]. This complex function consists of two components: the real part, ε1(ω), and the imaginary part, ε2(ω). The real part, ε1(ω), provides insights into how light propagates and polarizes within the material, reflecting its dispersive properties. On the other hand, the imaginary part, ε2(ω), is associated with the material’s ability to absorb electromagnetic radiation, highlighting its absorptive characteristics. Together, these components are critical for understanding the optical behavior of semiconductor materials, which are essential for applications in photonics, optoelectronics, and energy conversion technologies.
The investigated principal peaks of ε 1 ( ω ) are 4.91, 8.16, 8.77, 9.63, and 10.82, which correspond to the photon energy range at 5.7, 6.8, 6.8, 2.4, and 22 eV for (0, 5, 10, 15, and 20 GPa) pressure, as illustrated in Figure 6a. As the pressure increases, the curves exhibit an upward trend, transitioning from the UV to the visible spectrum or inferior energy. The decreasing tendency towards negative values of ε 1 ω elucidates the reflection of light’s incidence on the compound surface, resulting in the metallic behavior of the material. The spectrum range of the imaginary component, ε 2 ω , exhibits an increasing trend in which the peaks values are 4.29, 10.40, 11.02, 11.37, and 11.83 at 6.0, 10.97, 10.92, 10.70, and 10.62 eV for (0, 5, 10, 15, and 20 GPa) pressures. Figure 6b clearly illustrates that the curves demonstrate a little shift in confinement inside the visible region due to the pressure, transitioning from higher energies to lower energies. The ε 2 ω exhibits decreasing trends at a higher frequency region for various pressures (5–20 GPa), as shown in Figure 6b, respectively. Figure 6a,b enumerates the sharp and main peak’s for both ε 1 ( ω ) and ε 2 ω accompanied by the corresponding pressure (0–20 GPa).
The refractive index i ( ω ) = n ( ω ) + i k ( ω ) consists of two components: the real, n ( ω ) , and imaginary or extinction coefficient, k ( ω ) . When EMR propagates across a substance, the n ω provides insights into refraction, whereas the k ( ω ) elucidates the absorption aspect. There is a direct proportional relationship between the density and the refractive index [38]. The equations delineate a correlation between the dielectric function and the n ω and k ω [39]. The k ω and n ω are mathematically expressed in Equations [40]. The spectral range of n(ω) includes the prominent peaks seen at (5.7, 6.89, 6.76, 6.51, and 2.05) eV, which are 2.26, 3.07, 3.14, 3.31, and 3.45 with the hydraulic pressure (0–20 GPa), respectively. Figure 6c illustrates the alteration in the curves with increasing pressure, transitioning from the UV to the visible spectrum. The significant peaks of n(ω) demonstrate fluctuations due to differing energy level transition velocities at various pressures (0–20 GPa). Conversely, k(ω) attains its minimum at 6.21 eV, with 1.31 for 0 GPa, while the refractive index trends were completely changed due to higher pressure (5–20 GPa) in which the peaks values were 2.87, 2.95, 2.99, and 3.02 at 13.56 eV, as illustrated in Figure 6d.
Optical conductivity σ(ω) is crucial for the electrical transition strength between filled and empty levels during optical transitions. A comprehensive examination was performed on the σ(ω) of the cubic structure vacancy based DP under pressure (0–20 GPa). The σ(ω) is expressed in the following manner. Figure 6e graphically depicts numerous peaks seen at pressures (0–20 GPa). The peaks values of σ (ω) at (9.3, 9.6, 9.4, 9.4, and 9.4) eV were 5065.86, 10,942.9, 12,296.8, 13,321.3, and 13,583.7 with pressure (0–20 GPa). The σ(ω) peaks at various pressure (0–20 GPa) are dominant at higher frequency (UV) limits, as listed in Figure 6e. The permanent peaks of the computed compounds at various pressures (0–20 GPa) are highly recommended for UV photon sensor, high pressure optics, and tunable optical devices [41,42].
The absorption coefficient α(ω) defines the depth to which the EMR may contact any substance. The α(ω) can be determined, utilizing the relation (1) [43]:
α ω = 2 π ω c R e ω + ε 2 .
The α(ω) evaluates the capacity of a photon to infiltrate a compound at a particular wavelength prior to absorption. Furthermore, it offers essential insights into the material’s efficacy in maximum solar energy retention, a vital characteristic in solar cell applications. The under-pressure (0–20 GPa) computed compound shows prominent peaks in the UV domain as shown in Figure 6f. The graphic illustration shows visual absorption at all pressures, but the strength and range decrease with the pressure. This decrease is due to lower energy gaps. The additional UV absorption region, where the absorption effectiveness improves with pressure (0–20 GPa), makes Cs2HfCl6 useful for optoelectronic device uses and solar cell enhancements.
Reflectivity describes how a compound surface responds to incoming photon energy. Figure 6g shows Cs2HfCl6 under the impact of pressure (0–20 GPa). The static reflectivity of the under-pressure Cs2HfCl6 shows an increasing trend of 0.06, 0.15, 0.18, 0.21, and 0.22 for 0, 5, 10, 15, and 20 GPa, as illustrated in Figure 6g. Discrete high points were noticed in the UV spectrum at pressure (0–20 GPa) impacts. The highest fluctuating peaks were observed at 8 eV to 14 eV. As the pressure rose, several notable reflectance spectra peaks switched to smaller wavelengths. This implies that electronic shifts or vibrational patterns demand more energy. Pressure often changes the gap between energy levels in a substance. The blue shift was observed in the high frequency limits (UV) region, as shown in Figure 6g. The frequency ranges associated with the plasma resonance are determined by the energy loss function L(ω). Plasma resonance occurs when the frequency of the incident photons aligns with that of the plasmon [44]. The L(ω) of the compound dictates the energy dissipated as an electron traverses it. The L(ω) function exhibits its most significant peaks within the near-UV spectrum at 0 GPa. Plasmon peaks are characterized by distinct maxima, denoted as ωp. As compared to the 5–20 GPa pressure, the 0 GPa pressure of Cs2HfCl6 shows higher losses, with a value of 1.25 at 6.5 eV, as illustrated in Figure 6h. As the pressure escalates from 0 to 20 GPa, these peaks consistently shift to elevated energies, as illustrated in Figure 6h. A link exists between the minimum value of the energy loss function and the maximum value of ε 2 ω . This is an essential characteristic of semiconductors. The computed optical results of Cs2HfCl6 at various pressures (0–20 GPa) were compared with the vacancy-based double perovskite literature; the current work is effective for optoelectronic applications [45].

3. Experimental Section

Computational Method

The current work employed the FP-LAPW (full-potential linearized augmented plane wave) method in the WIEN2k calculation package for all analyses [46]. This method uses DFT [47], which reflects a system’s electron density as a function of its energy. Only atomic constants are used to solve the Schrödinger equation. In structural, electrical, and optical property computations, the local density approximation (LDA) [48] estimates the electron exchange energy and correlation. The FP-LAPW approach divides the crystal unit cell into muffin-tin (MT) spheres focused on atomic sites and interstitial regions. In the interstitial area, wave functions are constructed using plane waves and a cutoff parameter of RMT × Kmax = 7 for the precise energy convergence of eigenvalues. K max is the greatest k vector used for plane-wave expansion, while RMT is the muffin-tin sphere’s minimum radius. The k-point mesh in the first irreducible Brillouin zone (BZ) was optimized to 1000 points using GGA-PBEsol. The computations converged when the system energy reached 0.0001 Ry.

4. Conclusions

The present research employed the FP-LAPW approach using Wien2k code with the PBEsol-GGA and mBJ approximations to investigate the structure, electronic, and optical properties of the Cs2HfCl6. A wide range of pressure (0–20 GPa) was applied to investigate the vacancy-based double perovskite Cs2HfCl6. The escalating pressure (0–20 GPa) reduced the lattice constant, which altered the electronic band structure. At 0 GPa, Cs2HfCl6 exhibited a direct bandgap; however, as the pressure varied from 0 to 20 GPa, the compound transitioned from a direct to an indirect bandgap (L-X). The compound had a decreased bandgap as the pressure increased due to narrowing of the conduction band minimums (CBMs) and valence band maximums (VBMs). Additionally, the electron density and density of states (DOS) were investigated at pressures ranging from 0 to 20 GPa. When the pressure was 0 GPa, the electron density distribution showed mixed covalent bonds and ionic bonds; furthermore, with increasing pressure, the covalent bonding became stronger. The optical properties of Cs2HfCl6 were calculated at pressures ranging from 0 to 20 GPa and photon energies from 0 to 14 eV. The enhanced absorption and conductivity under increased pressure indicate that the compounds could be employed in a strained condition for high-frequency applications in optoelectronic devices.

Author Contributions

Conceptualization, U.F. and N.I.; methodology, U.F. and N.I.; software, U.F., N.I. and W.u.R.; validation, U.F. and N.I.; formal analysis, W.u.R.; investigation, U.F. and N.I.; resources, U.F.; data curation, U.F. and N.I.; writing—original draft preparation, U.F. and W.u.R.; writing—review and editing, W.u.R.; visualization, W.u.R. and Y.-L.W.; supervision, W.u.R., B.H. and A.A.; project administration, M.K.; funding acquisition, W.u.R. and Y.-L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was jointly supported by the Natural Science Foundation of Shandong Province of China (Grant No. ZR2024MA038), the National Nature Science Foundation of China (Grant No. 12475019), and the Deanship of Scientific Research at Northern Border University, Arar, KSA, through the project number NBU-FFR-2025-2193-11.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tang, X.; Hu, Z.; Yuan, W.; Hu, W.; Shao, H.; Han, D.; Zheng, J.; Hao, J.; Zang, Z.; Du, J.; et al. Perovskite CsPb2Br5 Microplate Laser with Enhanced Stability and Tunable Properties. Adv. Opt. Mater. 2017, 5, 1600788. [Google Scholar] [CrossRef]
  2. Li, N.; Song, L.; Jia, Y.; Dong, Y.; Xie, F.; Wang, L.; Tao, S.; Zhao, N. Stabilizing Perovskite Light-Emitting Diodes by Incorporation of Binary Alkali Cations. Adv. Mater. 2020, 32, 1907786. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, T.; Yang, Y.; Gong, X. Recent Advancements and Challenges for Low-Toxicity Perovskite Materials. ACS Appl. Mater. Interfaces 2020, 12, 26776–26811. [Google Scholar] [CrossRef]
  4. Chen, B.; Yu, Z.J.; Manzoor, S.; Wang, S.; Weigand, W.; Yu, Z.; Yang, G.; Ni, Z.; Dai, X.; Holman, Z.C.; et al. Blade-Coated Perovskites on Textured Silicon for 26%-Efficient Monolithic Perovskite/Silicon Tandem Solar Cells. Joule 2020, 4, 850–864. [Google Scholar] [CrossRef]
  5. Noman, M.; Neffati, R.; Khan, S.; Murad, K.; Ashraf, M.W.; Murtaza, G. The Halide Ion Replacement Effects on the Physical Properties of Cs2BX6 Variant Perovskites. Phys. B Condens. Matter 2023, 656, 414779. [Google Scholar] [CrossRef]
  6. Cheel, H.J.; Capper, P. (Eds.) Crystal Growth Technology: From Fundamentals and Simulation to Large-Scale Production, 1st ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008; ISBN 978-3-527-31762-2. [Google Scholar] [CrossRef]
  7. Lin, Z.; Gilbert, B.; Liu, Q.; Ren, G.; Huang, F. A Thermodynamically Stable Nanophase Material. J. Am. Chem. Soc. 2006, 128, 6126–6131. [Google Scholar] [CrossRef]
  8. Correa-Baena, J.-P.; Nienhaus, L.; Kurchin, R.C.; Shin, S.S.; Wieghold, S.; Putri Hartono, N.T.; Layurova, M.; Klein, N.D.; Poindexter, J.R.; Polizzotti, A.; et al. A-Site Cation in Inorganic A3Sb2I9 Perovskite Influences Structural Dimensionality, Exciton Binding Energy, and Solar Cell Performance. Chem. Mater. 2018, 30, 3734–3742. [Google Scholar] [CrossRef]
  9. Nishikubo, R.; Kanda, H.; García-Benito, I.; Molina-Ontoria, A.; Pozzi, G.; Asiri, A.M.; Nazeeruddin, M.K.; Saeki, A. Optoelectronic and Energy Level Exploration of Bismuth and Antimony-Based Materials for Lead-Free Solar Cells. Chem. Mater. 2020, 32, 6416–6424. [Google Scholar] [CrossRef]
  10. Zhang, F.; Zhou, Y.; Chen, Z.; Wang, M.; Ma, Z.; Chen, X.; Jia, M.; Wu, D.; Xiao, J.; Li, X.; et al. Thermally Activated Delayed Fluorescence Zirconium-Based Perovskites for Large-Area and Ultraflexible X-Ray Scintillator Screens. Adv. Mater. 2022, 34, 2204801. [Google Scholar] [CrossRef]
  11. Cao, M.; Li, Z.; Zhao, X.; Gong, X. Achieving Ultrahigh Efficiency Vacancy-Ordered Double Perovskite Microcrystals via Ionic Liquids. Small 2022, 18, 2204198. [Google Scholar] [CrossRef]
  12. Vishnoi, P.; Zuo, J.L.; Cooley, J.A.; Kautzsch, L.; Gómez-Torres, A.; Murillo, J.; Fortier, S.; Wilson, S.D.; Seshadri, R.; Cheetham, A.K. Chemical Control of Spin-Orbit Coupling and Charge Transfer in Vacancy-Ordered Ruthenium(IV) Halide Perovskites. Angew. Chem. 2021, 133, 5244–5248. [Google Scholar] [CrossRef]
  13. Hamdan, M.; Chandiran, A.K. Cs2PtI6 Halide Perovskite Is Stable to Air, Moisture, and Extreme pH: Application to Photoelectrochemical Solar Water Oxidation. Angew. Chem. 2020, 132, 16167–16172. [Google Scholar] [CrossRef]
  14. Sun, J.; Zheng, W.; Huang, P.; Zhang, M.; Zhang, W.; Deng, Z.; Yu, S.; Jin, M.; Chen, X. Efficient Near-Infrared Luminescence in Lanthanide-Doped Vacancy-Ordered Double Perovskite Cs2ZrCl6 Phosphors via Te4+ Sensitization. Angew. Chem. 2022, 134, e202201993. [Google Scholar] [CrossRef]
  15. Saparov, B.; Sun, J.-P.; Meng, W.; Xiao, Z.; Duan, H.-S.; Gunawan, O.; Shin, D.; Hill, I.G.; Yan, Y.; Mitzi, D.B. Thin-Film Deposition and Characterization of a Sn-Deficient Perovskite Derivative Cs2 SnI6. Chem. Mater. 2016, 28, 2315–2322. [Google Scholar] [CrossRef]
  16. Xiao, Z.; Zhou, Y.; Hosono, H.; Kamiya, T. Intrinsic Defects in a Photovoltaic Perovskite Variant Cs2 SnI6. Phys. Chem. Chem. Phys. 2015, 17, 18900–18903. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, S.; Zhu, S.; Guan, J.; Wang, R.; Zheng, W.; Gao, P.; Lu, X. Confronting the Air Instability of Cesium Tin Halide Perovskites by Metal Ion Incorporation. J. Phys. Chem. Lett. 2021, 12, 10996–11004. [Google Scholar] [CrossRef]
  18. Chen, M.; Ju, M.-G.; Carl, A.D.; Zong, Y.; Grimm, R.L.; Gu, J.; Zeng, X.C.; Zhou, Y.; Padture, N.P. Cesium Titanium(IV) Bromide Thin Films Based Stable Lead-Free Perovskite Solar Cells. Joule 2018, 2, 558–570. [Google Scholar] [CrossRef]
  19. Qiu, X.; Cao, B.; Yuan, S.; Chen, X.; Qiu, Z.; Jiang, Y.; Ye, Q.; Wang, H.; Zeng, H.; Liu, J.; et al. From Unstable CsSnI3 to Air-Stable Cs2SnI6: A Lead-Free Perovskite Solar Cell Light Absorber with Bandgap of 1.48 eV and High Absorption Coefficient. Sol. Energy Mater. Sol. Cells 2017, 159, 227–234. [Google Scholar] [CrossRef]
  20. Saeki, K.; Fujimoto, Y.; Koshimizu, M.; Yanagida, T.; Asai, K. Comparative Study of Scintillation Properties of Cs2HfCl6 and Cs2ZrCl6. Appl. Phys. Express 2016, 9, 042602. [Google Scholar] [CrossRef]
  21. Kang, B.; Biswas, K. Carrier Self-Trapping and Luminescence in Intrinsically Activated Scintillator: Cesium Hafnium Chloride (Cs2HfCl6). J. Phys. Chem. C 2016, 120, 12187–12195. [Google Scholar] [CrossRef]
  22. Wang, X.; Zhang, H.; Jiang, W.; Zhuang, Q.; Feng, S.; Li, P.; Ou, T.; Ma, X. The First Principles Study on Double Halide Perovskites Cs2HfX6 (X = Cl, Br and I): Structural, Electronic, Optical Properties, and Strain Effects. J. Solid State Chem. 2025, 348, 125353. [Google Scholar] [CrossRef]
  23. Vaněček, V. Kombinatorický Vývoj Scintilátorů Na Bázi Komplexních Halogenidů. Ph.D. Thesis, Czech Technical University in Prague, Prague, Czech Republic, 2022. [Google Scholar]
  24. Kaewmeechai, C.; Laosiritaworn, Y.; Jaroenjittichai, A.P. First-Principles Study on Structural Stability and Reaction with H2O and O2 of Vacancy-Ordered Double Perovskite Halides: Cs2(Ti,Zr,Hf)X6. Results Phys. 2021, 25, 104225. [Google Scholar] [CrossRef]
  25. Liu, D.; Dang, P.; Zhang, G.; Lian, H.; Li, G.; Lin, J. Near-Infrared Emitting Metal Halide Materials: Luminescence Design and Applications. InfoMat 2024, 6, e12542. [Google Scholar] [CrossRef]
  26. Naji, S.; Belhaj, A.; Labrim, H.; Bhihi, M.; Benyoussef, A.; El Kenz, A. Electronic and Magnetic Properties of Iron Adsorption on Graphene with Double Hexagonal Geometry. Int. J. Quantum Chem. 2014, 114, 463–467. [Google Scholar] [CrossRef]
  27. El Hallani, F.; Naji, S.; Ez-Zahraouy, H.; Benyoussef, A. First-Principles Study of the Magnetic Stability and the Exchange Couplings of LaMn2O5. J. Appl. Phys. 2013, 114, 163909. [Google Scholar] [CrossRef]
  28. Errandonea, D.; Popescu, C.; Garg, A.B.; Botella, P.; Martinez-García, D.; Pellicer-Porres, J.; Rodríguez-Hernández, P.; Muñoz, A.; Cuenca-Gotor, V.; Sans, J.A. Pressure-Induced Phase Transition and Band-Gap Collapse in the Wide-Band-Gap Semiconductor InTaO4. Phys. Rev. B 2016, 93, 035204. [Google Scholar] [CrossRef]
  29. Tyuterev, V.G.; Vast, N. Murnaghan’s Equation of State for the Electronic Ground State Energy. Comput. Mater. Sci. 2006, 38, 350–353. [Google Scholar] [CrossRef]
  30. Jiang, J.; Zheng, J.; Fu, H.; Zhang, H.; Ou, D.; Chen, Q.; Wang, K.; Cao, S.; Zhao, J.; Du, Z.; et al. Scalable and Room-Temperature Preparation of Cs2HfCl6 Double Perovskites with Recorded Photoluminescence Efficiency and Robust Stability. Chem. Eng. J. 2024, 479, 147543. [Google Scholar] [CrossRef]
  31. Král, R.; Babin, V.; Mihóková, E.; Buryi, M.; Laguta, V.V.; Nitsch, K.; Nikl, M. Luminescence and Charge Trapping in Cs2HfCl6 Single Crystals: Optical and Magnetic Resonance Spectroscopy Study. J. Phys. Chem. C 2017, 121, 12375–12382. [Google Scholar] [CrossRef]
  32. Chakraborty, K.; Choudhury, M.G.; Paul, S. Numerical Study of Cs2TiX6 (X = Br, I, F and Cl) Based Perovskite Solar Cell Using SCAPS-1D Device Simulation. Sol. Energy 2019, 194, 886–892. [Google Scholar] [CrossRef]
  33. Charifi, Z.; Baaziz, H.; Hussain Reshak, A. Ab-Initio Investigation of Structural, Electronic and Optical Properties for Three Phases of ZnO Compound. Phys. Status Solidi (B) 2007, 244, 3154–3167. [Google Scholar] [CrossRef]
  34. Nagorny, S. Novel Cs2HfCl6 Crystal Scintillator: Recent Progress and Perspectives. Physics 2021, 3, 320–351. [Google Scholar] [CrossRef]
  35. Wuttig, M.; Deringer, V.L.; Gonze, X.; Bichara, C.; Raty, J. Incipient Metals: Functional Materials with a Unique Bonding Mechanism. Adv. Mater. 2018, 30, 1803777. [Google Scholar] [CrossRef] [PubMed]
  36. Shojaei-Oghani, M.; Yavari, M.H. Modeling the Effects of Interband and Intraband Transitions on Phase and Gain Stabilities of Quantum Dot Semiconductor Optical Amplifiers. Opt. Quant. Electron. 2018, 50, 374. [Google Scholar] [CrossRef]
  37. Cappellini, G.; Del Sole, R.; Reining, L.; Bechstedt, F. Model Dielectric Function for Semiconductors. Phys. Rev. B 1993, 47, 9892–9895. [Google Scholar] [CrossRef]
  38. Reyes-Alberto, M.; García-Valenzuela, A.; Gutierrez-Herrera, E. Method for Measuring the Extinction Coefficient of Fluorescing Media within the Emission Band. Appl. Opt. 2023, 62, C106. [Google Scholar] [CrossRef]
  39. Muhammad, F.F.; Yahya, M.Y.; Aziz, F.; Rasheed, M.A.; Sulaiman, K. Tuning the Extinction Coefficient, Refractive Index, Dielectric Constant and Optical Conductivity of Gaq3 Films for the Application of OLED Displays Technology. J. Mater. Sci. Mater. Electron. 2017, 28, 14777–14786. [Google Scholar] [CrossRef]
  40. Toptygin, D. Effects of the Solvent Refractive Index and Its Dispersion on the Radiative Decay Rate and Extinction Coefficient of a Fluorescent Solute. J. Fluoresc. 2003, 13, 201–219. [Google Scholar] [CrossRef]
  41. Nofriandi, A.; Hamdi; Ratnawulan; Yulkifli. Ultra-Sensitive Light Detection Technologies Based on Single-Photon Detectors: A Review. Sens. Technol. 2024, 2, 2404268. [Google Scholar] [CrossRef]
  42. Zhu, Y.; Wang, A. Miniature Fiber-Optic Pressure Sensor. IEEE Photon. Technol. Lett. 2005, 17, 447–449. [Google Scholar] [CrossRef]
  43. Hu, C.; Muller-Karger, F.E.; Zepp, R.G. Absorbance, Absorption Coefficient, and Apparent Quantum Yield: A Comment on Common Ambiguity in the Use of These Optical Concepts. Limnol. Oceanogr. 2002, 47, 1261–1267. [Google Scholar] [CrossRef]
  44. Lalanne, P.; Yan, W.; Vynck, K.; Sauvan, C.; Hugonin, J. Light Interaction with Photonic and Plasmonic Resonances. Laser Photonics Rev. 2018, 12, 1700113. [Google Scholar] [CrossRef]
  45. Zhao, X.-H.; Wang, F.; Hu, D.-Y.; Lu, L.-M.; Li, L.; Tang, T.-Y.; Tang, Y.-L. Effect of Hydrostatic Pressure on the Structural, Elastic, and Optoelectronic Properties of Vacancy-Ordered Double Perovskite Cs2PdBr6. J. Mol. Model. 2022, 28, 337. [Google Scholar] [CrossRef] [PubMed]
  46. Schwarz, K.; Blaha, P. Solid State Calculations Using WIEN2k. Comput. Mater. Sci. 2003, 28, 259–273. [Google Scholar] [CrossRef]
  47. Orio, M.; Pantazis, D.A.; Neese, F. Density Functional Theory. Photosynth. Res. 2009, 102, 443–453. [Google Scholar] [CrossRef]
  48. Gould, T.; Pittalis, S. Local Density Approximation for Excited States. Phys. Rev. X 2024, 14, 041045. [Google Scholar] [CrossRef]
Figure 1. (a,b) The under-pressure crystal structure (0–20 GPa) and optimized energy–volume plot of Cs2HfCl6.
Figure 1. (a,b) The under-pressure crystal structure (0–20 GPa) and optimized energy–volume plot of Cs2HfCl6.
Crystals 15 00395 g001
Figure 2. (a,b) The pressure verses bandgap and lattice parameter plots for Cs2HfCl6.
Figure 2. (a,b) The pressure verses bandgap and lattice parameter plots for Cs2HfCl6.
Crystals 15 00395 g002
Figure 3. (a,b) The pressure variance (0–20 GPa) electronic band structure of Cs2HfCl6.
Figure 3. (a,b) The pressure variance (0–20 GPa) electronic band structure of Cs2HfCl6.
Crystals 15 00395 g003
Figure 4. (a,b) The DOS at (0–20 GPa) for Cs2HfCl6.
Figure 4. (a,b) The DOS at (0–20 GPa) for Cs2HfCl6.
Crystals 15 00395 g004
Figure 5. (a,b) The under pressure (0–20 GPa) electron density plot of Cs2HfCl6.
Figure 5. (a,b) The under pressure (0–20 GPa) electron density plot of Cs2HfCl6.
Crystals 15 00395 g005
Figure 6. (ah) The real and imaginary dielectric function, refractive index, extinction coefficient, optical conductivity, absorption, reflectivity, and energy loss under-pressure (0–20 GPa) electron density plot of Cs2HfCl6.
Figure 6. (ah) The real and imaginary dielectric function, refractive index, extinction coefficient, optical conductivity, absorption, reflectivity, and energy loss under-pressure (0–20 GPa) electron density plot of Cs2HfCl6.
Crystals 15 00395 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Farooq, U.; Israr, N.; Hassan, B.; Alnakhlani, A.; Kallel, M.; Rehman, W.u.; Wang, Y.-L. Unveiling the Impact of 0–20 Gpa Hydrostatic Pressure on the Physical Properties of (Cs2HfCl6) Double Perovskite. Crystals 2025, 15, 395. https://doi.org/10.3390/cryst15050395

AMA Style

Farooq U, Israr N, Hassan B, Alnakhlani A, Kallel M, Rehman Wu, Wang Y-L. Unveiling the Impact of 0–20 Gpa Hydrostatic Pressure on the Physical Properties of (Cs2HfCl6) Double Perovskite. Crystals. 2025; 15(5):395. https://doi.org/10.3390/cryst15050395

Chicago/Turabian Style

Farooq, Umar, Nabeel Israr, Belqees Hassan, Ali Alnakhlani, Mohamed Kallel, Wasif ur Rehman, and Yong-Long Wang. 2025. "Unveiling the Impact of 0–20 Gpa Hydrostatic Pressure on the Physical Properties of (Cs2HfCl6) Double Perovskite" Crystals 15, no. 5: 395. https://doi.org/10.3390/cryst15050395

APA Style

Farooq, U., Israr, N., Hassan, B., Alnakhlani, A., Kallel, M., Rehman, W. u., & Wang, Y.-L. (2025). Unveiling the Impact of 0–20 Gpa Hydrostatic Pressure on the Physical Properties of (Cs2HfCl6) Double Perovskite. Crystals, 15(5), 395. https://doi.org/10.3390/cryst15050395

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