Next Article in Journal
Fabrication and Characterization of Physical and Mechanical Properties of Carbon Nanotubes—Graphene-Based Sandwich Composite Pressure Sensor
Next Article in Special Issue
Structural Transitions and Stability of FAPbI3 and MAPbI3: The Role of Interstitial Water
Previous Article in Journal
Effect of Graphene and Graphene Oxide on Airway Barrier and Differential Phosphorylation of Proteins in Tight and Adherens Junction Pathways
Previous Article in Special Issue
In Situ Raman Microdroplet Spectroelectrochemical Investigation of CuSCN Electrodeposited on Different Substrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Structural Competition between the Black and the Yellow Phase of CsPbI3

by
Ioannis Deretzis
1,
Corrado Bongiorno
1,
Giovanni Mannino
1,
Emanuele Smecca
1,
Salvatore Sanzaro
1,
Salvatore Valastro
1,2,
Giuseppe Fisicaro
1,
Antonino La Magna
1,* and
Alessandra Alberti
1
1
Istituto per la Microelettronica e Microsistemi, Dipartimento di Scienze Fisiche e Tecnologie della Materia, Consiglio Nazionale delle Ricerche, Z.I. VIII Strada 5, 95121 Catania, Italy
2
Dipartimento di Fisica e Astronomia “Ettore Majorana”, Università di Catania, Via S. Sofia 64, 95123 Catania, Italy
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(5), 1282; https://doi.org/10.3390/nano11051282
Submission received: 14 April 2021 / Revised: 10 May 2021 / Accepted: 11 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue Mesoporous Materials and Nanoscale Phenomena in Hybrid Photovoltaics)

Abstract

:
The realization of stable inorganic perovskites is crucial to enable low-cost solution-processed photovoltaics. However, the main candidate material, CsPbI3, suffers from a spontaneous phase transition at room temperature towards a photo-inactive orthorhombic δ-phase (yellow phase). Here we used theoretical and experimental methods to study the structural and electronic features that determine the stability of the CsPbI3 perovskite. We argued that the two physical characteristics that favor the black perovskite phase at low temperatures are the strong spatial confinement in nanocrystalline structures and the level of electron doping in the material. Within this context, we discussed practical procedures for the realization of long-lasting inorganic lead halide perovskites.

Graphical Abstract

1. Introduction

Lead halide perovskites are materials that recently emerged as candidates for next generation solar cells, combining high yields with low-cost processing methods. The photovoltaic properties of hybrid halide perovskites were first demonstrated by Miyasaka et al. [1] who showed a photoactive response in methylammonium-based lead iodides and bromides. Since then, ongoing improvements in the composition [2], mixing [3,4], interface engineering [5,6], and architectural optimization [7,8] have resulted in record-high solar cell efficiencies, directly comparable or even superior to present-day silicon-based technology [9]. The realization of these breakthroughs was largely accomplished by the partial or total occupation of the A-site cation in the ABX3 perovskite structure by a monovalent organic molecule, like methylammonium (MA) or formamidinium (FA). MAPbI3 and FAPbI3 perovskite solar cells showed excellent photovoltaic performance [10,11] but suffered from fast degradation under humidity, electromagnetic radiation, or heat [12,13,14]. Therefore, since the very beginning, the presence of the organic component in the perovskite matrix raised questions on the long-term stability of the respective devices. This stability deficit was partially mitigated by compositional mixing (e.g., by forming mixed Cs-FA-MA-Pb-I-Br perovskites), where the more robust lead-bromide bonding [15] and the reduced rotational disorder [16] increased the lifetime of the respective perovskites. However, even in this case, phase segregation (i.e., the separation of the mixed perovskite into elementary non-perovskite phases) [17] remained a critical issue that has yet to be settled.
A different strategy for the long-term stability of lead halide perovskites is the substitution of the organic cation from the ABX3 matrix with a monovalent inorganic cation having a large ionic radius, like cesium. The CsPbI3 perovskite could ideally be used in both tandem or single-junction solar cells, having a direct bandgap of ~1.73 eV and favorable optoelectronic characteristics [18] (whereas the CsPbBr3 perovskite has a bandgap of ~2.35 eV [19], which significantly restricts its application range in optoelectronic devices). The perovskite phase of CsPbI3 is however not the most stable phase at room temperature, where a photo-inactive orthorhombic yellow phase prevails (δ-phase) [20]. The phase diagram of the material indicates a spontaneous transformation from this yellow phase to the black α-perovskite structure only above 300 °C, which is impracticable for solar cell applications. In this respect, the main challenge for the utilization of CsPbI3 as an active material in light absorbers is the stabilization of the perovskite phase at a temperature range useful for common optoelectronic devices (20–80 °C). This issue has been a subject of investigation in numerous literature studies, where various strategies have been applied to engineer a stable perovskite phase for CsPbI3 close to room temperature. Some of the most prominent approaches are regarding the introduction of bivalent cations like Eu2+ in the perovskite lattice [21], the stabilization through strain engineering [22], or the enhancement of the process control using HI (or HPbI3) [18,23], water [24], or other additives [25] during the perovskite synthesis. Although these appear to be viable routes for the preservation of the perovskite phase at device operating temperatures, understanding the physical properties that regulate the balance between the two competing phases in CsPbI3 is still unclear.
In this paper, we use theoretical (density functional theory) and experimental (X-ray diffraction, high-resolution scanning transmission electron microscopy, spectroscopic ellipsometry) techniques to investigate the impact of various physical characteristics in competition between the yellow and the perovskite phase of CsPbI3. We find that the two physical parameters that can shift the phase equilibrium towards the perovskite structure are the electron doping of the material and its spatial confinement in nanometer-sized crystals. We additionally discuss practical process recipes to achieve these objectives. The paper is organized as follows: in Section 2 we describe the theoretical and experimental methodology used in this study, in Section 3 we show calculation outcomes and experimental characterizations of our samples, in Section 4 we critically discuss our results, while in Section 5 we provide the conclusions.

2. Materials and Methods

CsPbI3 films were obtained though solution processing of 1 M of PbI2 and 1 M of CsI (Tokyo Chemical Industry Co. Ltd., Tokyo, Japan) in a composite solvent made of DMF and DMSO (3:1 v/v). An EuCl3 solution of 0.1 M (Sigma-Aldrich, Darmstadt, Germany) was separately added in a mixed solvent of DMF and DMSO with the same composition as before. After separately stirring the solutions at room temperature for 1 h, 1 mL of PbI2/CsI was mixed with 0.5 mL of EuCl3 and stirred for a further 1 h. Subsequently, the mixture was deposited on glass substrates and spin-coated in two steps: 1000 rpm for 10 s followed by 5000 rpm for 25 s in ambient environment. The CsPbI3 film with Eu/Pb = 5% was then annealed at 80 °C on a hot plate for 20 s and cooled to 30 °C at 0.5–1 °C/s to form the black perovskite phase. The final CsPbI3 layer had a thickness of ~80 nm. A complete discussion on the preparation procedure can be found in Ref. [26].
XRD patterns were collected with a D8Discoved diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) equipped with a high precision goniometer (0.0001 Å) and a Cu-Ka source with an instrumental broadening of 0.07°. In situ XRD was performed using an DHS 900 sample holder (Anton Paar GmbH, Graz, Austria), equipped with a PolyEther-Ether-Ketone dome filled with nitrogen at a pressure slightly higher than the atmospheric pressure (+0.3 bar). Spectroscopic ellipsometry was performed with a VASE Ellipsometer (J. A. Woollam Co. Inc., Lincoln, NE, USA). Samples were kept in a closed chamber with an overpressure of N2. High-resolution imaging was performed by Scanning Transmission Electron Microscopy (STEM) at an accelerating voltage of 200 kV, using a Cs-probe corrected ARM200 (JEOL Ltd, Akishima, Tokyo, Japan), equipped with a cold field emission gun.
Density functional theory calculations were performed with the plane-wave Quantum Espresso code [27]. We used the Perdew–Burke–Ernzerhof (PBE) implementation [28] of the generalized gradient approximation for the description of the exchange-correlation functional and scalar relativistic ultrasoft pseudopotentials [29] for the description of the ionic cores. Pb-5d semicore electrons were dynamically treated in the same way as valence electrons. Convergence was achieved with a plane-wave cutoff kinetic energy of 49 Ry and an augmented charge density cutoff of 258 Ry. Calculations were performed on CsPbI3 unit cell structures comprised of 20 atoms in either the perovskite or the yellow orthorhombic symmetry. A (6 × 6 × 4) Monkhorst-Pack grid [30] was considered for Brillouin zone sampling of the perovskite phase, whereas a (8 × 5 × 2) grid was used for the yellow phase. Both atomic positions and lattice parameters were allowed to fully relax.

3. Results

3.1. Intrinsic Bulk Properties of the Yellow and Black Phase in CsPbI3

CsPbI3 can be crystallized in four polymorphic phases that are distinguished by different structures and symmetries. The α, β, and γ phases crystallize in a perovskite structure with either cubic (α), tetragonal (β), or orthorhombic (γ) symmetry. Out of the three, the γ phase has the smallest ground-state energy and it is therefore the most favorable perovskite phase for bulk crystals at relatively low temperatures. At high temperatures instead (>300 °C), the cubic α-perovskite spontaneously forms due to a combined beneficial effect of lattice vibrations/disorder and thermal expansion that lowers the free energy of this perovskite system with respect to all other phases. The orthorhombic (δ) structure instead is a non-perovskite crystal with a formation energy that is even lower than the γ perovskite phase [31]. Figure 1a,b shows the structural characteristics of the γ and the δ phases, respectively, which, under certain conditions, can be both experimentally observed at room temperature. The γ-perovskite is characterized by the typical [PbI6]4− octahedra in perovskite crystals, forming Pb-I cages that accommodate the Cs+ monovalent cations. The δ phase is instead characterized by quasi-one-dimensional [PbI3] chains that are surrounded by Cs+ cations, making the structure highly anisotropic towards the three crystallographic directions. The band structures of the two phases by means of the DFT (Figure 1c,d) provide evidence of why the former is suitable for optoelectronic devices whereas the second lacks photoactivity. For the γ phase, a direct band gap with a value of ~1.75 eV is calculated (Figure 1c), which makes this system ideal for tandem solar cells and potentially useful in other optoelectronic devices. The δ phase instead shows an indirect wide band gap of ~2.6 eV (Figure 1d) that makes it inappropriate for light harvesting applications. We note here that bandgaps within the DFT-PBE approach for these two systems approximate well the experimental values (see Figure 2b,d).
The DFT can also give a clear first-order evaluation of the energetic equilibrium between the two phases (Figure 1e), where the ground-state energy of the yellow phase is lower by 54 meV per formula unit (corresponding to 216 meV per unit cell) from that of the γ black phase. This important difference is practically translated in the spontaneous phase transformation of the γ-perovskite towards the δ phase at room temperature. It is interesting to note here that this preference exists only for the iodide perovskite, whereas for bromide and chloride structures the ground-state energy of the perovskite γ-phase prevails over the δ phase. Indeed, experimentally, CsPbBr3 and CsPbCl3 perovskites are stable and the manifestation of a δ-orthorhombic phase similar to CsPbI3 is not observed [19,32].
In this respect, the main structural and optical characteristics of the two CsPbI3 phases can be observed in Figure 2, by means of X-ray diffraction and spectroscopic ellipsometry. In both cases a perovskite polycrystalline layer was obtained through solution processing of a mixture of PbI2, CsI, and EuCl3, where the presence of the Eu cation guarantees the formation of a relatively stable low-temperature perovskite phase at 80 °C [21], which persists after cooling down to room temperature for several days in a nitrogen ambient. We note here that similar results are obtained if EuI2 is used in the solution instead of EuCl3, as the europium cation acquires the +2 oxidation state within the perovskite, regardless of its initial state within the solution (see Ref. [26] for a comprehensive discussion on this issue). The XRD data (Figure 2a) show the characteristics of this low-temperature perovskite phase, with diagnostic peaks in the angular range 2θ~14–22°. A deconvolution of the peak at ~20.3° indicates the presence of three separate peak components that correspond to the (200), (112), and (020) planes of the orthorhombic γ-CsPbI3. It is therefore important to note that the prevailing perovskite phase at room temperature retains orthorhombic characteristics, in agreement with the lower ground-state energy of this structure as compared to the α and β polymorphs. By fitting the entire pattern with a γ-orthorhombic CsPbI3 lattice, the lattice parameters were extracted by means of a standard Rietveld refinement: a = 8.632 Å, b = 8.927 Å and c = 12.327 Å (Pbnm). This dark-colored perovskite phase unavoidably turns into the yellow δ phase at room temperature, either by a gradual transformation within a nitrogen environment (from days to weeks), or rapidly, if exposed to humid air. The main XRD characteristics of the γ-to-δ transformation (Figure 2c) are the peaks at low angles and in particular the diagnostic peaks at 2θ~9.9° and ~13.1°. The Rietveld refinement in this case reveals the following lattice parameters: a = 10.421 Å, b = 4.604 Å, c = 17.873 Å (Pnma). By calculating the respective unit cell volume, it becomes clear that a γ-to-δ transformation induces a volume contraction in the material.
Figure 2b,d shows the real (ε1) and imaginary (ε2) parts of the dielectric function for the previously discussed phases of CsPbI3 measured by spectroscopic ellipsometry. Optically, only the dark perovskite phase has a useful bandgap for devices and applications (~1.77 eV), whereas the yellow crystal is a wide band gap semiconductor (~2.9 eV), in good agreement with our theoretical calculations. Further optical differences appear in the determination of the critical points of the complex dielectric function and on the absorption capacity of the two phases at different photon energies.
By defining the basic structural and optical characteristics of the dark and yellow CsPbI3 phases at room temperature, a spontaneous question that arises is in regard to the physical features and quantities that could favor the preservation of the former under device-operating conditions. Interesting insights may come up from the microstructural investigation of the synthesized CsPbI3 films through high-resolution scanning transmission electron microscopy, or by theoretical calculations.

3.2. Role of Spatial Confinement in the Preservation of the Perovskite Phase of CsPbI3

To gain an atomic-scale understanding of the structural characteristics of our CsPbI3 films, CsPbI3 samples were prepared for STEM analysis directly on a C-coated Cu-grid to avoid specific preparation procedures that could alter the material properties. Figure 3a shows a microstructural arrangement of the low-temperature γ-perovskite phase, which is characterized by a continuous network of nanometer-sized grains (~5–20 nm diameter) that are identified from their typical perovskite diffraction pattern, along with inclusions of larger grains that correspond to the yellow phase. The perovskite nanograins are surrounded by a material with amorphous characteristics, which acts as an interconnecting network and renders the entire film compact [26]. The co-presence of both perovskite and yellow grains in the same area of this sample is indicative of an early transition stage and allows for a first consideration on the stability of the perovskite phase, which appears to depend on the small size of the perovskite crystals. Indeed, an analysis of all available samples by STEM showed that larger grains have diffraction patterns that undoubtably unveil the characteristic peaks of the yellow δ-phase. The presence of perovskite nanograins agrees with literature studies that have examined the microstructural arrangement of CsPbI3 films starting from different growth methods [24,33]. Indeed, it has been theoretically argued that the surface energy for many low-index Miller planes is significantly lower for the perovskite phase as compared to the yellow orthorhombic structure [24]. Therefore, it appears that a viable route for the stabilization of the perovskite phase at room temperature is by increasing the surface-to-volume ratio of CsPbI3 grains, i.e., by confining CsPbI3 in nanostructures. This is most probably one of the key roles of europium or other additives during the formation of the CsPbI3 perovskite at temperatures that are lower than 300 °C, above which a spontaneous transition to the α-phase takes place.
Figure 3b shows an atomic-resolution STEM image of a perovskite nanograin with a diameter of ~9 nm within a thin film that has almost globally transited into the δ-phase. The persistence of this perovskite nanostructure within the yellow layer further supports the role of spatial confinement for the preservation of the perovskite phase and indicates that the transformation from γ to δ may occur through a gradual merging of the initially small grains into bigger structures, where the surface-to-volume ratio decreases. The atomic-scale details of this image are the following: the (1 1 0) perovskite planes are clearly visible from the [1–11] zone axis direction, whereas surfaces seem to be surrounded by an amorphous-like shell. By performing a fine-scale analysis of the atomic structure, it emerges that the [PbI6]4− octahedra are tilted (resulting in a zigzag-like configuration for iodine atoms), as expected for the γ-orthorhombic polymorph of CsPbI3 (see also Figure 4, where a scheme of a CsPbI3 nanograin in the orthorhombic phase is shown). It is important to note here that tilting aids in the stabilization of the CsPbI3 perovskite phase, as our theoretical calculations show that the non-tilted α-polymorph has a ground-state energy that is higher than the γ-polymorph by ~124 meV per formula unit. Consequently, the stabilization of the α-phase at high temperatures should largely depend solely on vibrational and volumetric features.

3.3. Role of Electron Concentration and Volume Expansion in the Stabilization of the Perovskite Phase of CsPbI3

The high surface-to-volume ratio is an important feature that can shift the energetic equilibrium towards the perovskite phase of CsPbI3 at room temperature. However, the formation of thin films that are principally composed of nanograins raises questions on the transport properties of photogenerated carriers in the respective devices, as the presence of a nonuniform lattice could limit the carrier diffusion length and increase the electron-hole recombination rate. In this sense, it is important to investigate alternatives beyond the effect of spatial confinement that could lead to the stabilization of the perovskite phase in crystal grains of bigger dimensions. Within this context, ab initio calculations can be important for exploring some fundamental features of the γ and the δ phases with respect to charge and volume.
Figure 5a shows the energetic competition (at 0 K) between the γ and the δ phase upon an increase of the electron concentration in the material. The uncharged δ-structure has a ground-state energy that is lower by ~216 meV per unit cell with respect to the γ- structure. Upon electron doping, the energy difference tends to diminish and for an electron concentration of ~0.4 e per unit cell the γ perovskite becomes more stable than the yellow δ-phase. Electron doping could be ideally engineered by partially substituting one or both cationic species of CsPbI3 with elements that have a higher oxidation number with respect to Cs+ and Pb2+, while maintaining a proper coordination with the iodine atoms of the octahedral framework. This practically means that a viable way for the stabilization of the perovskite phase could be the partial substitution of monovalent Cs+ cations with divalent cations that can maintain the octahedral perovskite arrangement, or similarly by the substitution of Pb2+ with trivalent cations. The possibility of an n-doped CsPbI3 material could also open novel possibilities for the engineering of p-n heterojunctions in perovskite solar cell architectures. However, we note that a full stabilization of the CsPbI3 perovskite would require a very high level of doping, which could be solely achieved with the formation of stable and uniform perovskite alloys. The latter is one of the most challenging issues in the processing of metal-halide perovskites.
In addition to electron doping, another feature that can create a more favorable energetic framework for the preservation of the perovskite phase is the volume expansion of the material. Figure 5b shows the energetic diagram of the γ and δ phases as a function of the unit-cell volume. For each calculation, only the volume of the structures was fixed, whereas both lattice parameters and atoms were freely allowed to relax. The equilibrium volume of the δ-structure is lower than the one of the γ-perovskite, in agreement with the experimental XRD measurements. By increasing the volume of the lattice (e.g., through artificial strain or thermal expansion at high temperatures) this energetic difference becomes lower. However, even for a lattice expansion as high as 5% with respect to the equilibrium value, the yellow phase remains more stable, indicating that volume expansion can reduce the energetic difference between the two phases, but not stabilize by itself the perovskite structure.

4. Discussion

Considering a critical analysis of the available experimental and theoretical data, the main physical aspects that could stabilize the CsPbI3 material at room temperature are the spatial confinement in nanocrystalline structures and the level of electron doping in the material. Both features could be practically achieved through the control of the growth process, e.g., by inserting additives in the precursors that can accelerate the perovskite formation at relatively low temperatures though the increase of perovskite nucleation centers during solution processing, while concomitantly substituting the cationic species of the material with atoms that have a higher oxidation number. However, even if the resulting perovskite films are encapsulated in a quasi-inert ambient like nitrogen, a further obstacle that could compromise the medium to long-term stability (i.e., the stability after weeks or months) is the intrinsic capacity of lattice reconstruction in lead-iodide perovskite systems [34], which strongly depends on the weakness of the Pb-I bonds. The latter could accelerate the kinetic merging of the perovskite nanograins (coalescence) when a stable inter-grain blocking layer is absent, resulting in a reduction of the surface-to-volume ratio and a consequent transformation of the films towards the yellow polymorph. Moreover, another characteristic that is closely related to the weak lead-halogen ionic bonding is the self-healing capacity of lead-halide perovskites [35], which should tend to diminish the tolerance for non-intrinsic species over time. Within this context, a primary objective for the long-term stabilization of CsPbI3, in addition to the prevalence of the perovskite phase at room temperature, should be the structural reinforcement of the inorganic network. This can be achieved by the partial substitution of iodine with, e.g., bromine [36,37] (considering though the possible implications for the band gap of the material in the related CsPbIxBr3-x alloys), the addition of rubidium [38], or by engineering novel processes that could mitigate the facile Pb-I bond breaking and the subsequent iodide kinetics (which should also contribute to ionic currents during device operation).

5. Conclusions

We studied the physical aspects that could shift the balance in the energetic equilibrium between the perovskite and the yellow phase of CsPbI3, leading to a stabilization of the former at room temperature, by means of theoretical and experimental techniques. We found that two important features that could assist in the stabilization of the perovskite structure are the spatial confinement of CsPbI3 nanocrystals and the electron doping of the material. Both features could be engineered during solution processing, e.g., by introducing additives that accelerate the perovskite formation kinetics at relatively low temperatures while concomitantly offering cationic species that can substitute the intrinsic atoms of the material, increasing the electron concentration. We expect further developments on the formation of stable and long-lasting perovskite alloys in the forthcoming years that could mitigate the long-term deterioration of the perovskite structure.

Author Contributions

Conceptualization, I.D. and A.A.; methodology, I.D., C.B., and A.A.; software, I.D., G.F.; validation, G.F., G.M., E.S., S.S., S.V., and A.L.M.; investigation, I.D., C.B., G.M., and A.A.; resources, C.B., G.M., S.S., S.V., E.S., A.L.M., A.A.; writing, I.D.; supervision, A.L.M.; funding acquisition, A.L.M., A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially supported by the Italian ministry of education and research within the project BEST-4U (CUP B88D19000160005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Computational resources were provided by the Italian CINECA Consortium under the ISCRA project HYPERSOL (Ref. n. HP10BRZPQ8).

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. 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]
  2. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. Compositional engineering of perovskite materials for high-performance solar cells. Nature 2015, 517, 476–480. [Google Scholar] [CrossRef]
  3. McMeekin, D.P.; Sadoughi, G.; Rehman, W.; Eperon, G.E.; Saliba, M.; Hörantner, M.T.; Haghighirad, A.; Sakai, N.; Korte, L.; Rech, B.; et al. A mixed-cation lead mixed-halide perovskite absorber for tandem solar cells. Science 2016, 351, 151–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bi, D.; Tress, W.; Dar, M.I.; Gao, P.; Luo, J.; Renevier, C.; Schenk, K.; Abate, A.; Giordano, F.; Correa Baena, J.-P.; et al. Efficient luminescent solar cells based on tailored mixed-cation perovskites. Sci. Adv. 2016, 2, e1501170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Grancini, G.; Roldán-Carmona, C.; Zimmermann, I.; Mosconi, E.; Lee, X.; Martineau, D.; Narbey, S.; Oswald, F.; De Angelis, F.; Grätzel, M.; et al. One-Year stable perovskite solar cells by 2D/3D interface engineering. Nat. Commun. 2017, 8, 15684. [Google Scholar] [CrossRef] [PubMed]
  6. Colella, S.; Mosconi, E.; Fedeli, P.; Listorti, A.; Gazza, F.; Orlandi, F.; Ferro, P.; Besagni, T.; Rizzo, A.; Calestani, G.; et al. MAPbI3-xClx Mixed Halide Perovskite for Hybrid Solar Cells: The Role of Chloride as Dopant on the Transport and Structural Properties. Chem. Mater. 2013, 25, 4613–4618. [Google Scholar] [CrossRef]
  7. Wu, C.G.; Chiang, C.H.; Tseng, Z.L.; Nazeeruddin, M.K.; Hagfeldt, A.; Grätzel, M. High efficiency stable inverted perovskite solar cells without current hysteresis. Energy Environ. Sci. 2015, 8, 2725–2733. [Google Scholar] [CrossRef]
  8. Eperon, G.E.; Hörantner, M.T.; Snaith, H.J. Metal halide perovskite tandem and multiple-junction photovoltaics. Nat. Rev. Chem. 2017, 1, 1–18. [Google Scholar] [CrossRef]
  9. Best Research-Cell Efficiency Chart. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 15 January 2021).
  10. Holzhey, P.; Yadav, P.; Turren-Cruz, S.H.; Ummadisingu, A.; Grätzel, M.; Hagfeldt, A.; Saliba, M. A chain is as strong as its weakest link–stability study of MAPbI3 under light and temperature. Mater. Today 2019, 29, 10–19. [Google Scholar] [CrossRef]
  11. Pang, S.; Hu, H.; Zhang, J.; Lv, S.; Yu, Y.; Wei, F.; Qin, T.; Xu, H.; Liu, Z.; Cui, G. NH2CHNH2PbI3: An alternative organolead iodide perovskite sensitizer for mesoscopic solar cells. Chem. Mater. 2014, 26, 1485–1491. [Google Scholar] [CrossRef]
  12. Deretzis, I.; Smecca, E.; Mannino, G.; La Magna, A.; Miyasaka, T.; Alberti, A. Stability and degradation in hybrid perovskites: Is the glass half-empty or half-full? J. Phys. Chem. Lett. 2018, 9, 3000–3007. [Google Scholar] [CrossRef] [PubMed]
  13. Smecca, E.; Numata, Y.; Deretzis, I.; Pellegrino, G.; Boninelli, S.; Miyasaka, T.; La Magna, A.; Alberti, A. Stability of solution-processed MAPbI 3 and FAPbI 3 layers. Phys. Chem. Chem. Phys. 2016, 18, 13413–13422. [Google Scholar] [CrossRef] [PubMed]
  14. Alberti, A.; Bongiorno, C.; Smecca, E.; Deretzis, I.; La Magna, A.; Spinella, C. Pb clustering and PbI 2 nanofragmentation during methylammonium lead iodide perovskite degradation. Nat. Commun. 2019, 10, 2196. [Google Scholar] [CrossRef] [PubMed]
  15. Aziz, A.; Aristidou, N.; Bu, X.; Westbrook, R.J.; Haque, S.A.; Islam, M.S. Understanding the enhanced stability of bromide substitution in lead iodide perovskites. Chem. Mater. 2020, 32, 400–409. [Google Scholar] [CrossRef]
  16. Fisicaro, G.; La Magna, A.; Alberti, A.; Smecca, E.; Mannino, G.; Deretzis, I. Local Order and Rotational Dynamics in Mixed A-Cation Lead Iodide Perovskites. J. Phys. Chem. Lett. 2020, 11, 1068–1074. [Google Scholar] [CrossRef]
  17. Schelhas, L.T.; Li, Z.; Christians, J.A.; Goyal, A.; Kairys, P.; Harvey, S.P.; Kim, D.H.; Stone, K.H.; Luther, J.M.; Zhu, K.; et al. Insights into operational stability and processing of halide perovskite active layers. Energy Environ. Sci. 2019, 12, 1341–1348. [Google Scholar] [CrossRef]
  18. Eperon, G.E.; Paternò, G.M.; Sutton, R.J.; Zampetti, A.; Haghighirad, A.A.; Cacialli, F.; Snaith, H.J. Inorganic caesium lead iodide perovskite solar cells. J. Mater. Chem. A 2015, 3, 19688–19695. [Google Scholar] [CrossRef]
  19. Mannino, G.; Deretzis, I.; Smecca, E.; La Magna, A.; Alberti, A.; Ceratti, D.; Cahen, D. Temperature-Dependent Optical Band Gap in CsPbBr3, MAPbBr3, and FAPbBr3 Single Crystals. J. Phys. Chem. Lett. 2020, 11, 2490–2496. [Google Scholar] [CrossRef]
  20. Ke, W.; Spanopoulos, I.; Stoumpos, C.C.; Kanatzidis, M.G. Myths and reality of HPbI 3 in halide perovskite solar cells. Nat. Commun. 2018, 9, 4785. [Google Scholar] [CrossRef]
  21. Jena, A.K.; Kulkarni, A.; Sanehira, Y.; Ikegami, M.; Miyasaka, T. Stabilization of α-CsPbI3 in ambient room temperature conditions by incorporating Eu into CsPbI3. Chem. Mater. 2018, 30, 6668–6674. [Google Scholar] [CrossRef]
  22. Steele, J.A.; Jin, H.; Dovgaliuk, I.; Berger, R.F.; Braeckevelt, T.; Yuan, H.; Martin, C.; Solano, E.; Lejaeghere, K.; Rogge, S.M.; et al. Thermal unequilibrium of strained black CsPbI3 thin films. Science 2019, 365, 679–684. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, T.; Dar, M.I.; Li, G.; Xu, F.; Guo, N.; Grätzel, M.; Zhao, Y. Bication lead iodide 2D perovskite component to stabilize inorganic α-CsPbI3 perovskite phase for high-efficiency solar cells. Sci. Adv. 2017, 3, e1700841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhao, B.; Jin, S.F.; Huang, S.; Liu, N.; Ma, J.Y.; Xue, D.J.; Han, Q.; Ding, J.; Ge, Q.-Q.; Feng, Y.; et al. Thermodynamically stable orthorhombic γ-CsPbI3 thin films for high-performance photovoltaics. J. Am. Chem. Soc. 2018, 140, 11716–11725. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, P.; Zhang, X.; Zhou, Y.; Jiang, Q.; Ye, Q.; Chu, Z.; Li, X.; Yang, X.; Yin, Z.; You, J. Solvent-controlled growth of inorganic perovskite films in dry environment for efficient and stable solar cells. Nat. Commun. 2018, 9, 2225. [Google Scholar] [CrossRef] [PubMed]
  26. Alberti, A.; Smecca, E.; Deretzis, I.; Mannino, G.; Bongiorno, C.; Valastro, S.; Sanzaro, S.; Fisicaro, G.; Jena, A.K.; Numata, Y.; et al. Formation of CsPbI3 γ-phase at 80 °C by Europium assisted snowplow Effect. Adv. Energy Sustain. Res. 2021, 2100091. [Google Scholar] [CrossRef]
  27. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Guido, L.C.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter. 2009, 21, 395502. [Google Scholar] [CrossRef]
  28. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  29. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892. [Google Scholar] [CrossRef]
  30. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  31. Sutton, R.J.; Filip, M.R.; Haghighirad, A.A.; Sakai, N.; Wenger, B.; Giustino, F.; Snaith, H.J. Cubic or orthorhombic? Revealing the crystal structure of metastable black-phase CsPbI3 by theory and experiment. ACS Energy Lett. 2018, 3, 1787–1794. [Google Scholar] [CrossRef]
  32. Cohen, M.I.; Young, K.F.; Chang, T.T.; Brower Jr, W.S. Phase transitions in CsPbCl3. J. Appl. Phys. 1971, 42, 5267–5272. [Google Scholar] [CrossRef]
  33. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Alberti, A.; Deretzis, I.; Mannino, G.; Smecca, E.; Giannazzo, F.; Listorti, A.; Colella, S.; Masi, S.; La Magna, A. Nitrogen soaking promotes lattice recovery in polycrystalline hybrid perovskites. Adv. Energy Mater. 2019, 9, 1803450. [Google Scholar] [CrossRef]
  35. Ceratti, D.R.; Rakita, Y.; Cremonesi, L.; Tenne, R.; Kalchenko, V.; Elbaum, M.; Oron, D.; Potenza, M.A.C.; Hodes, G.; Cahen, D. Self-Healing Inside APbBr3 Halide Perovskite Crystals. Adv. Mater. 2018, 30, 1706273. [Google Scholar] [CrossRef]
  36. Guo, Z.; Jena, A.K.; Takei, I.; Kim, G.M.; Kamarudin, M.A.; Sanehira, Y.; Ishii, A.; Numata, Y.; Hayase, S.; Miyasaka, T. VOC Over 1.4 V for Amorphous Tin-Oxide-Based Dopant-Free CsPbI2Br Perovskite Solar Cells. J. Am. Chem. Soc. 2020, 142, 9725–9734. [Google Scholar] [CrossRef]
  37. Mannino, G.; Deretzis, I.; Smecca, E.; Giannazzo, F.; Valastro, S.; Fisicaro, G.; La Magna, A.; Ceratti, D.; Alberti, A. CsPbBr3, MAPbBr3, and FAPbBr3 Bromide Perovskite Single Crystals: Interband Critical Points under Dry N2 and Optical Degradation under Humid Air. J. Phys. Chem. C 2021, 125, 4938–4945. [Google Scholar] [CrossRef]
  38. Bernasconi, A.; Rizzo, A.; Listorti, A.; Mahata, A.; Mosconi, E.; De Angelis, F.; Malavasi, L. Synthesis, Properties, and Modeling of Cs1−xRbxSnBr3 Solid Solution: A New Mixed-Cation Lead-Free All-Inorganic Perovskite System. Chem. Mater. 2019, 31, 3527–3533. [Google Scholar] [CrossRef]
Figure 1. (a) Scheme of the γ-CsPbI3 perovskite phase seen from the (001) plane; (b) Scheme of the orthorhombic δ-CsPbI3 yellow phase seen from the (010) plane; (c) Band structure of the γ-CsPbI3 perovskite phase; (d) Band structure of the orthorhombic δ-CsPbI3 yellow phase; (e) Difference in the calculated ground state energy (at 0 K) between the γ and the δ phases of CsPbX3 perovskites (where X = I, Br, Cl).
Figure 1. (a) Scheme of the γ-CsPbI3 perovskite phase seen from the (001) plane; (b) Scheme of the orthorhombic δ-CsPbI3 yellow phase seen from the (010) plane; (c) Band structure of the γ-CsPbI3 perovskite phase; (d) Band structure of the orthorhombic δ-CsPbI3 yellow phase; (e) Difference in the calculated ground state energy (at 0 K) between the γ and the δ phases of CsPbX3 perovskites (where X = I, Br, Cl).
Nanomaterials 11 01282 g001
Figure 2. Structural and optical properties of the γ-CsPbI3 perovskite phase and its transition towards the δ-CsPbI3 yellow phase at room temperature: (a) XRD spectrum of the γ-CsPbI3 phase; (b) Real (ε1) and imaginary (ε2) parts of the dielectric function for the γ-CsPbI3 phase; (c) XRD spectrum of the δ-CsPbI3 yellow phase; (d) Real (ε1) and imaginary (ε2) parts of the dielectric function for the δ-CsPbI3 yellow phase.
Figure 2. Structural and optical properties of the γ-CsPbI3 perovskite phase and its transition towards the δ-CsPbI3 yellow phase at room temperature: (a) XRD spectrum of the γ-CsPbI3 phase; (b) Real (ε1) and imaginary (ε2) parts of the dielectric function for the γ-CsPbI3 phase; (c) XRD spectrum of the δ-CsPbI3 yellow phase; (d) Real (ε1) and imaginary (ε2) parts of the dielectric function for the δ-CsPbI3 yellow phase.
Nanomaterials 11 01282 g002
Figure 3. (a) STEM image of a γ-CsPbI3 perovskite thin film with δ-CsPbI3 yellow phase inclusions; (b) High-resolution STEM image of a γ-CsPbI3 nanograin within a δ-CsPbI3 thin film.
Figure 3. (a) STEM image of a γ-CsPbI3 perovskite thin film with δ-CsPbI3 yellow phase inclusions; (b) High-resolution STEM image of a γ-CsPbI3 nanograin within a δ-CsPbI3 thin film.
Nanomaterials 11 01282 g003
Figure 4. Scheme of a CsPbI3 nanograin from the [1–11] zone axis showing the (110) crystallographic planes and the two different types of inorganic layers (Pb-I and Cs-I).
Figure 4. Scheme of a CsPbI3 nanograin from the [1–11] zone axis showing the (110) crystallographic planes and the two different types of inorganic layers (Pb-I and Cs-I).
Nanomaterials 11 01282 g004
Figure 5. (a) Difference in the calculated ground-state energy per unit cell between the γ and the δ phases of CsPbI3 as a function of electron doping; (b) Relative ground-state energies of γ-CsPbI3 and δ-CsPbI3 upon volume expansion and contraction. The reference energy at 0 eV has been set to the minimum energy of the δ phase.
Figure 5. (a) Difference in the calculated ground-state energy per unit cell between the γ and the δ phases of CsPbI3 as a function of electron doping; (b) Relative ground-state energies of γ-CsPbI3 and δ-CsPbI3 upon volume expansion and contraction. The reference energy at 0 eV has been set to the minimum energy of the δ phase.
Nanomaterials 11 01282 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Deretzis, I.; Bongiorno, C.; Mannino, G.; Smecca, E.; Sanzaro, S.; Valastro, S.; Fisicaro, G.; La Magna, A.; Alberti, A. Exploring the Structural Competition between the Black and the Yellow Phase of CsPbI3. Nanomaterials 2021, 11, 1282. https://doi.org/10.3390/nano11051282

AMA Style

Deretzis I, Bongiorno C, Mannino G, Smecca E, Sanzaro S, Valastro S, Fisicaro G, La Magna A, Alberti A. Exploring the Structural Competition between the Black and the Yellow Phase of CsPbI3. Nanomaterials. 2021; 11(5):1282. https://doi.org/10.3390/nano11051282

Chicago/Turabian Style

Deretzis, Ioannis, Corrado Bongiorno, Giovanni Mannino, Emanuele Smecca, Salvatore Sanzaro, Salvatore Valastro, Giuseppe Fisicaro, Antonino La Magna, and Alessandra Alberti. 2021. "Exploring the Structural Competition between the Black and the Yellow Phase of CsPbI3" Nanomaterials 11, no. 5: 1282. https://doi.org/10.3390/nano11051282

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