Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2014 Apr 30;111(19):6900–6904. doi: 10.1073/pnas.1405780111

Switchable S = 1/2 and J = 1/2 Rashba bands in ferroelectric halide perovskites

Minsung Kim a,1, Jino Im b,1, Arthur J Freeman b, Jisoon Ihm a,2, Hosub Jin a,c,2
PMCID: PMC4024856  PMID: 24785294

Significance

Manipulation of electron spins is an important research subject to achieve breakthroughs in current electronics. The Rashba effect is the spin energy level splitting due to broken inversion symmetry and spin–orbit coupling, and the controllability of Rashba splitting is crucial to spin manipulation. Here, we report a theoretical investigation of Rashba band splitting in ferroelectric halide perovskite materials. Since the polarization direction in ferroelectric materials can be switched by external electric fields, Rashba splitting can be controlled conveniently in these materials. Interestingly, ferroelctric polarizations give rise to two distinct Rashba bands of contrasting orbital and spin characters. Taking advantage of the unique band characteristics, these materials prove promising candidates for the switchable Rashba effect.

Keywords: electronic structure, density functional theory, effective Hamiltonian

Abstract

The Rashba effect is spin degeneracy lift originated from spin–orbit coupling under inversion symmetry breaking and has been intensively studied for spintronics applications. However, easily implementable methods and corresponding materials for directional controls of Rashba splitting are still lacking. Here, we propose organic–inorganic hybrid metal halide perovskites as 3D Rashba systems driven by bulk ferroelectricity. In these materials, it is shown that the helical direction of the angular momentum texture in the Rashba band can be controlled by external electric fields via ferroelectric switching. Our tight-binding analysis and first-principles calculations indicate that Inline graphic and Inline graphic Rashba bands directly coupled to ferroelectric polarization emerge at the valence and conduction band edges, respectively. The coexistence of two contrasting Rashba bands having different compositions of the spin and orbital angular momentum is a distinctive feature of these materials. With recent experimental evidence for the ferroelectric response, the halide perovskites will be, to our knowledge, the first practical realization of the ferroelectric-coupled Rashba effect, suggesting novel applications to spintronic devices.


The Rashba effect has been widely investigated in 2D surfaces, interfaces, quantum wells, and 3D bulk systems (16). The essential requisite for the Rashba effect is that the spin degeneracy is lifted by the inversion symmetry-breaking (ISB) field in the presence of the spin–orbit coupling (SOC). To date, major concerns have focused on enlarging Rashba strength characterized by the Rashba coefficient Inline graphic (3, 5, 6). The controllability in the direction of the ISB field, on the other hand, has not been seriously considered. In a surface or an interface, the potential gradient generated by the structural inversion asymmetry results in the loss of controllability; the field direction is mainly fixed according to the preformed surface or interface configuration. The situation is similar in recently discovered 3D Rashba material BiTeI (6), because this material has the compositional ISB field between Te and I layers.

Controlling the ISB field and ultimately Rashba-type band splitting can be achieved by using a novel ferroelectric Rashba material (7, 8). In a ferroelectric system, the bulk polarization controlled by external electric fields generates the ISB field. Therefore, the ferroelectric polarization directly couples to the spin splitting and the helical spin texture in the ferroelectric Rashba material, enabling the helicity reversal via the ferroelectric switching. Recent theoretical study suggested GeTe as a possible candidate for this mechanism, but the direct measurement of the ferroelectric polarization and switching is still missing due to the sizable conductivity of bulk GeTe (7, 8).

We consider organic–inorganic hybrid metal halide perovskites as promising ferroelectric Rashba materials. The general formula for this material class is Inline graphic where Inline graphic, i.e., methylammonium (MA); M = Pb and Sn; and X = I and Br. The materials have several advantages over GeTe. Firstly, polar distortions and ferroelectric responses in the halide perovskite have been clearly observed in experiments (9, 10). On the other hand, GeTe inevitably generates Ge vacancies to give considerable bulk conductivity, which in turn hinders the polarization switching (11). Secondly, unlike the multiple band edges and strong hexagonal warping in GeTe (7, 8), the band edge states of the halide perovskites lie at a single point in the Brillouin zone with nearly isotropic Rashba bands. Thus, the halide perovskites have ideal Rashba-split bands with proper material quality. Finally, whereas GeTe has an indirect band gap (11), the halide perovskites have a direct one. This direct gap becomes important when we consider the transition between the valence and conduction bands in optical devices.

Interestingly, the above halide perovskite series has two contrasting types of Rashba bands simultaneously: the Inline graphic and Inline graphic Rashba bands at the valence and conduction bands, respectively (Fig. 1). This originates from the different band characters in terms of the angular momentum at the valance and conduction band edges common in halide perovskite compounds (12). The angular momentum character of the individual band can be affected by relative energy scales of the crystal field and SOC; the crystal field quenches orbital degrees of freedom, whereas SOC entangles spin and orbital degrees. Through the competition between the two energy scales, the band character on which the Rashba Hamiltonian is based can vary from the fully spin–orbital entangled total angular momentum state (J) to the spin state (S) (13), which causes a significant distinction in the angular momentum texture of the Rashba band.

Fig. 1.

Fig. 1.

Schematic illustration of the switchable Rashba effect. The polarization-coupled Rashba-type band splitting of the Inline graphic and Inline graphic manifolds in hybrid metal halide perovskites.

In this work, we examine the electronic structures of the hybrid metal halide perovskites as candidates for the ferroelectric Rashba materials. By constructing a minimal tight-binding (TB) model Hamiltonian, we can capture the key features of the band structures; the low-energy effective Hamiltonian gives rise to the ferroelectric-coupled Inline graphic and Inline graphic Rashba bands at the valence band maximum (VBM) and the conduction band minimum (CBM), respectively. The Rashba-type splitting of the fully spin–orbital-entangled Inline graphic subspace stems from the bandgap-independent intraorbital as well as the bandgap-dependent interorbital terms. On the contrary, Inline graphic Rashba splitting only consists of the interorbital term (14). We present several examples of possible ferroelectric Rashba materials (β-MAPbI3, β-MASnI3, and ortho-MASbBr3) by adopting first-principles electronic structure calculations based on density functional theory. These halide perovskites are shown to have the characteristic features predicted in the TB model with the sizable Rashba coefficient Inline graphic. A different type of controllability on the relative helicity between the two Rashba bands is also discussed according to the positions of the lateral halide atoms.

Results and Discussion

To reveal the microscopic origin of the ferroelectric-driven Rashba effect, we construct a TB Hamiltonian which contains additional hopping channels allowed by broken inversion symmetry. In the cubic phase of Inline graphic halide perovskites, CBM (VBM) consists of triply degenerate bands (a single band) in the absence of SOC, whose orbital angular momentum L = 1 (L = 0). The triple bands come from the p orbitals of the M-site atom, and the single band from the s orbital of M and p orbitals of the surrounding halogens. In the presence of SOC, the triply degenerate L = 1 states split into lower Inline graphic and upper Inline graphic bands whereas SOC is ineffective in the single band, resulting in the Inline graphic CBM and Inline graphic VBM (Fig. 2A). Under the ISB field, the TB Hamiltonian reads

graphic file with name pnas.1405780111eq1.jpg

with the s and p orbital states at the M site as the minimal basis set. Here, the first two terms describe the electronic structure of the high symmetric cubic perovskites (12), and the last term, Inline graphic, breaks the inversion symmetry by ferroelectric polarization. Most importantly, additional orbital mixings between the nearest neighbors emerge in Inline graphic; for example, the ISB field along the z direction allows the hopping between the nearest s and Inline graphic orbitals, Inline graphic, as well as the nearest Inline graphic (or Inline graphic) and Inline graphic orbitals, Inline graphic. The general form of the hopping parameters induced by the ISB field in the arbitrary direction reads Inline graphic (Inline graphic or Inline graphic). For simplicity, we omit the on-site mixing of s and p states. (For details on the TB model, see SI Text.)

Fig. 2.

Fig. 2.

The band structures and angular momentum textures of the TB model. The band structures are depicted (A) without an ISB field and (B) with Inline graphic and Inline graphic having the ferroelectric moment along the [001] direction. In B, the helical direction of the orbital angular momenta is denoted by CW (clockwise) or CCW (counterclockwise) in red color near CBM whereas that of the spin angular momenta is shown in blue near VBM. The orbital (red solid arrows) and spin (blue dotted arrows) angular momenta are presented in (C) VBM and (D) CBM bands near the A point of the Brillouin zone for Inline graphic with the constant energy contours at the horizontal energy cuts in B.

The eight-band TB model in Eq. 1 is solved numerically as shown in Fig. 2. When the ISB term is introduced in the z direction, i.e., with the ferroelectric moment along the [001] direction, Rashba-type band splitting occurs near the A point of the tetragonal Brillouin zone (Fig. 2B). To understand the splitting mechanism, we derive the Inline graphic effective Hamiltonian for the Inline graphic manifold near VBM up to Inline graphic

graphic file with name pnas.1405780111eq2.jpg

with Inline graphic, and the Inline graphic manifold near CBM

graphic file with name pnas.1405780111eq3.jpg

with Inline graphic, and Inline graphic, where Inline graphic is the band gap without the ISB term, λ is the SOC strength, t are hopping parameters, I is the Inline graphic identity matrix, and Inline graphic are the Pauli matrices describing the Inline graphic degrees of freedom. The Inline graphic and Inline graphic are functions that describe the parabolic band dispersion. The form of the splitting terms corresponds to the Bychkov–Rashba Hamiltonian (1), where the ISB field term is represented by the ferroelectric polarization and Inline graphic and Inline graphic determine the direction and magnitude of the splitting. The ferroelectric switching flips the direction of Inline graphic and in turn reverses the helical angular momentum texture of the Rashba band. [This helicity reversal does not depend on the specific form of the effective Hamiltonian (SI Text).]

Although the splitting terms in the effective Hamiltonian for VBM and CBM have the same form as the ferroelectric-coupled Rashba-type band splitting, they have quite different aspects. Firstly, the splitting occurs mainly in the Inline graphic space in VBM whereas it does in the Inline graphic space in CBM, which results from the original band characters. Consequently, the spin angular momentum is conspicuous in VBM, whereas the orbital angular momentum dominates in CBM (Fig. 2 C and D). Secondly, the origin of the splitting terms in CBM (Eq. 3) is the intraorbital coupling within the Inline graphic manifold as well as the sp coupling, whereas that in VBM (Eq. 2) is only the interorbital coupling between the s and p states. It is remarkable that the intraorbital splitting term within the Inline graphic manifold (the Inline graphic term) contains no other energy scale except for Inline graphic, which can survive in a large band gap system. Note that the splitting terms from the sp coupling (the Inline graphic terms in Eqs. 2 and 3) inevitably include the prefactors containing the hopping parameter and the energy difference of the s and p states in contrast.

With the qualitative understanding in the TB model, we investigate a series of the ferroelectric metal halide perovskites using first-principles calculations. Perovskites are known to show various structural phase transitions that may include the ferroelectric phase (15). Relative sizes of constituent atoms are important for the ferroelectricity in the perovskite structure (16). Indeed, there have been experimental attempts to establish ferroelectric phases by using organic molecules at the A sites (9, 10, 17). Among them, β-MAPbI3 was found to have polar distortion, and a clear signature of the ferroelectric response was observed at room temperature (9). Other compounds such as MASnI3 make transitions to the β-phase at a lower temperature and are expected to show similar behavior. Once the ferroelectricity is established, hybrid metal halide perovskites could be good candidates for ferroelectric Rashba materials due to the large SOC inherent in Pb and Sn.

β-MAPbI3 has a Inline graphic tetragonal unit cell with four formula units. There are octahedron rotations in the basal plane, and small polar distortions occur along the z direction. The difference between the longer and shorter Pb—I bond length resulting from the off-centering of the Pb atom is Inline graphic Å. The atomic and electronic structures of β-MAPbI3 are depicted in Fig. 3 A and B. VBM and CBM are located near the Γ point where clear Rashba-type band splitting appears. Due to the original character of each band, namely Inline graphic valence and Inline graphic conduction bands, main contributions to the angular momenta of VBM come from the spin, whereas CBM is dominated by orbital angular momentum components (Fig. 3 CE) in accordance with our TB analysis. The two Rashba bands are contrahelical, i.e., they have the opposite helicity and the constant energy contours show a nearly isotropic shape. Here, we assign the helical direction to the spin (orbital) angular momenta in VBM (CBM) as the principal contributions come from the spin (orbital). Because the material intrinsically has a 3D structure, it has a dispersive band without band splitting in the polarization direction as shown in the dispersion along the Γ–Z line (Fig. 3B). To see the Inline graphic dependence of Rashba splitting, the band structures in three different Inline graphic planes are depicted in Fig. 3F. The size of the splitting diminishes as Inline graphic increases.

Fig. 3.

Fig. 3.

Electronic and atomic structures of β-MAPbI3 from the first-principles calculations. (A) The atomic and (B) band structures of β-MAPbI3 are shown. The spin (blue dotted arrows) and orbital (red solid arrows) angular momenta are depicted with the constant energy contours near (C) Inline graphic and (D) Inline graphic with Inline graphic and Inline graphic. (E) The magnitudes of the spin (blue dotted circles) and orbital (red solid circles) angular momenta for the VBM and CBM bands. (F) The band structure with different Inline graphic values. Here, Inline graphic is the reciprocal lattice vector along the z direction.

The strength of the Rashba effect is characterized by the Rashba coefficient Inline graphic. By using the energy difference (Inline graphic, Rashba energy) and the momentum offset Inline graphic between the band crossing point and the band extremum, the Rashba coefficient Inline graphic can be obtained. The estimated values for β-MAPbI3 near CBM are Inline graphic, Inline graphic, and Inline graphic for Γ–X. The calculation for Γ–M gives practically the same value, indicating the isotropic structure of the band splitting. Rashba coefficients of other candidates in the hybrid metal halide perovskite series are listed in Table 1, where the Rashba coefficients are of order 1. The sizes of the Rashba coefficients are comparable to those in the giant Rashba materials, BiTeI and GeTe (68). When replacing Pb with Sn, various physical quantities (SOC strength, the positions of Sn and I atoms, and the band gap) can affect the amount of the splitting. In β-MASnI3, the SOC strength is reduced but the off-centering of Sn is slightly larger (the bond length difference Inline graphic) than β-MAPbI3, resulting in the similar Rashba coefficient. In case of CBM in ortho-MASnBr3, the Rashba energy is larger, but the angular momentum texture does not correspond to the characteristic Inline graphic Rashba band, possibly due to the significant distortions in the compound.

Table 1.

Estimated Rashba energy Inline graphic, momentum offset Inline graphic, and Rashba coefficient Inline graphic

Compound Band Direction Inline graphic Inline graphic Inline graphic
β-MAPbI3 CBM Γ–X 12 0.015 1.5
VBM Γ–X 11 0.016 1.4
β-MASnI3 CBM Γ–X 11 0.011 1.9
VBM Γ–X 4 0.006 1.2
ortho-MASnBr3 VBM X–Γ 15 0.031 1.0
VBM X–S 13 0.031 0.84
VBM X–Y 14 0.031 0.92

According to the effective Hamiltonians in Eqs. 2 and 3, the helicity of the angular momentum in CBM is determined by both Inline graphic and Inline graphic, whereas that in VBM depends on Inline graphic only. This provides additional controllability in the relative helicity between VBM and CBM Rashba bands. According to the relative sizes and directions of Inline graphic and Inline graphic, the two Rashba bands can have the same helicity (cohelical) or an opposite one (contrahelical). A series of band structures is plotted from the TB model by varying the value of Inline graphic but with a fixed Inline graphic; the relative helicity between VBM and CBM can be tuned from contrahelical (Fig. 2B) to cohelical (Fig. 4B). In the intermediate region, the band crossing occurs in the Inline graphic manifold (Fig. 4A). In real materials, the positions of the lateral halide atoms can determine the relative helicity (SI Text). It is to be noted that, although the helicity of VBM is determined only by Inline graphic in our simplified model, there might be other terms if the VBM band couples to other bands (for example, the halogen p bands) appropriately. In general, the Rashba splitting term of the effective Hamiltonian can be written as

graphic file with name pnas.1405780111eq4.jpg

where Inline graphic originate from the intra- or interband coupling under the ferroelectric-induced ISB field. The relative magnitudes of Inline graphic and directions of Inline graphic will determine the helicity of the Rashba band.

Fig. 4.

Fig. 4.

Relative helicity change. The band structures of the TB model (A) with Inline graphic and Inline graphic, and (B) with Inline graphic and Inline graphic.

We also want to mention a characteristic aspect of Rashba band splitting in the halide perovskites. In 2D Rashba systems, the asymmetry of the crystal field is relatively large, resulting in, for example, Inline graphic-like surface states with interband Rashba splitting (18). In contrast, the starting point of the effective Hamiltonian near the Fermi energy in the halide perovskites is the high-symmetric cubic structure where the triple degeneracy of CBM is present due to the cubic symmetry. Thus, the inclusion of SOC under the ISB field gives rise to the almost fully spin–orbital entangled Inline graphic Rashba band that contains the bandgap-independent intraband splitting term. In the intraband splitting, the orbital components of the Inline graphic states are essential as the ISB field mixes only the orbital part. This unique intraband splitting mechanism would be important in large gap Inline graphic Rashba materials.

It is interesting to consider the effect of low dimensionality in layered perovskite materials. In the case of the 2D limit, the threefold degeneracy at CBM is broken due to the crystal field. For a monolayer of PbI6 octahedra with polar distortions, we find that CBM states have different orbital composition from the Inline graphic bands in the 3D limit. (The electronic structure calculation for the 2D layer is presented in SI Text.) Thus, in 2D layered structures the Rashba-type splitting near the band edge states occurs with a different band character in CBM. The ferroelectric 2D Rashba system could be realized in hybrid layered halide perovskites experimentally (19, 20).

Conclusions

In conclusion, we have investigated the electronic structures of hybrid metal halide perovskites with ferroelectric distortions. We have shown that ferroelectric polarization switching enables the angular momentum helicity reversal in the band edge states. The band structure is intriguing in that two contrasting Inline graphic and Inline graphic Rashba bands appear simultaneously and the Inline graphic Rashba band contains the bandgap-independent intraband splitting term. The coexistence of Inline graphic and Inline graphic Rashba bands would critically influence the optical transitions between the two Rashba bands. Thus, the controllable Rashba effect could provide opportunities to exploit the spin and orbital degrees of freedom in the photoinduced effect of photovoltaics (2124) that has increasing importance in the energy materials research.

Methods

The density functional theory calculations were performed using Vienna Ab initio Simulation Package (25, 26) with projector augmented-wave potentials (27). The PBEsol (Perdew–Burke–Ernzerhof revised for solids) exchange-correlation functional was adopted (28). A plane–wave basis set with the energy cutoff of 520 eV was used. Experimental lattice constants and internal coordinates were used. The effect of the SOC was included using the second variational method (29). The constant energy surfaces were obtained from the interpolated band structures by maximally localized Wannier functions as implemented in the Wannier90 package (3032).

Supplementary Material

Supporting Information

Acknowledgments

We thank Cheol-Hwan Park for fruitful discussions and comments. Computations were performed through the support of the Korea Institute of Science and Technology Information (KISTI). This work was supported by the National Research Foundation (NRF) of Korea through Ministry of Education, Science, and Technology (MEST) Grant 2006-0093853 and the Institute for Basic Science (IBS) in Korea.

Footnotes

The authors declare no conflict of interest.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1405780111/-/DCSupplemental.

References

  • 1.Bychokov YA, Rashba EI. Properties of a 2D electron gas with lifted spectral degeneracy. JETP Lett. 1984;39(2):78. [Google Scholar]
  • 2.LaShell S, McDougall BA, Jensen E. Spin splitting of an Au(111) surface state band observed with angle resolved photoelectron spectroscopy. Phys Rev Lett. 1996;77(16):3419–3422. doi: 10.1103/PhysRevLett.77.3419. [DOI] [PubMed] [Google Scholar]
  • 3.Ast CR, et al. Giant spin splitting through surface alloying. Phys Rev Lett. 2007;98(18):186807. doi: 10.1103/PhysRevLett.98.186807. [DOI] [PubMed] [Google Scholar]
  • 4.Dil JH, et al. Rashba-type spin-orbit splitting of quantum well states in ultrathin Pb films. Phys Rev Lett. 2008;101(26):266802. doi: 10.1103/PhysRevLett.101.266802. [DOI] [PubMed] [Google Scholar]
  • 5.Mathias S, et al. Quantum-well-induced giant spin-orbit splitting. Phys Rev Lett. 2010;104(6):066802. doi: 10.1103/PhysRevLett.104.066802. [DOI] [PubMed] [Google Scholar]
  • 6.Ishizaka K, et al. Giant Rashba-type spin splitting in bulk BiTeI. Nat Mater. 2011;10(7):521–526. doi: 10.1038/nmat3051. [DOI] [PubMed] [Google Scholar]
  • 7.Di Sante D, Barone P, Bertacco R, Picozzi S. Electric control of the giant Rashba effect in bulk GeTe. Adv Mater. 2013;25(4):509–513. doi: 10.1002/adma.201203199. [DOI] [PubMed] [Google Scholar]
  • 8.Di Sante D, Barone P, Bertacco R, Picozzi S. Correction: Electric control of the giant Rashba effect in bulk GeTe. Adv Mater. 2013;25(27):3625–3626. doi: 10.1002/adma.201203199. [DOI] [PubMed] [Google Scholar]
  • 9.Stoumpos CC, Malliakas CD, Kanatzidis MG. Semiconducting tin and lead iodide perovskites with organic cations: Phase transitions, high mobilities, and near-infrared photoluminescent properties. Inorg Chem. 2013;52(15):9019–9038. doi: 10.1021/ic401215x. [DOI] [PubMed] [Google Scholar]
  • 10.Swainson I, et al. Orientational ordering, tilting and lone-pair activity in the perovskite methylammonium tin bromide, CH3NH3SnBr3. Acta Crystallogr B. 2010;66(Pt 4):422–429. doi: 10.1107/S0108768110014734. [DOI] [PubMed] [Google Scholar]
  • 11.Picozzi S. 2013. Ferroelectric Rashba semiconductors as a novel class of multifunctional materials. arXiv:1312.0095.
  • 12.Jin H, Im J, Freeman AJ. Topological insulator phase in halide perovskite structures. Phys Rev B. 2012;86(12):121102. [Google Scholar]
  • 13.Park SR, Kim CH, Yu J, Han JH, Kim C. Orbital-angular-momentum based origin of Rashba-type surface band splitting. Phys Rev Lett. 2011;107(15):156803. doi: 10.1103/PhysRevLett.107.156803. [DOI] [PubMed] [Google Scholar]
  • 14.Bahramy MS, Arita R, Nagaosa N. Origin of giant bulk Rashba splitting: Application to BiTeI. Phys Rev B. 2011;84(4):041202. [Google Scholar]
  • 15.Cohen RE. Origin of ferroelectricity in perovskite oxides. Nature. 1992;358(6382):136–138. [Google Scholar]
  • 16.Singh DJ, Ghita M, Fornari M, Halilov SV. Role of A-site and B-site ions in perovskite ferroelectricity. Ferroelectrics. 2006;338(1):73–79. [Google Scholar]
  • 17.Hang T, Zhang W, Ye HY, Xiong RG. Metal-organic complex ferroelectrics. Chem Soc Rev. 2011;40(7):3577–3598. doi: 10.1039/c0cs00226g. [DOI] [PubMed] [Google Scholar]
  • 18.Petersen L, Hedegård P. A simple tight-binding model of spin-orbit splitting of sp-derived surface states. Surf Sci. 2000;459(1-2):49–56. [Google Scholar]
  • 19.Mitzi DB. Synthesis, crystal structure, and optical and thermal properties of (C4H9NH3)2MI4 (M = Ge, Sn, Pb) Chem Mater. 1996;8(3):791–800. [Google Scholar]
  • 20.Tang Z, Guan J, Guloy AM. Synthesis and crystal structure of new organic-based layered perovskites with 2,2’-biimidazolium cations. J Mater Chem. 2001;11(2):479–482. [Google Scholar]
  • 21.Burschka J, et al. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature. 2013;499(7458):316–319. doi: 10.1038/nature12340. [DOI] [PubMed] [Google Scholar]
  • 22.Liu M, Johnston MB, Snaith HJ. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature. 2013;501(7467):395–398. doi: 10.1038/nature12509. [DOI] [PubMed] [Google Scholar]
  • 23.Stranks SD, et al. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science. 2013;342(6156):341–344. doi: 10.1126/science.1243982. [DOI] [PubMed] [Google Scholar]
  • 24.Xing G, et al. Long-range balanced electron- and hole-transport lengths in organic-inorganic CH3NH3PbI3. Science. 2013;342(6156):344–347. doi: 10.1126/science.1243167. [DOI] [PubMed] [Google Scholar]
  • 25.Kresse G, Hafner J. Ab initio molecular dynamics for liquid metals. Phys Rev B Condens Matter. 1993;47(1):558–561. doi: 10.1103/physrevb.47.558. [DOI] [PubMed] [Google Scholar]
  • 26.Kresse G, Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys Rev B Condens Matter. 1996;54(16):11169–11186. doi: 10.1103/physrevb.54.11169. [DOI] [PubMed] [Google Scholar]
  • 27.Blöchl PE. Projector augmented-wave method. Phys Rev B Condens Matter. 1994;50(24):17953–17979. doi: 10.1103/physrevb.50.17953. [DOI] [PubMed] [Google Scholar]
  • 28.Perdew JP, et al. Restoring the density-gradient expansion for exchange in solids and surfaces. Phys Rev Lett. 2008;100(13):136406. doi: 10.1103/PhysRevLett.100.136406. [DOI] [PubMed] [Google Scholar]
  • 29.Koelling DD, Harmon BN. A technique for relativistic spin-polarised calculations. J Phys C Solid State Phys. 1977;10(16):3107. [Google Scholar]
  • 30.Marzari N, Vanderbilt D. Maximally localized generalized Wannier functions for composite energy bands. Phys Rev B. 1997;56(20):12847–12865. [Google Scholar]
  • 31.Souza I, Marzari N, Vanderbilt D. Maximally localized Wannier functions for entangled energy bands. Phys Rev B. 2001;65(3):035109. [Google Scholar]
  • 32.Mostofi AA, et al. Wannier90: A tool for obtaining maximally-localised Wannier functions. Comput Phys Commun. 2008;178(9):685–699. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES