Adding to the Perovskite Universe: Inverse-Hybrid ... - ACS Publications

Oct 18, 2017 - The electronic structure of this most stable corner- and edge- sharing structure is shown in Figure 3e,f. Gold B-site contributions are...
0 downloads 9 Views 3MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Letter

Adding to the Perovskite Universe - Inverse Hybrid Perovskites Julian Gebhardt, and Andrew M. Rappe ACS Energy Lett., Just Accepted Manuscript • DOI: 10.1021/acsenergylett.7b00966 • Publication Date (Web): 18 Oct 2017 Downloaded from http://pubs.acs.org on October 19, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Energy Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Adding to the Perovskite Universe - Inverse Hybrid Perovskites Julian Gebhardt∗ and Andrew M. Rappe Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104-6323, United States E-mail: [email protected]

1 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The perovskites are a rich family of functional materials with many interesting physical properties. Usually, they contain two cationic species on the A- and B-sites, surrounded by anionic species on the X-site, but compounds are also known that invert the ion types on the respective lattice sites (inverse perovskites). Recently, conventional perovskites with one inorganic cation substituted by an organic molecule are intensively studied, due to the promising performance of CH3 NH3 PbI3 based solar-cells. Here, for the first time, we take the concept of inverse perovskites to organic-inorganic hybrid materials, investigating the properties of inverse-hybrid perovskites by first-principles calculations, adding yet another structural variant to the perovskite universe. We present results for various compositions with a wide range of band gaps from metallic systems over small and intermediate band gaps to large band-gap semiconductors. Due to the changed location of the organic ion, the inverse structure could overcome stability problems of current hybrid perovskite photovoltaics. In addition, inverse hybrid perovskites show inherent off-center displacement of ions, leading to polar phases with large polarization.

Graphical TOC Entry

2 ACS Paragon Plus Environment

Page 2 of 15

Page 3 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Perovskite materials exhibit fascinating properties for various applications, and a wide variety of different compositions have been investigated. The perovskite ABX 3 structure contains cations on the A- and B- sites, with anions on the X-site balancing the positive charges, leading to a crystalline solid with strong ionic bonding. Besides the numerous inorganic perovskites, over the last decade there has been resurgent interest in perovskites with organic molecules, forming so-called organic-inorganic hybrid perovskites (HP). The organic part is usually employed on the A-site, 1,2 forming materials with promising photophysical properties for photovoltaic applications, but can also be found on the X-site or a combination of both. 3,4 However, it is also possible to invert the ionic charges, forming structures with anions on the A- and B-sites and cations on the X-site. Although less frequently investigated than perovskites, such inverse or anti perovskites (IP) have been studied since 1980, 5 showing many fascinating physical properties. 6–12 Herein, we combine the two concepts of hybrid and inverse perovskites, proposing a new class of inverse-hybrid perovskites (IHP). We present results for a variety of different X 3 BA − 2− compounds combining X=CH3 NH+ 3 (MA) with monovalent (a ) and divalent anions (a )

by accurate density-functional calculations using Quantum Espresso 13 (for details on the computational setup see the supporting information (SI)). We demonstrate the flexibility of these compounds as well as their limitations to form the perovskite structure, estimate their formation energies, and show that the compositional variety leads to materials with metallic and semiconducting properties, with band gaps tunable from very small values due to SOC, over an intermediate range interesting for photovoltaic applications, up to wide band gap semiconductors. The structural alignment naturally includes a B-site off-center displacement, giving rise to polar structures. Traditionally, the Goldschmidt tolerance factor 14 is used as a key indicator for the preference of ABX3 compositions to crystallize as perovskite. This concept has also been extended to organic-molecules in HP. ? We can further extend this concept for IPs and IHPs. However, we must keep in mind that anionic radii are less established than cationic radii and

3 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

that the tolerance factor is only an indicator, 15 with acceptable values ranging typically from t = 0.75 − 1.0. Common X-site anions in perovskites have ionic radii of 1.26 ˚ A to 2.06 ˚ A (ionic radii between O2- and I- ). 16 In order to be suitable to build perovskites with inorganic anions, we choose a small cation within this range for the X-site. Due to its favorable character for HPs and the current intense focus we select MA. When calculating tolerance factors for structures that contain organic ions complications arise, because their effective ionic radii are less established and they are non-spherical. Since the organic cation in IHPs is located on the X-site and, thus, enters the calculation of t in numerator and denominator, fortunately t is quite insensitive to the exact value of the effective ionic radius of the organic cation (see Table SI1) and we can, therefore, use the proposed value of reff (MA) = 2.17 ˚ A with confidence. ? Radii for untabulated anionic oxidation states were obtained similarly to Shannon’s original approach, see Table SI2. We find that the resulting values for t are a good tool to estimate suitable ion compositions, which we will use below. To pair with a monovalent cation for the X-site, we are seeking B -/2- and A2-/- ions. On the one hand, a2− ions should generally be larger than a− anions and should favor the A-site. On the other hand, the B-site provides a closer proximity to the cations, which should be favorable for the anion in the higher oxidation state due to the greater Coulombic attraction. Thus, in contrast to conventional perovskites where these effects go hand in hand, they are competing in the case of IPs, and the preference of A- vs. B- site occupation must be investigated. Here, we focus on the most stable compounds of the compositions with the most interesting properties, whereas details on our materials search will be discussed elsewhere. Choosing A=I- as large anion provides structural flexibility, encouraging a search for suitable B 2− anions. Natural candidates are elements of the chalcogen group. For (MA)3 BI with B=O, S, Se, Te, all compounds have tolerance factors within the range of 0.71 – 1.0 and could, therefore, form stable IHPs, with Te and Se compounds on the fringe where other

4 ACS Paragon Plus Environment

Page 4 of 15

Page 5 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

competing phases become more likely. 14 We find that O and S become hydrogenated, and we focus, here, on the heavier chalcogenides so that this kind of hydrogenation is suppressed, due to the reduced electronegativity and reduced orbital overlap of larger anions with hydrogen. For (MA)3 SeI we find a structure (Fig. 1 (a, b)) with both corner- and edge-sharing octahedra (CaIrO3 structure), which is energetically preferred by 0.32 eV per formula unit (f.u.) compared to the cubic perovskite structure. In fact we find that this CaIrO3 phase is favored for all cases where the tolerance factor falls below a threshold of 0.76. In the CaIrO3 structure, the large B–X bonding distance is better accommodated, due to a stretching of bonds along one axis and a rotation of the BX2 plane by about 45◦ perpendicular to this axis. As a result, the H–N–H angle is perfectly suited to bridge two B-sites by a NH3 group. Along the direction of the BX2 plane that does not contain these NH3 bridges, an X-site double layer with only intermolecular interactions between CH3 units occurs. This leads to an elongated lattice in this direction analogous to layered perovskites in the Ruddlesden-Popper phase. 17–19 As such, this structural variant could host even larger organic groups. The electronic structure of (MA)3 SeI is shown in Fig. 2 (a, b). For inverse materials, it can be expected that the electronic structure can be deduced from traditional perovskites, by inverting the contributions at the valence band (VB) and conduction band (CB) edges, which are dominated in conventional perovskites by X-site anions and B-site cations, respectively. Despite the phase change to CaIrO3 , this inverted band structure is observed for (MA)3 SeI, with X-site cation contributions in the CB region and contributions from the B- and A-site anions in the VB region. In order to estimate the stability of our new materials towards decomposition, we compute the formation energy for the reaction (MA)2 a2− + (MA)a− −→ (MA)3 a2− a− , which corresponds to the synthetical route often used for HPs 20 (Table 1, see Figs. S1 and S2 for details). With respect to this reaction, (MA)3 SeI is predicted to be thermodynamically unstable by 0.28 eV/f.u.

5 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The ionic radius for Te2- is essentially identical to that of I- . This means that the site preference is unclear and that competing structural phase alternatives to the IHP become more likely. A-site Te2- and, thus, the perovskite structure should be stabilized by choosing a smaller B-site anion. Therefore, we focus on the fluoride compound here, which, gives the best tolerance factor for any halide-telluride compound. Indeed, (MA)3 FTe stabilizes the perovskite structure (Fig. 1 (c, d)), which is 1 eV/f.u. more stable than the CaIrO3 structure obtained for (MA)3 TeI and overall thermodynamically stable (formation energy of −0.71 eV/f.u.). In the obtained perovskite structure, MA units on the X-sites are tilted, allowing every molecule to bridge B-sites using both CH3 and NH3 groups via hydrogen bonds to the CH3 and the NH3 groups. The location change of the organic MA, from the relatively free A-site in HPs to the bridging X-site in IHPs, could potentially overcome many of the stability problems 21 of current HPs. Three MA units are oriented in order to point one of the acidic N–H hydrogen atoms towards the B-site center. The remaining three corners of the BX6 octahedron are occupied with CH3 groups, each pointing one H atom towards the B-site. The electronic structure (Fig. 2 (c, d)) is similar to (MA)3 SeI, but the band gap reduces from 3.39 eV ((MA)3 SeI) to 3.15 eV ((MA)3 TeI). To further demonstrate the stability of (MA)3 FTe, we compute the phonon band structure (Fig. S3), which does not show any instability throughout the Brillouin zone. For (MA)3 FTe, we also compare the PBE band gap with accurate hybrid-DFT and manybody calculations in order to provide a better quantitative estimate for these wide band gap semiconductors. Other than the expected underestimation of the band gap, PBE describes the bands qualitatively correctly compared to the HSE band structure (Fig. 2 (c)). Note that we obtain a significant difference between the HSE gap (Eg = 4.29 eV) and the G0 W0 quasiparticle gap (Eg = 5.88 eV). Such discrepancies for large-gap systems have been observed before, 22–27 i.e., pinpointing the band gap exactly remains difficult. However, the observed agreement between HSE and PBE band dispersion suggests that band structures can be interpreted by applying a simple scissors operator 28 for comparison with future experimental

6 ACS Paragon Plus Environment

Page 6 of 15

Page 7 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

data.

Figure 1: Representative structures of IHPs. Atoms are colored gray (C), blue (N), white (H), green (halogens), and brown (chalcogens). (a, b) show an alternative phase (CaIrO3 ) observed for cases with t ≤ 0.76. B–X–B bridges are observed by H–N–H bonds along two directions and an X-site double layer along the third. (c, d) show (MA)3 FTe as exemplary IHP, with B–X–B-sites being bridged in a H–C–N–H fashion via (tilted) MA molecules.

Table 1: Tolerance factor t, structure information, formation energies Eform , and band gap Eg for the investigated (MA)3 BA compounds. For cases with indirect band gap, the direct band gap Egdir is given in parentheses. P=perovskite System (MA)3 SeI (MA)3 FTe (MA)3 FTl (MA)3 FPb (MA)3 AuTe

3

t Structure V /atom/˚ A 0.75 CaIrO3 9.175 0.89 P 8.836 0.90 P 8.646 0.91 P 8.633 0.76 CaIrO3 ’ 8.671

a0 /˚ A 7.030 5.996 5.839 5.855 6.332

b0 /˚ A 6.931 6.460 6.481 6.453 6.274

c0 /˚ A 5.232 5.932 5.951 5.944 5.916

α 94.75◦ 91.28◦ 92.77◦ 91.76◦ 85.47◦

β 71.89◦ 90.28◦ 91.51◦ 90.76◦ 77.22◦

γ 99.88◦ 89.79◦ 91.23◦ 90.38◦ 79.90◦

Eg (Egdir )/eV 3.39 3.15 (3.38) metallic 0.07 (0.10) 2.35 (2.42)

Eform /eV/f.u. 0.28 -0.71 0.90 -0.58 0.21

In addition to lattice distortion from the cubic phase (Table 1), IHPs naturally show pronounced B-site off-center displacements due to the uneven binding strength of B-site anions to CH3 and NH3 units. Therefore, we expect IHPs to be polar and calculate the polarization for our model case (MA)3 FTe. We obtain a large total polarization of P = 44.18 µC/cm2 [P = (−9.2, 35.6, −24.5)T µC/cm2 ], which demonstrates the polar character of this new material and makes the study of ferroelectric IHPs an interesting future direction. 7 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: Band structure and DoS for (MA)3 SeI (a, b), and (MA)3 FTe (c, d). Bands are colored according to elemental species (Fig. 1, and H in pink for visibility). VB and CB regions are dominated by anion and cation contributions, respectively. The relative alignment of bands from the anion sites changes with the ion composition, allowing to tune the gap of these wide band-gap semiconductors. For a better quantitative estimate of the band structure we compare the PBE result with HSE (black) (c). Since chalcogenide halide IHPs feature large band gaps, we also tested substitution of halogens and chalcogens against other mono and divalent metal anions. For example, group 13 elements like Tl could potentially stabilize a 2- oxidation state by half filling their valence p shell. The ionic radii of Tl2- and Te2- are similar and, therefore, the best suitable halogen to stabilize perovskite is again F on the B-site. The obtained electronic structure for (MA)3 FTl differs greatly from structures with chalcogens. Note that the partially-filled valence p shell requires a spin-polarized treatment. Calculations including Tl (and Pb below) were carried out fully relativistically including spin-orbit coupling (SOC). We obtain a metallic band structure [Fig. 3 (a, b)], with Tl 6p states around the Fermi level. SOC leads to a splitting of these bands into two j = 1/2 and four j = 3/2 bands. The lower pair of j = 3/2 bands is fractionally occupied by about one electron per f.u., in line with the formal charge of Tl2- . Due to SOC and inversion symmetry breaking, doubly-degenerate Kramers-theorem enforced Weyl points 29 are found at Γ and each of the sampled high-symmetry points for

8 ACS Paragon Plus Environment

Page 8 of 15

Page 9 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

this semimetal. In the primitive unit cell we obtain a non-magnetic structure, but further stabilization of antiferromagnetic spin structures could occur in larger cells. In its present structure, (MA)3 FTl has a high formation energy of 0.9 eV/f.u., i.e., fabrication might require metastable phase engineering. 30 We further explore this concept and try to open a small gap by adding another electron, i.e., we substitute Tl by Pb. Once again, the tolerance factor is optimal with fluorine on the B-site. The band topology of (MA)3 FPb (Fig. 3 (c, d)) is identical to the Tl case, but due to the additional electron a small gap of 0.07 eV is successfully opened between the four j = 3/2 bands. In contrast to Tl, Pb leads to a thermodynamically favorable structure (−0.58 eV/f.u. fromation energy), demonstrating the feasibility of small band-gap IHPs.

Figure 3: Band structure and DoS for (MA)3 FTl (a, b), (MA)3 FPb (c, d), and (MA)3 AuTe (e, f). The insets in (a) and (c) show a close-up of the bands around the Fermi level (dashed gray line). While three valence p shell electrons lead to a fractional filling of the SOC split p bands for tellurium, in the lead case a small band gap appears between the two sets of j = 3/2 bands. For (MA)3 AuTe, the B-site anion (Au1- ) contributes to bands in both the VB and the CB region, leading to a smaller band gap compared to chalcogenide halide IHPs. 9 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Finally, we investigate the possibility to employ transition-metal anions, with the idea being to fill the valence d shell in order to form a1− or a2− anions, by employing group 10 or 11 elements. The ionic radii of these transition metals are of intermediate size and a perfect candidate that favors the perovskite structure is hard to find. Nevertheless, we show by the example of Au1- that this route is promising in order to obtain moderate band gaps. Based on ionic radii, Au1- favors the B-site in conjunction with Te (t = 0.76), whereas Se is even less suitable, giving almost identical tolerance factors of about 0.70 on either site. The small tolerance factor of (MA)3 AuTe leads to a strong distortion of the Au(MA)6 octahedrons in the CaIrO3 phase, which is more stable than an undistorted structure, the perovskite phase, or compounds with exchanged ion positions. The electronic structure of this most stable corner- and edge-sharing structure is shown in Fig. 3 (e, f). Gold B-site contributions are observed at both the VB and the CB edge. These contributions arise from the filled valence 6s and 5d orbitals as well as the empty 6p orbitals, confirming the formation of Au1- . With gold anions contributing to CB and VB, the band gap is significantly reduced compared to chalcogenide halide IHPs to 2.35 eV. However, due to the intermediate ionic radii, promising compositions that favor a perovskite structure are not easy to find and require further tuning. We introduce IHPs as a new materials class. Tolerance factors are used to guide our compositional search, since it provides a valuable estimate for the stability of the perovskite structure, with a threshold value of t ≥ 0.76. For smaller values, a mixed corner- and edge-sharing phase occurs, which can be further explored in conjunction with larger organic X-site cations. For all halogen compounds, pairing heavier elements with small halogens, i.e., exchanging anion sites, successfully stabilizes the perovskite structure. Chalcogenide halides form large band-gap semiconductors, with a band gap tunable by the ion composition. Electronically, the expected inversion of band character compared to conventional perovskites is observed. By changing the ion composition, a variety of electronic behavior is observed, ranging from metallic systems, over systems with very small band gaps due to SOC, to

10 ACS Paragon Plus Environment

Page 10 of 15

Page 11 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

semiconductors. The inclusion of transition metals seems to be a promising route for systems with photovoltaic potential, whereas others show interesting topological behavior. We predict that the suggested compounds can be thermodynamically stable against decomposition and demonstrate the absence of soft phonons for one case [(MA)3 FTe]. IHPs show great variety with respect to ion composition as well as electronic properties. Furthermore, IHPs have technologically interesting polar phases with large polarization, and we hope that our study triggers research efforts to further explore this new modification of functional perovskite materials. Due to the changed bonding site of the organic anion, the proposed IHPs could also be superior in terms of stability compared to traditional HPs.

Supporting Information Available Details on the computational methods, ionic radii, and employed reference states for the evaluation of formation energies. Phonon band structure for (MA)3 FTe.

Acknowledgement This work was supported by the U.S. Office of Naval Research, under Grant N00014-17-12574. J. G. thanks the German Research Foundation for support from Research Fellowship GE 2827/1-1 and Heng Gao for discussing topological physics. Computational support is provided by the HPCMO of the U.S. DOD.

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. (2) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. I. Chemical Management for 11 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Colorful, Efficient, and Stable Inorganic Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764–1769. (3) Zhang, W.; Cai, Y.; Xiong, R.-G.; Yoshikawa, H.; Awaga, K. Exceptional Dielectric Phase Transitions in a Perovskite-type Cage Compound. Angew. Chem. Int. Ed. 2010, 49, 6608–6610. (4) Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R. H.; Cheetham, A. K. Chemically Diverse and Multifunctional Hybrid OrganicInorganic Perovskites. Nat. Rev. Mater. 2017, 2, 1–18. ¨ (5) Widera, A.; Sch¨afer, H. Ubergangsformen Zwischen Zintlphasen und Echten Salzen: Die Verbindungen A3 BO (MIT A = Ca, Sr, Ba und B = Sn, Pb). Mater. Res. Bull. 1980, 15, 1805–1809. (6) He, T.; Huang, Q.; Ramirez, A. P.; Wang, Y.; Regan, K. A.; Rogado, N.; Hayward, M. A.; Haas, M. K.; Slusky, J. S.; Inumara, K. et al. Superconductivity in the Non-Oxide Perovskite MgCNi3 . Nature 2001, 411, 54–56. (7) Takenaka, K.; Takagi, H. Giant Negative Thermal Expansion in Ge-Doped AntiPerovskite Manganese Nitrides. Appl. Phys. Lett. 2005, 87, 261902. (8) Kamishima, K.; Goto, T.; Nakagawa, H.; Miura, N.; Ohashi, M.; Mori, N.; Sasaki, T.; Kanomata, T. Giant Magnetoresistance in the Intermetallic Compound Mn3 GaC. Phys. Rev. B 2000, 63, 024426. (9) G¨abler, F.; Schnelle, W.; Senyshyn, A.; Niewa, R. Magnetic Structure of the Inverse Perovskite (Ce3 N)In. Solid State Sci. 2008, 10, 1910–1915. (10) Wang, B. S.; Tong, P.; Sun, Y. P.; Tang, W.; Li, L. J.; Zhu, X. B.; Yang, Z. R.; Song, W. H. Structural, Magnetic Properties and Magnetocaloric Effect in Ni-Doped

12 ACS Paragon Plus Environment

Page 12 of 15

Page 13 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Antiperovskite Compounds GaCMn3−x Nix (0 ≤ x ≤ 0.10). Physica B 2010, 405, 2427– 2430. (11) Gomonaj, E. V.; L’Vov, V. A. A Theory of Spin Reorientation and Piezomagnetic Effect in Noncollinear Mn3 AgN Antiferromagnet. Phase Transitions 1992, 40, 225–237. (12) Sun, Y.; Chen, X.-Q.; Yunoki, S.; Li, D.; Li, Y. New Family of Three-Dimensional Topological Insulators with Antiperovskite Structure. Phys. Rev. Lett. 2010, 105, 216406. (13) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; 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. (14) Goldschmidt, V. M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 21, 477–485. (15) Travis, W.; Glover, E. N. K.; Bronstein, H.; Scanlon, D. O.; Palgrave, R. G. On the Application of the Tolerance Factor to Inorganic and Hybrid Halide Perovskites: a Revised System. Chem. Sci. 2016, 7, 4548–4556. (16) Shannon, R. D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. (17) Ruddlesden, S. N.; Popper, P. New Compounds of the K2 NiF4 Type. Acta Cryst. 1957, 10, 538–539. (18) Calabrese, J.; Jones, N. L.; Harlow, R. L.; Herron, N.; Thorn, D. L.; Wang, Y. Preparation and Characterization of Layered Lead Halide Compounds. J. Am. Chem. Soc. 1991, 113, 2328–2330. (19) Gebhardt, J.; Kim, Y.; Rappe, A. M. Influence of the Dimensionality and Organic

13 ACS Paragon Plus Environment

ACS Energy Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Cation on Crystal and Electronic Structure of Organometalic Halide Perovskites. J. Phys. Chem. C 2017, 121, 6569–6574. (20) Im, J.-H.; Lee, C.-R.; Lee, J.-W.; Park, S.-W.; Park, N.-G. 6.5% Efficient Perovskite Quantum-Dot-Sensitized Solar Cell. Nanoscale 2011, 3, 4088–4093. (21) Niu, G.; Guo, X.; Wang, L. Review of Recent Progress in Chemical Stability of Perovskite Solar Cells. J. Mater. Chem. A 2015, 3, 8970–8980. (22) Karlick´ y, F.; Zboril, R.; Otyepka, M. Band Gaps and Structural Properties of Graphene Halides and Their Derivates: A Hybrid Functional Study with Localized Orbital Basis Sets. J. Chem. Phys. 2012, 137, 034709. (23) L¨ uftner, D.; Refaely-Abramson, S.; Pachler, M.; Resel, R.; Ramsey, M. G.; Kronik, L.; Puschnig, P. Experimental and Theoretical Electronic Structure of Quinacridone. Phys. Rev. B 2014, 90, 075204. (24) Camargo-Martinez, J. A.; Baquero, R. The Band Gap Problem: The Accuracy of the Wien2k Code Confronted. Rev. Mex. Fis. 2013, 59, 453–459. (25) Fuchs, F.; Furthm¨ uller, J.; Bechstedt, F.; Shishkin, M.; Kresse, G. Quasiparticle Band Structure Based on a Generalized Kohn-Sham Scheme. Phys. Rev. B 2007, 76, 115109. (26) Tran, F.; Blaha, P. Accurate Band Gaps of Semiconductors and Insulators with a Semilocal Exchange-Correlation Potential. Phys. Rev. Lett. 2009, 102, 226401. (27) Crowley, J. M.; Tahir-Kheli, J.; Goddard III, W. A. Resolution of the Band Gap Prediction Problem for Materials Design. J. Phys. Chem. Lett. 2016, 7, 1198–1203. (28) Levine, Z. H.; Allan, D. C. Linear Optical Response in Silicon and Germanium Including Self-Energy Effects. Phys. Rev. Lett. 1989, 63, 1719–1722. (29) Chang, G.; Sanchez, D. S.; Wieder, B. J.; Xu, S.-Y.; Schindler, F.; Belopolski, I.; Huang, S.-M.; Singh, B.; Wu, D.; Neupert, T. et al. Kramers Theorem-Enforced Weyl 14 ACS Paragon Plus Environment

Page 14 of 15

Page 15 of 15

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Energy Letters

Fermions: Theory and Materials Predictions (Ag3 BO3 , TlTe2 O6 and Ag2 Se Related Families). arXiv:1611.07925 2016, 1–26. (30) Gopalakrishnan, J. Chimie Douce Approaches to the Synthesis of Metastable Oxide Material. Chem. Mater. 1995, 7, 1265–1275.

15 ACS Paragon Plus Environment