Materials Design of Green Light-emitting Semiconductors: Perovskite

2 hours ago - Both the undoped and doped SrHfS3 exhibited an intense green photoluminescence (PL) at 2.37 eV. From the PL blue shift and short lifetim...
0 downloads 0 Views 902KB Size
Subscriber access provided by Washington University | Libraries

Article

Materials Design of Green Light-emitting Semiconductors: Perovskite-type Sulfide SrHfS

3

Kota Hanzawa, Soshi Iimura, Hidenori Hiramatsu, and Hideo Hosono J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 06 Mar 2019 Downloaded from http://pubs.acs.org on March 6, 2019

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 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 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.

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 34 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

Journal of the American Chemical Society

ACS Paragon Plus Environment

Journal of the American Chemical Society 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 Paragon Plus Environment

Page 2 of 34

Page 3 of 34 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

Journal of the American Chemical Society

ACS Paragon Plus Environment

Journal of the American Chemical Society 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 Paragon Plus Environment

Page 4 of 34

Page 5 of 34 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

Journal of the American Chemical Society

ACS Paragon Plus Environment

Journal of the American Chemical Society 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 Paragon Plus Environment

Page 6 of 34

Page 7 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Materials Design of Green Light-emitting Semiconductors: Perovskite-type Sulfide SrHfS3

Kota Hanzawa,1 Soshi Iimura,1 Hidenori Hiramatsu,1,2,* and Hideo Hosono1,2

1Laboratory

for Materials and Structures, Institute of Innovative Research, Tokyo

Institute of Technology, Mailbox R3-3, 4259 Nagatsuta-cho, Midori-ku, Yokohama 226-8503, Japan 2Materials

Research Center for Element Strategy, Tokyo Institute of Technology,

Mailbox SE-1, 4259 Nagatsuta-cho, Midori-ku, Yokohama 226-8503, Japan

1

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 8 of 34

K. Hanzawa et al.,

Abstract A current issue facing light-emitting devices is a missing suitable material for green emission. To overcome this, we explore semiconductors possessing (i) a deep conduction band minimum (CBM) and shallow valence band maximum (VBM), (ii) a good controllability of electronic conductivity and carrier polarity, and (iii) a direct-allowed band gap corresponding to a green emission. We focus on early transition metal (eTM)-based perovskites. The eTM cation’s high and stable valence state makes its carrier controllability easy, and the eTM’s non-bonding d-orbital and the anion’s p-orbital, which constitute the deep CBM and shallow VBM, are favourable for n- and p-type doping, respectively. To obtain a direct band gap, we applied a scheme that folds the bands constituting the VBM at the zone boundary to the zone centre where the CBM appears. Orthorhombic SrHfS3 was chosen as the candidate. Electrical conductivity was tuned from 6×10−7 S·cm−1 to 7×10−1 S·cm−1 with lanthanum (La)-doping and to 2×10−4 S·cm−1 with phosphorus (P)-doping. Simultaneously, the major carrier polarity was controlled into n-type by La-doping, and into p-type by P-doping. Both the undoped and doped SrHfS3 exhibited an intense green photoluminescence (PL) at 2.37 eV. From the PL blue shift and short lifetime, we attributed the emission to a band-to-band transition and/or exciton. These results demonstrate that SrHfS3 is a promising green light-emitting semiconductor.

2

ACS Paragon Plus Environment

Page 9 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Introduction Light-emitting semiconductors are key components in practical optoelectronic devices such as light-emitting diodes, laser diodes, and solar cells. Although light-emitting technologies have widely been applicable, there is a strong demand on next-generation materials to exhibit higher performance in terms of light brightness, quantum efficiency, and colour accuracy1. Currently, GaN- and GaAs-based IIIb-Vb semiconductors have commercially been employed in light-emitting devices. However, they have a serious issue where the emission quantum efficiency drastically decreases in a particular wavelength region around green and yellow, that is, it decreases from ~70%–80% in the red- and blue-light regions to ~15% in the green region (~530 nm) or ~20% in the yellow region (~570 nm)2, 3. This is called the ‘green gap problem’ and exists because there is currently no green-light source bright enough to realize the next-generation optoelectronic devices. Therefore, it is strongly required to explore new semiconductor materials with a high quantum efficiency of emission in the green wavelength region (Eg = 2.18−2.50 eV, λ = 495−570 nm) and n- and p-type dopability. To meet these demands, we consider three basic electronic characteristics of semiconductors; (i) a deep conduction band minimum (CBM; large electron affinity) and shallow valence band maximum (VBM; small ionization potential) to achieve both n- and p-type dopings, respectively, (ii) controllability of electronic conductivity and carrier polarity, and (iii) a direct allowed band gap to realize highly efficient light emission. Recently, our research group proposed that early transition metal (eTM)-based semiconductors with a high symmetry crystal structure can satisfy both the first and second requirements4−6. Usually, the empty d-orbital of the eTM and p-orbital of the 3

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 10 of 34

K. Hanzawa et al.,

anion form a strong antibonding state, so that the conduction band (CB) of the eTM-based compound possesses a too shallow energy level (i.e., too small electron affinity) for electron doping (Fig. 1a). However, by introducing a deep non-bonding state of the d-orbital of eTM, we demonstrated that a deep CBM is formed at a specific k point, which enables doping with an electron and reduces the effective mass. Figure 1b shows an example of the d-orbital configuration in cubic perovskite at the Γ point (k = 0). At this point, the phase of the t2g orbitals of transition metals (e.g., eTM in A-eTM-X3, where A and X denote a cation and an anion, respectively) is invariant from one unit cell to the next. This configuration of the d-orbital can form neither a bonding nor anti-bonding state with the neighbouring p-orbitals of anions owing to their symmetry constraint. As a result, the deep unoccupied non-bonding state consists of the CBM. The use of eTMs has advantages in terms of controllability of the electronic conductivity and carrier polarity. The eTM cations tend to occupy a high coordination number site owing to their large valence states, such as Zr4+, which suppresses the formation of a cation vacancy. In addition, the stable valence state inhibits the localization of the generated electron at the eTM cation site through the change of valence state. These characteristics are in contrast to those of post-transition metal (pTM, e.g., Cu, In, and Sn) cations, which are constituents of conventional oxide semiconductors. These semiconductors suffer from a poor doping controllability owing to their low and variable valence changes of the cation7−9. Therefore, eTM-based cubic perovskites A-eTM-X3 seem to be promising candidates in terms of the dopability and controllability of the carrier density. A unique feature of the cubic perovskite is that their CBM and their VBM are composed of 4

ACS Paragon Plus Environment

Page 11 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

non-bonding states. Figure 1c depicts the orbital configuration of a fully electron-occupied p-orbital of anion X at the R point. At the R point of the primitive cubic structure, the phase of each orbital in the unit cell is inverted by the primitive translation along the x, y, or z directions. This configuration of p orbitals can form neither a bonding nor anti-bonding state with the neighbouring d-orbitals of eTM. Consequently, this shallow energy level of the non-bonding state pushes up the VBM and makes it easy to dope the hole carriers. Although non-bonding deep d- and shallow p-orbitals levels are favourable for doping of electrons and holes, respectively, the indirect band gap in the cubic perovskite definitely inhibits an efficient light emission (Fig. 1d). Thus, we introduce another strategy, that is, band folding, to obtain a direct band gap suitable for light emission. Figure 1e and 1f show the relationship between the size of the unit cell in real space and the Brillouin zone (BZ) in reciprocal space10, 11. Here, we consider how the band structure of a one-dimensional s-orbital chain changes by choosing the unit cell containing one orbital or two orbitals. In the case of a one orbital cell with a lattice parameter of a, a single s-band appears in the first BZ ranging from −π/a to π/a. By doubling the unit cell size to a’ = 2a, the first BZ halves. Consequently, bands located outside the first BZ ranging from −π/2a to π/2a (i.e., located in the second BZ) are folded back to the first BZ and the number of bands doubles. This relation tells us that the VBM at the R point of cubic perovskite can be transferred to the Γ point by choosing a proper superstructure. Note that such band folding we suppose occurs only when the number of superstructures is an even multiple of a primitive unit cell. In a perovskite-type structure, a superstructure is introduced by tilting of the eTM-X6 octahedra. Indeed, it is known that octahedron tilting and the resultant 5

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 12 of 34

K. Hanzawa et al.,

superstructure are well governed by the tolerance factor12, therefore we can find an optimum electronic structure with the direct-type transition by choosing constituent elements with different ionic radii. In this paper, we investigate the electronic and optical properties of ternary alkali earth metal (AE)-eTM-S systems with an orthorhombic structure whose direct band gap corresponds to a colour ranging from green to red. Through screening by first-principles calculations and experiments, we identify that SrHfS3 is a promising candidate for the green light emission among the AE-eTM-S systems. We then experimentally demonstrate that SrHfS3 is a novel n- and p-type dopable semiconductor exhibiting an intense green emission at room temperature.

6

ACS Paragon Plus Environment

Page 13 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Figure 1. Design of n- and p-type dopable semiconductors with direct band gaps in eTM-based perovskite-type compounds. (a) Schematic energy diagram of diatomic molecule. The configurations of non-bonding (b) d-orbitals at Γ point and (c) p-orbitals at R point in cubic perovskite-type structure. The former constitutes CBM, while the latter constitutes VBM. Red cloverleafs and green dumbbells represent d-orbitals of eTM and p-orbitals of X (anion), respectively. (d) Schematic band structure of cubic perovskite-type compounds. Black solid curves denote their bands at CBM and VBM. Red band at Γ point is a band transferred from R point after band folding. The inset illustrates the BZ of primitive cubic cell. (e) Two different unit cells in one-dimensional s-orbital chain containing one orbital (top) and two orbitals (bottom). (f) Band structures of one-orbital unit cell (top) and two-orbital unit cell (bottom). For the two-orbital model, bands in the second BZ are folded back to the first BZ of the one-orbital model.

7

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 14 of 34

K. Hanzawa et al.,

Results and discussion  Electronic structures of AE-eTM-S3 From screening based on our concepts above, we chose AE-eTM-S3 (AE = Ca, Sr or Ba, eTM = Zr or Hf) owing to their bandgaps suitable for green light emission. The crystal structure shown in Fig. 2a forms an orthorhombic perovskite-type structure with a space group of Pnma. Their lattice constructs a √2 × 2 × √2 superstructure of a cubic perovskite-type cell, so that bands at the R point of the cubic lattice are folded to the Γ point of the orthorhombic lattice. The relationship between formation of the superstructure and the band folding is explained in Fig. S1. Figure 2b and 2c respectively show the electron density maps at the CBM and VBM of SrHfS3 as an example. At the CBM, a d-orbital-like electron density appears at the Hf site, while at the VBM, a p-orbital-like population is also visible at the S site. The other orbital contributions are negligibly small both for the CBM and VBM, which indicates that the non-bonding state in cubic perovskite is preserved in this orthorhombic structure. Figure 2d shows the band structure of SrHfS3. It is obvious that SrHfS3 has a direct-type band gap at the Γ point owing to the band folding, as explained in Fig. S1. Band structures of the other AE-eTM-S3 compounds are displayed in Fig. S2. To check whether the optical transition from the CBM to VBM is symmetrically allowed or not, we calculated the corresponding matrix momentum elements (Fig. S3). The finite and large value at the Γ point indicates that the band-edge transition is allowed. The calculated effective masses of electrons in SrHfS3 were 0.40, 0.26, and 0.43 m0 for Γ–X, Γ–Y, and Γ–Z, respectively, while those for holes were 0.70, 0.19, and 0.61 m0, where m0 represents the effective mass of a free electron. For comparison, the calculated effective masses of other AE-eTM-S3 are summarized in Table S1. Figure 2e shows the 8

ACS Paragon Plus Environment

Page 15 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

density of states (DOS) of SrHfS3. The CB consists of Hf-5d orbitals and the Sr-4d is located at a higher energy than the CB of Hf-5d orbitals. Because Sr-4d orbitals do not hybridize with Hf-5d orbitals, the CBM is determined almost only by Hf-5d orbitals. The VBs are composed mainly of S-3p orbitals. These results indicates that the electronic structures of CBM (i.e., d orbitals of an eTM) and VBM (i.e., p orbitals of an anion) are as we expected. Figure 2f summarizes the band alignments of AE-eTM-S3 and other semiconductors that exhibits both n- and p-type conductivity. The energy levels of CBM and VBM for conventional semiconductors (i.e., Si and GaAs13, 14) are located within the typical doping limit region, as shown by the yellow area13. Note that the energy levels of CBM and VBM of AE-eTM-S3 are both located around the boundary or inside of the limit region, which suggests that n- and p-type dopings should be possible. The calculated band gaps of CaHfS3, SrHfS3, and BaHfS3 were 2.21, 2.18, and 1.90 eV, respectively, while those of CaZrS3, SrZrS3, and BaZrS3 were 1.95, 1.94, and 1.72 eV, respectively, which are close to the calculated values reported previously15. Results of the band gap calculations for other AE-eTM-S3 are also summarized in Table S1. Because the band gaps of AE-HfS3 are more suitable for a green emission than the Zr-system, we chose the Hf-system, and investigated the optoelectronic properties.

9

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 16 of 34

K. Hanzawa et al.,

Figure 2. Electronic structures of AE-eTM-S3. (a) Crystal structure of orthorhombic perovskite-type sulfide AE-eTM-S3. Blue, red, and green spheres correspond to AE, eTM, and sulfur ions, respectively. Charge density maps at (b) CBM and (c) VBM of SrHfS3 calculated using the Heyd–Scuseria–Ernzerhof (HSE) hybrid-functional. Vertical colour scale bar indicates the intensity of the charge density. (d) Band structure of SrHfS3 projected by maximally localized Wannier functions. (e) Density of states (DOS). Black, blue, red, and green lines correspond to total DOS and contributions from Sr 4d-, Hf 5d-, and S 3p-orbitals, respectively. (f) Band alignment of AE-eTM-S3 and three n- and p-type dopable semiconductors. Except for AE-eTM-S3, all data are experimental results taken from Refs 5, 13, 14. The yellow area denotes a conventional carrier dopable range located from −4 to −6 eV with respect to Evac13.

10

ACS Paragon Plus Environment

Page 17 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

 SrHfS3 with an appropriate band gap for green light emission To examine the optical and carrier transport properties of AE-HfS3, we first prepared the undoped polycrystalline samples. AE-HfS3 was previously prepared through the direct sulfurization of a mixture of AE-CO3 and HfO216. However, we employed a conventional solid-state reaction using binary sulfide precursors to obtain high-density bulk (relative density >80%), which are advantageous especially for reliable carrier transport property measurements. Figure 3a and 3b show the X-ray diffraction (XRD) patterns of SrHfS3 and BaHfS3. They indicate that SrHfS3 and BaHfS3 were successfully synthesized, which was crosschecked by chemical composition analysis using an electron micro-probe analyser. For synthesis of AE = Ca, however, we could not obtain the CaHfS3 phase by the same method. From the Rietveld analysis on the XRD patterns, we observed that major parts of the obtained bulk consisted of SrHfS3 (purity = 94.0 mol%) and BaHfS3 (79.1 mol%), and the minor impurities were HfO2 and AE-S for both samples. Insets in Fig. 3a and 3b show photographs of the SrHfS3 and BaHfS3 polycrystalline bulk. The sample colours were vivid yellow and deep orange for SrHfS3 and BaHfS3, respectively. Note that both polycrystalline bulk samples were quite stable in ambient air atmosphere despite that they are sulfides. We measured the diffuse reflectance at room temperature in the visible– near-infrared wavelength region to examine the optical band gap. The raw spectra of the diffuse reflectance and the Kubelka–Munk plots are displayed in Fig. S4a and S4b. Figure 3c summarizes the calculated and observed optical band gaps of AE-HfS3. Even though the calculated band gaps were slightly underestimated, they reproduced well the observed optical band gaps. The observed gaps of SrHfS3 and BaHfS3 were respectively 11

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 18 of 34

K. Hanzawa et al.,

2.32 (λ = 534 nm) and 2.06 eV (λ = 602 nm), which was consistent with the sample colours, as shown in the inset photographs of Fig. 3a and 3b. Because the band gaps of SrHfS3 and BaHfS3 respectively correspond to the green- and orange-light emission wavelengths, we focussed on SrHfS3 hereafter.

12

ACS Paragon Plus Environment

Page 19 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Figure 3. XRD patterns of (a) SrHfS3 and (b) BaHfS3. Each inset is a photo of the polycrystalline bulk. (c) Optical band gaps of AE-HfS3. Red circles and blue squares represent observed optical and calculated band gaps, respectively. Vertical colour bar represents colours of the absorbed light. The raw diffuse reflectance spectra are shown in Fig. S4. The dashed lines indicate the green-light region, which is the main target of this study. 13

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 20 of 34

K. Hanzawa et al.,

 Electronic transport properties of SrHfS3 Next, we performed electron and hole dopings of SrHfS3 to examine the carrier controllability (i.e., control of carrier polarity and concentration). Among the several candidates for appropriate dopants, we chose La3+ (1.16 Å17) to act as a donor at the Sr2+ (1.26 Å17)-site and P3− (2.12 Å18) as an acceptor at the S2− (1.84 Å18)-site, owing to their suitable ionic radii. The nominal concentrations of La were set to 0.5%, 1%, 2%, 3%, 4%, and 6%, while those of P-substitution were 1%, 2%, 2.5%, 3%, 4%, and 4.5%. Hereafter, the doping concentrations are defined as the nominal concentrations. Before carrier transport property measurements, we confirmed replacement of La and P at the Sr- and S-sites by XRD. Figure 4a shows the doping concentration dependence of the lattice volume (V). The change in each lattice parameter is summarized in Fig. S5a for a-, S5b for b-, and S5c for c-axes. In the case of La substitution, V gradually decreased from 461.43 Å3 (undoped) to 461.09 Å3 (6% La-doped) owing to the smaller ionic radius of La than Sr. However, for P substitution, V expanded to 461.79 Å3 (4.5% P-doped) because of the larger ionic radius of P than S. This structure variation suggests that La and P progressively occupied the Sr- and S-sites with an increase in the doping concentrations, respectively. Figure 4b shows the electrical conductivities at room temperature as a function of each doping concentration. Before the measurements, we confirmed the ohmic contacts using Au electrodes for La- and undoped SrHfS3, and Pt electrodes for P-doped SrHfS3. Although undoped SrHfS3 exhibited a very low conductivity of 6×10−7 S·cm−1, the conductivities sharply increased with both La- and P-substitutions. In the case of La, the conductivity reached 7×10−1 S·cm−1 at a La doping of 6%; while in the case of P, it reached 2×10−4 S·cm−1 at a P doping of 4%. 14

ACS Paragon Plus Environment

Page 21 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

We checked the carrier polarity by thermopower measurements for both Laand P-doped SrHfS3. The raw data, that is, the detected thermoelectromotive forces (ΔV) as a function of the difference in temperature (ΔT), are shown in Fig. S6. Probably owing to the high electrical resistance, we could not obtain a clear linear relation for 1% and 2% P-doped SrHfS3. Figure 4c shows the Seebeck coefficients obtained from the slopes in Fig. S6. For all La- and P-concentrations, negative and positive Seebeck coefficients were observed, respectively, which indicates that La- and P-substitutions effectively supplied electrons and holes into SrHfS3. The negative Seebeck coefficients varied from −450 up to −230 μV/K with an increase in La concentration from 0.5% to 6%, while the positive Seebeck coefficients varied from +880 to +60 μV/K with P concentrations from 2.5% to 4.5%, which implies that carrier concentration increased with an increase in doping levels. The above results of electrical conductivity and Seebeck coefficients represent that carriers were continuously and widely activated along with the change in the cell volumes and the doping concentrations. These electronic transport properties demonstrate the good controllability of the carrier polarity as well as the concentrations in SrHfS3 by intentional impurity doping.

15

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 22 of 34

K. Hanzawa et al.,

Figure 4. Electron- and hole-doping effects on SrHfS3. (a) Lattice volume (V), (b) electrical conductivity (σ), and (c) Seebeck coefficients (S) of La- (blue squares), P- (red circles), and undoped (black triangle) SrHfS3. Horizontal axis represents nominal doping concentrations of La and P. All these measurements were performed at room temperature. The V was calculated from each axis lattice parameter shown in Fig. S5. 16

ACS Paragon Plus Environment

Page 23 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

 Intense green luminescence characteristics in undoped and doped SrHfS3 Finally, we performed photoluminescence (PL) measurements to examine the optical properties. Figure 5a shows PL spectra of undoped SrHfS3 at 30–300 K. SrHfS3 exhibited intense green emission (peaking at E = 2.37 eV, λ = 523 nm at 300 K) at all temperatures measured, where no other emission peaks were detected. We confirmed good agreement of the emission photon energy at 300 K with the band gap energy estimated from diffuse reflectance (2.32 eV, see Figs. 3c and S4). The full width at half maxima were 0.18 and 0.25 eV at 30 and 300 K, respectively. Note that the green emission could be seen by the human eye even at room temperature (i.e., 300 K) as shown in the inset picture of Fig. 5a, which suggests that the quantum efficiency of emission should be high. To elucidate the origin of the emission, we investigated the variation in the peak energy at different temperatures and constructed the decay curve. Figure 5b shows the temperature dependence of the emission peak energy. The energy increased from 2.37 eV at 300 K to 2.40 eV at 30 K; that is, a blue shift was clearly observed and the shift was well reproduced using an empirical model for the band edge19. From the decay curves, we estimated the emission lifetime to be 7.6 ns at 300 K and 3.7 ns at 30 K (see Fig. S7), which are comparable or shorter than the detection limit of our PL measurement system (i.e., pulse width of the excitation laser was 7−8 ns). This result indicates that the actual life time of the emission was shorter than 7 ns. These results strongly support that this intense emission originates from a band-to-band transition and/or an exciton. Indeed, eTM-based compounds are more ionic than pTM-based compounds owing to the lower electron negativity of eTMs than pTMs. In addition, the electron effective mass of the d-orbital in eTMs is usually heavier than that of the s-orbital in pTMs. These intrinsic natures of eTM-based semiconductors would be 17

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 24 of 34

K. Hanzawa et al.,

related to the efficient and bright emission, because both the ionicity and heavy electron effective mass are essential to enhance the exciton binding energy. Furthermore, any other emissions in the band gap (e.g., donor-acceptor pair emission) were not detected, which suggests that the number of defects should be very low even in a polycrystalline sample, probably as a result of the stable valence state of eTMs as we expected. This result is consistent with the low electrical conductivity (6×10−7 S·cm−1) of the undoped SrHfS3 (Fig. 4b). We then investigated the PL behaviour of the doped samples to clarify their potential as electroluminescence materials. Figure 5c summarizes the room temperature PL spectra of doped (6% for La and 4% for P) and undoped SrHfS3. In both La- and P-doped SrHfS3, only the band-edge and/or exciton emissions, similar to the undoped SrHfS3, were observed even at the high doping concentrations. Additionally, at other doping concentrations, similar emission spectra were observed, as shown in Fig. S8a (La-doping) and S8b (P-doping). Moreover, in all the doped samples, the green emissions were detected by human eyes at room temperature. It is noteworthy that highly doped carriers of electrons or holes do not quench the emission. These features denote that SrHfS3 can effectively emit green light even under the carrier-doped states.

18

ACS Paragon Plus Environment

Page 25 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Figure 5. Optical emission properties of doped and undoped SrHfS3. (a) Photoluminescence (PL) spectra of undoped SrHfS3 observed at 30−300 K. The inset is a picture of undoped SrHfS3 excited at room temperature. (b) Temperature dependence of emission peak energies. Red line is a fitting result using the following equation; Eg = E0 – αT2/(T+β), where E0 is the Eg at 0 K and α and β denote material constants19. (c) PL spectra of La 6%- (blue), P 4.5%- (red), and un-doped (black) SrHfS3 at 300 K.

19

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 26 of 34

K. Hanzawa et al.,

 Conclusions We explored candidate compounds for green-light emitting electron- and hole-dopable semiconductors. To achieve good carrier dopability and controllability, we first focused on eTM-based cubic perovskites, in which the high and stable valence state of the eTM cation makes the carrier controllability easy, and the deep non-bonding d-orbitals of eTM and shallow p-orbitals of the anion enable electron- and hole-doping, respectively. To obtain the direct band gap in the perovskites, we applied the band folding scheme that folds bands at the zone boundary to the zone centre. From the screening using the first-principles calculations, we identified that AE-HfS3 has an optimum, allowed band gap for green emission and appropriate band alignment to nand p-type dopings. The experimentally estimated optical band gaps of synthesized SrHfS3 and BaHfS3 were 2.3 eV (λ = 534 nm) and 2.1 eV, which correspond to green and orange emissions, respectively. The electrical conductivities of SrHfS3 were sharply and widely changed from 6×10−7 S·cm−1 at 0% to 7×10−1 S·cm−1 at 6% La3+ doping and 2×10−4 S·cm−1 at 4% P3− doping. The Seebeck effect measurements confirmed that substituting La3+ at the Sr2+-site successfully introduced n-type carriers, while the P3− substitution at the S2−-site exhibits p-type conductivity. From the PL measurements, SrHfS3 exhibited an intense green PL at all observed temperatures (30−300 K), which originates from a band-to-band transition and/or an exciton. The PL could be seen by the human eye even at room temperature and under carrier-doped states. Furthermore, other emissions could not be detected other than the green emission even under the doped conditions, which suggests that SrHfS3 should be strong against defects. These electronic and optical properties reveal that SrHfS3 is a promising candidate for novel green light-emitting semiconductors. 20

ACS Paragon Plus Environment

Page 27 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Computational and Experimental Procedures  Theoretical calculations Electronic structures of AE-eTM-S3 were calculated by using the HSE hybrid-functional with the HSE06 Vienna ab initio simulation package (VASP) code20. The mesh sampling in the Brillouin zone and cutoff energy were set to 5×4×5 and 600 eV, respectively, for all AE-eTM-S3. The DOS of tetragonal ZrOS was also calculated as a normalizer for band alignments of AE-eTM-S3 with the HSE hybrid-functional, where the BZ sampling and cutoff energy were set to 6×6×4 and 600 eV, respectively. The positions of CBM and VBM of AE-eTM-S3 were normalized by adjusting the energy levels of the core S 3s-band to that of ZrOS, where the energy levels were determined by estimating the centre of gravity for the DOS peak of S 3s-states. Then, all the normalized levels were shifted in parallel by adjusting the calculated VBM level of ZrOS to that experimentally observed5. For SrHfS3, BaHfS3, and BaZrS3, maximally localized Wannier functions as projection functions were constructed from the d-bands of AE and eTM, and the p-bands of S by using Wannier90 code21; while for CaHfS3, CaZrS3, and SrZrS3, they were from the d-bands of eTM and p-bands of S. The band structure and DOS shown in Fig. 2c and 2d, respectively, were depicted using those Wannier functions. In Fig. S3, the band structures of AE-eTM-S3 calculated by VASP code is compared with those calculated by Wannier interpolation.  Sample preparation Undoped polycrystalline AE-HfS3 (AE = Sr and Ba) was synthesized by conventional solid-state reaction from mixed precursors of AE-S and HfS2. The mixtures were pelletized and heated at 1100 °C for 48 hours in Ar-sealed stainless-steel tubes. For electron and hole dopings, we employed La2S3 and SrP, respectively. The 21

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 28 of 34

K. Hanzawa et al.,

chemical reaction formulae are as follows; (1−x)SrS + 0.5xLa2S3 + HfS2 → Sr(1−x)LaxHfS3 + 0.5xS↑ and (1−y)SrS + ySrP + HfS2 → SrHfS(3−y)Py. All compounds used as precursors and dopants were prepared by ourselves. AE-S and HfS2 were synthesized by mixing ~1−2 mm2 dendritic pieces of AE metal (purity: 99.99%) or Hf metal (99.99%) scraped from a metal-rod with sulfur powder (99.9999%) and by heating at 750 °C and 950 °C, respectively. La2S3 was prepared from scraped La metal (99.9%) and sulfur at 950 °C, and SrP was synthesized from Sr metal and phosphorous (99.999%) at 750 °C.  Structure analysis The crystal structure was analysed from powder X-ray diffraction (XRD) using a Cu-Kα radiation source with a one-dimensional silicon strip detector. The structure parameters (e.g., lattice parameter) and amounts of impurity phases were evaluated by Rietveld refinement using TOPAS code (Bruker AXS, version 4.2)22. We crosschecked the obtained phase by chemical compositions analysis using an electron probe micro analyser with a wave dispersion detector, where AE-S and HfS2 were used as standard samples.  Optical properties Diffuse reflectance spectra were obtained in the visible–near-infrared wavelength region at room temperature. Then we performed the Kubelka–Munk transformation for the obtained spectra to estimate the optical band gaps. To examine the emission characteristics, PL spectra of undoped SrHfS3 were obtained in the temperature range of 30−300 K; whereas for La- and P-doped SrHfS3, the temperature was set to 300 K. As the excitation source, we employed a third-harmonic Nd-doped YAG laser (3ω-Nd:YAG, λ = 355 nm), whose pulse width was 7−8 ns. Decay curve was measured 22

ACS Paragon Plus Environment

Page 29 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

for evaluation of the emission lifetime with a gate delay between 0–20 ns.  Electronic transport properties We investigated the electrical conductivities by the four- and two-probe methods for La- and P-doped SrHfS3, respectively. For ohmic contacts, Au and Pt electrodes were deposited by DC sputtering on La- and P-doped SrHfS3. Then, thermopower measurements were performed to determine the carrier polarity, where we deposited a Pt electrode only on p-type SrHfS3.

23

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 30 of 34

K. Hanzawa et al.,

Author information Corresponding author *

[email protected]

Notes The authors declare no competing financial interests.

Acknowledgements This work was supported by the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) through the Element Strategy Initiative to Form Core Research Center. S. I. was also supported by the Japan Society for the Promotion of Science (JSPS) through the Grant-in-Aid for Early-Career Scientists (Grant No. 18K13499). H. Hi. was also supported by the JSPS through the Grants-in-Aid for Scientific Research (A) and (B) (Grant Nos. 17H01318 and 18H01700), and Support for Tokyotech Advanced Research (STAR).

Supporting Information Supporting Information is available free of charge via the Internet at (http://pubs.acs.org) for descriptions of band folding of this system, calculation results, structure analysis data, raw data of optical and electronic transport properties data.

24

ACS Paragon Plus Environment

Page 31 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

References 1.

Tsao, J. Y.; Crawford, M. H.; Coltrin, M. E.; Fischer, A. J.; Koleske, D. D.; Subramania, G. S.; Wang, G. T.; Wierer, J. J.; Karlicek Jr, R. F. Toward smart and ultra-efficient solid-state lighting. Adv. Opt. Mater. 2014, 2, 809–836.

2.

O’Donnell, K. P.; Auf der Maur, M.; Di Carlo, A.; Lorenz, K.; the SORBET consortium. It’s not easy being green: Strategies for all-nitrides, all-colour solid state lighting. Phys. Status Solidi-Rapid Res. Lett. 2012, 6, 49–52.

3.

Auf der Maur, M.; Pecchia, A.; Penazzi, G.; Rodrigues, W.; Di Carlo, A. Efficiency drop in green InGaN/GaN light emitting diodes: The role of random alloy fluctuations. Phys. Rev. Lett. 2016, 116, 027401.

4.

Mizoguchi, H.; Kamiya, T.; Matsuishi. S.; Hosono, H. A germanate transparent conductive oxide. Nat. Commun. 2011, 2, 470.

5.

Arai, T.; Iimura, S.; Kim, J.; Toda, Y.; Ueda, S.; Hosono, H. Chemical design and example of transparent bipolar semiconductors. J. Am. Chem. Soc. 2017, 139, 17175–17180.

6.

Arai, T.; Iimura, S.; Hosono, H. Doping induced polymorph and carrier polarity changes in LaSeF. Chem. Mater. 2018, 30, 597–601.

7.

Togo, A.; Oba, F.; Tanaka, I.; Tatsumi, K. First-principles calculations of native defects in tin monoxide. Phys. Rev. B 2006, 74, 195128.

8.

Raebiger, H.; Lany, S.; Zunger, A. Origins of the p-type nature and cation deficiency in Cu2O and related materials. Phys. Rev. B 2007, 76, 045209.

9.

Pohl, J.; Albe, K. Intrinsic point defects in CuInSe2 and CuGaSe2 as seen via screened-exchange hybrid density functional theory. Phys. Rev. B 2013, 87, 25

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 32 of 34

K. Hanzawa et al.,

245203. 10. Hoffmann, R. A chemical and theoretical way to look at bonding on surfaces. Rev. Mod. Phys. 1988, 60, 601. 11. Burdett, J. K. From bonds to bands and molecules to solids. Prog. Solid State Chem. 1984, 15, 173−255. 12. Goldschmidt, V. M. Die gesetze der krystallochemie. Naturwissenschaften 1926, 14, 477–485. 13. Hosono, H. Exploring electro-active functionality of transparent oxide materials. Jpn. J. Appl. Phys. 2013, 52, 090001. 14. Mӧnch, W. Semiconductor surfaces and interfaces. Springer, Berlin 2001. 15. Sun, Y. Y.; Agiorgousis, M. L.; Zhang, P.; Zhang, S. Chalcogenide perovskites for photovoltaics. Nano Lett. 2015, 15, 581–585. 16. Lelieveld, R.; IJdo, D. J. W. Sulphides with the GdFeO3 structure. Acta Cryst. B 1980, 36, 2223–2226. 17. Shannon, R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 1976, 32, 751–767. 18. Pauling, L. The nature of the chemical bond, 3rd Edn. Cornell University Press, N. Y. 1960. 19. Varshni, Y. P. Temperature dependence of the energy gap in semiconductors. Physica 1967, 34, 149–154. 20. Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207. 21. Mostofi, A. A.; Yates, J.; Lee, Y. S.; Souza, I.; Vanderbilt, D.; Marzari, N. Wannier90: A tool for obtaining maximally-localised Wannier functions. Comput. 26

ACS Paragon Plus Environment

Page 33 of 34 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

Journal of the American Chemical Society

K. Hanzawa et al.,

Phys. Commun. 2008, 178, 685−699. 22. Bruker AXS. TOPAS version 4.2: General profile and structure analysis software for powder diffraction data. (Bruker AXS, Karlsruhe, Germany, 2009).

27

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 34 of 34

K. Hanzawa et al.,

TOC

28

ACS Paragon Plus Environment