Crystal structure and luminescence properties of lead-free metal halides (C6H5CH2NH3)3MBr6 (M = Bi and Sb)

Da Chen a, Fulong Dai a, Shiqiang Hao b, Guojun Zhou a, Quanlin Liu a, Christopher Wolverton b, Jing Zhao *a and Zhiguo Xia *ac
aThe Beijing Municipal Key Laboratory of New Energy Materials and Technologies, School of Materials Sciences and Engineering, University of Science and Technology Beijing, Beijing 100083, China. E-mail: jingzhao@ustb.edu.cn; xiazg@ustb.edu.cn
bDepartment of Materials Science and Engineering, Northwestern University, Evanston, Illinois 60208, USA
cState Key Laboratory of Luminescent Materials and Devices and Institute of Optical Communication Materials, South China University of Technology, Guangzhou 510641, China

Received 2nd February 2020 , Accepted 17th March 2020

First published on 19th March 2020


Abstract

Organic–inorganic hybrid metal halides have received extensive attention owing to their versatile structures and optoelectronic properties. Herein, we report two lead-free metal halides, (PMA)3BiBr6 and (PMA)3SbBr6 [PMA+: (C6H5CH2NH3)+, as the abbreviation of phenylmethylammonium], which possess iso-structural zero-dimensional structures and crystallize in the monoclinic space group P21/c. (PMA)3BiBr6 and (PMA)3SbBr6 exhibit optical band gaps of ∼3.50 and 3.40 eV, respectively, and density functional theory calculations reveal their indirect bandgap behaviors. Upon 350 and 425 nm excitation, (PMA)3BiBr6 and (PMA)3SbBr6 exhibit broadband emission peaking at 510 nm and 625 nm with wide full-widths at half-maximum of ∼153 and 175 nm, respectively. The emission mechanism of the metal halides is attributed to self-trapped exciton emission. The relationship between the crystal structure and luminescence intensity is also discussed. Finally, both metal halides have high decomposition temperatures and are stable for long-term storage under ambient conditions, demonstrating their potential for optoelectronic applications.


1. Introduction

Recently, organic–inorganic metal halides (OIMHs) have attracted extensive attention, and great progress has been made in the fields of material development, interface engineering, and device design. This can be attributed to the excellent optoelectronic properties of OIMHs originating from their band gap tunability, low defect density, long carrier lifetime, and high carrier mobility.1 Moreover, their solution processability at room temperature makes OIMHs attractive for various applications,2 including solar cells,3 lasers,4 photodetectors,5 quantum dots,6 and light-emitting diodes.7

Despite the great potential of OIMHs for commercial application in solar cells and optoelectronic devices, two shortcomings should be addressed: (i) most OIMHs contain highly toxic Pb element and (ii) their instability against moisture and temperature shortens the lifetime of the devices. Two approaches have been proposed to overcome these issues: (i) the design of lead-free metal halides by replacing Pb2+ with other cations and (ii) the exploration of low-dimensional metal halide materials by introducing hydrophobic macromolecular organic cations. Specifically, for approach (i), attempts have been made to replace Pb2+ with less toxic Sn2+ or Ge2+; however, both Sn2+ and Ge2+ are easily oxidized to Sn4+ and Ge4+ in air.8,9 In addition, monovalent and trivalent cations can be used to replace Pb2+ to synthesize double perovskites with the general formula of A2B′B′′X6 (A = CH3NH3+ or Cs+; B′ = Ag+ or Cu+; B′′ = Bi3+, Sb3+, or In3+).10 Bi3+ or Sb3+ and a vacancy co-substitution of Pb2+ is another appealing option, as Bi3+ and Sb3+ possess the same electronic structure as Pb2+,11,12 preserving the advantages of Pb2+ with highly anisotropic electronic behavior, low conduction-band effective mass, and high electron mobility.13,14 For approach (ii), the bulky organic cations can cut the inorganic octahedra into 2-, 1-, or even 0-dimensional structures.15 This reduction of the dimension is beneficial for improving the stability.16 Additionally, low-dimensional structures can induce efficient wide-band luminescence, which originates from the enhanced self-trapped excitons (STEs).17,18 These approaches provide various options for improving metal halides in terms of the structural dimension and the selection of organic cations.

Under the guidance of the above ideas, we selected Bi3+/Sb3+ and C6H5CH2NH3+ [phenylmethylammonium, abbreviated as PMA+] to design new low-dimensional OIMHs and studied the photoelectric properties. Presently, some Sb- or Bi-based OIMHs have been reported to show great potential for optoelectronic applications, such as (C4N2H14Br)4SnX6 (X = Br or I)16 and (C8NH12)4Bi0.57Sb0.43Br7·H2O.19 Other reported Sb- and Bi-based OIMHs include (CH3NH3)3Sb2I9,20 (C6H20N3)BiI6·H2O,21 (C5H10N)2BiBr5,22 (C7H7)MI4 (M = Bi or Sb),23 (C7H10NO)2BiBr5,24 (C7H18N2O)3Sb4I18·H2O,25 and (C7H18N2O)Sb2I8·H2O;25 however, despite their crystal structures being reported, their luminescence properties have not yet been comprehensively studied.

Herein, we prepared two environmentally friendly lead-free 0-dimensional (0D) OIMHs with the chemical formulae (PMA)3BiBr6 (1) and (PMA)3SbBr6 (2) using a facile solution method. Both compounds exhibit typical semiconducting behavior with experimental band gaps of ∼3.50 and 3.40 eV for 1 and 2, respectively. Upon 350 and 425 nm excitation, large Stokes-shifted broadband emission was detected, with very wide full-widths at half-maximum (FWHMs). In addition, both compounds exhibited good thermal and ambient stability. The luminescence mechanism of 1 and 2 and the relationship between the crystal structure and luminescence efficiency are discussed in detail.

2. Experimental section

Synthesis

Materials. Phenylmethylammonium bromide (PMABr) (≥99.5%) was purchased from Xi’an Polymer Light Technology Corp. Bismuth(III) bromide (BiBr3) (99.9%) and antimony(III) bromide (SbBr3) (99.9%) were purchased from Aladdin Company. Hydrogen bromide (HBr) (48 wt% in H2O) was purchased from Beijing Chemical Works.
Synthesis of 1. A stoichiometric mixture of PMABr (0.334 g) and BiBr3 (0.266 g) was dissolved in 4.6 mL of HBr under heating and continuous stirring at ∼92 °C. After all the reactants dissolved in the solution, an orange-yellow solution was obtained, which was slowly cooled to ∼25 °C. Finally, light-yellow, long rod-like crystals with sizes up to 1 cm were obtained. The yield of the reaction based on Bi element was ∼72%.
Synthesis of 2. A stoichiometric mixture of PMABr (0.366 g) and SbBr3 (0.234 g) was dissolved in 4.8 mL of HBr and heated to 92 °C under continuous stirring, forming a light-yellow solution, which was slowly cooled to ∼25 °C. Finally, large bulk light-yellow crystals with sizes up to 1 cm were obtained. The yield of the reaction based on Sb element was ∼60%.

Characterization

A selected single crystal was subjected to X-ray diffraction (SXRD) analysis using a Rigaku XtaLAB PRO single-crystal diffractometer with Mo Kα radiation. XPREP software was used to determine the crystal structure using the direct method, and the SHELXTL package was used for structure refinement.26 Powder X-ray diffraction (PXRD) patterns were collected using a D8 Advance powder diffractometer (Bruker Corporation, Germany) equipped with a Cu Kα radiation source, operating at 40 kV and 40 mA. Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS) analysis were performed using a JEOL JSM-6510A scanning electron microscope. Thermogravimetric analysis (TGA) was performed using a Setaram Labsys Evo at a heating rate of 10 °C min−1 and a maximum temperature of 500 °C. Absorption spectra were collected using an ultraviolet-visible-near-infrared (UV-vis-NIR) spectrophotometer (Hitachi UH4150) at room temperature. Steady-state photoluminescence (PL) spectra and luminescence decay curves were obtained using an FLSP9200 fluorescence spectrophotometer (Edinburgh Instruments Ltd, UK) with a PMT detector and a 150 W Xe lamp.

DFT calculation

Density functional theory (DFT) calculation was performed under periodic boundary conditions using a plane-wave basis set as implemented in the Vienna ab initio simulation package.27 The DFT calculations were converged to approximately 3 meV per cation using a basis set energy cutoff of 500 eV and dense k-meshes corresponding to 4000 k-points per reciprocal atom in the Brillouin zone.

3. Results and discussion

Synthetic methods

Both compounds exhibited good crystal quality, with crystal sizes reaching the centimeter level (Fig. 1). They were synthesized via direct combination of the organic cation precursor and MBr3 (M = Bi or Sb) in an aqueous solution of HBr at 92 °C followed by slow cooling of the solution to ∼25 °C at a cooling rate of 1.0 °C h−1. Fig. S1 (ESI) presents the SEM images and EDS analysis results of compounds 1 and 2. The EDS elemental analysis indicates that the atomic ratio of M[thin space (1/6-em)]:[thin space (1/6-em)]Br (M = Bi and Sb) is approximately 1[thin space (1/6-em)]:[thin space (1/6-em)]5.6, which is consistent with the crystallographic analysis.
image file: d0tc00562b-f1.tif
Fig. 1 (a) Crystal structures of 1 and 2. (b) Structure of the organic cation of PMA+, photographs of single crystals of 1 and 2, and distorted octahedra in 1 and 2, respectively.

Crystal structure

Compounds 1 and 2 are isostructural, and the unit cells and detailed crystallographic data are shown in Table 1. Both compounds crystallized in the monoclinic space group P21/c; other structural details (Tables S1–S8, ESI) and crystallographic information files (CIFs) for 1 and 2 are presented in the Supporting Information. As illustrated in Fig. 1a, compounds 1 and 2 possess a 0D structure with isolated octahedra. Each Bi atom is coordinated with six adjacent Br atoms, forming distorted octahedra with bond lengths ranging from 2.829 to 2.941 Å (Fig. 1b). These bond lengths are in the same range as those reported for [C6H20N3]BiBr6·H2O (2.829–2.8643 Å) and C4H16N3BiBr6 (2.829–2.8643 Å).28,29 The Br–Bi–Br bond angles range from 85.75° to 99.69° (Fig. 1b). The deviation of these values from 90° indicates that the inorganic octahedra are distorted. In compound 2, the Sb–Br bond length ranges from 2.732 to 2.932 Å (Fig. 1b), which is slightly shorter than the Bi–Br bond length; this result is consistent with the Shannon radii of 1.03 Å for Bi and 0.76 Å for Sb in six coordination environments.30 The Br–Sb–Br angle ranges from 85.33° to 99.77°, which also indicates that the inorganic octahedra are distorted (Fig. 1b).
Table 1 Crystal data and structure refinement of 1 and 2 at 100(2) K
Compound (PMA)3BiBr6 (PMA)3SbBr6
R = Σ||Fo| − |Fc||/Σ|Fo|, wR = {Σ[w(|Fo|2 − |Fc|2)2]}/Σ[w(|Fo|4)]1/2 and w = 1/[σ2(Fo2) + (0.0462P)2] where P = (Fo2 + 2Fc2)/3.
Empirical formula (C6H5CH2NH3)3BiBr6 (C6H5CH2NH3)3SbBr6
Formula weight 1011.91 925.69
Temperature 100(2) K 100(2) K
Wavelength 0.71069 Å 0.71073 Å
Crystal system Monoclinic Monoclinic
Space group P21/c P21/c
Unit cell dimensions a = 15.425 Å, α = 90° a = 15.246 Å, α = 90°
b = 8.174 Å, β = 99.56° b = 8.0632 Å, β = 100.005°
c = 24.610 Å, γ = 90° c = 24.3042 Å, γ = 90°
Volume 3059.6 Å3 2938.19 Å3
Z 4 4
Density (calculated) 2.197 g cm−3 2.093 g cm−3
Absorption coefficient 13.612 mm−1 9.113 mm−1
F(000) 1876 1752
θ range for data collection 2.314 to 25.00° 1.984° to 29.346°
Index ranges −18 ≤ h ≤ 18, −9 ≤ k ≤ 9, −29 ≤ l ≤ 29 −18 ≤ h ≤ 20, −10 ≤ k ≤ 10, −32 ≤ l ≤ 32
Reflections collected 25[thin space (1/6-em)]982 35[thin space (1/6-em)]727
Independent reflections 5375 [Rint = 0.0553] 7269 [Rint = 0.0304]
Completeness to θ = 25.00° 99.7% 100%
Refinement method Full-matrix least-squares on F2
Data/restraints/parameters 5378/10/283 7296/18/283
Goodness-of-fit 1.130 1.019
Final R indices [I > 2σ(I)] R obs = 0.0801, wRobs = 0.1869 R obs = 0.0334, wRobs = 0.0559
R indices [all data] R all = 0.1019, wRall = 0.2023 R all = 0.0506, wRall = 0.0593
Largest diff. peak and hole 2.770 and −0.836 e Å−3 0.673 and −0.648 e Å−3


Comparing 1 and 2 with the reported (C6H5CH2CH2NH3)3BiBr6 with the I2/a space group, all of them belong to the centrosymmetric monoclinic system with a 0D structure consisting of alternating [MBr6]3− (M = Bi or Sb) octahedra and organic cations linked by hydrogen bonding. As illustrated in Fig. S2 (ESI), the hydrogen bond length ranges from 2.541 to 2.986 Å (Fig. S2a, ESI) for compound 1 and from 2.556 to 2.990 Å (Fig. S2b, ESI) for compound 2, which are in accordance with the bond lengths reported for [(CH3)2NH2]6(BiBr6)2 (2.500–2.770 Å),31 (C6H20N3)BiBr6·H2O (2.440–3.110 Å),28 and (C7H10NO)2BiBr5 (2.470–2.980 Å).24

Stability studies

The experimental PXRD patterns of 1 and 2 and calculated patterns based on single-crystal structures are compared in Fig. 2a and b. The experimental PXRD and calculated peaks fit well, and no extra peaks were observed, indicating the high purity of the obtained samples. In addition, we compared the PXRD patterns of 1 and 2 samples stored in air for 3 months to those of the as-synthesized samples. No peak position or intensity changes were observed, indicating that both 1 and 2 are stable in air.
image file: d0tc00562b-f2.tif
Fig. 2 (a) Comparison of calculated PXRD patterns based on SXRD data and experimental PXRD patterns of polycrystalline powders of 1 and 2 measured on fresh samples and after storage in air for 3 months. (b) Selected diffraction peaks in the range of 5°–10° of 1 and 2.

To investigate the thermal stability of the two compounds at elevated temperature, TGA measurements were conducted under nitrogen flow from room temperature to 500 °C. Sample 1 was stable below ∼273 °C, and a multiple-step decomposition profile was obtained for the compound (Fig. 3a). In the first stage, approximately 32% of the initial weight was lost, which can be attributed to the volatilization of the PMA (C6H5CH2NH2) (the theoretical calculation is approximately 32.00%). There was an obvious platform when the temperature was higher than 325 °C, and until 325 °C, the weight loss was approximately 61%. To verify the residual content, fresh samples were heated to ~325 °C in a muffle furnace and held for 2 h; then, the residues were collected and examined using PXRD (Fig. S3, ESI). After searching in the Cambridge Crystallographic Data Centre (CCDC) database, the product was identified as a mixture of BiBr3 and Bi6Br7. Thus, the weight loss of sample 1 may be attributed to the volatilization of PMA and subsequently Br2. TGA revealed that compound 2 was stable up to ∼232 °C (Fig. 3b), showing a two-step decomposition. After the first step, it lost about ∼35% of the initial weight, corresponding to the volatilization of PMA with a calculated value of ∼35.02%. After the second step, only 18% of the weight remained.


image file: d0tc00562b-f3.tif
Fig. 3 TGA data for powder samples of 1 (a) and 2 (b) under a N2 atmosphere.

Compounds 1 and 2 exhibited higher thermal stability than reported OIMHs with small organic molecules such as MA3Bi2Br9, (C5H10N)2BiBr5, and (C8NH12)4Bi0.57Sb0.43Br7·H2O, which start to decompose at approximately 200, 140, and 120 °C, respectively.19,22,32 Thus, the thermal stability of OIMHs containing large organic molecules appears to be higher than that of OIMHs containing small organic molecules.

Optical properties and DFT calculations

The Raman spectra of compound 1 was studied by Fehlner et al.,33 but no structure or other properties were characterized. Herein, the spectroscopic properties of both 1 and 2 were investigated in detail. Fig. 5a presents the absorption spectra of polycrystal powders 1 and 2 and the band gaps of 1 and 2 were estimated to be ∼3.50 and ∼3.40 eV, respectively. As illustrated in Fig. 5a, the absorption peaks are slightly lower than the band gaps of compounds 1 and 2 corresponding to the exciton absorption peaks with peak energies of ∼3.28 and 3.38 eV, respectively, at room temperature. The presence of sharp exciton peaks at room temperature indicates that the compound has high exciton binding energy.34 The absorption band at higher energy corresponds to the interband absorption. These absorption peaks are similar to those of previously reported OIMH compounds, including [(CH3)2NH2]3[BiI6] (∼2.60 and 3.04 eV),35 (C6H20N3)BiBr6·H2O (∼2.36 and 3.10 eV),28 C4H16N3BiBr6 (∼3.22 and 4.08 eV),29 and [(CH3)2NH2]6(BiBr6)2 (∼3.30 and 3.97 eV).31

The specific bandgap types were determined by DFT calculation. The band structures of compounds 1 and 2 consist of indirect (H to X) bandgaps of 3.24 and 2.77 eV, respectively, thus exhibiting weak PL emission (Fig. 4a and b). For compound 1, the valence band maximum (VBM) consists of Bi 6s and Br 4p states at H, and the conduction band minimum (CBM) consists of Bi 6p and Br 4p states at X (Fig. 4c). Similarly, the VBM of 2 mainly consists of Sb 5s and Br 4p states at H, and the CBM consists of Sb 5p and Br 4p states at X (Fig. 4d).


image file: d0tc00562b-f4.tif
Fig. 4 Band structures of 1 (a) and 2 (b). Density of states of 1 (c) and 2 (d).

Fig. 5b presents the normalized PL emission spectra of compounds 1 and 2. Upon 350 nm excitation, the PL spectrum of compound 1 exhibited two broad emission bands peaking at 405 and 510 nm; and the one with low energy is much stronger and possesses a large Stokes shift of 160 nm and a FWHM of ∼153 nm. Upon excitation with 425 nm light, the PL spectrum of compound 2 also exhibited two bands peaking at 453 and 625 nm, and the one with low energy is stronger with a large Stokes shift of 200 nm and a large FWHM of ∼175 nm. This broad emission is commonly observed in low-dimensional OIMHs with strong electron–phonon coupling.19,36 For example, (C8NH12)4Bi0.57Sb0.43Br7·H2O exhibits white light emission with a FWHM of ∼160 nm,19 and (C6H20N3)BiBr6·H2O exhibits broadband emission with a FWHM of ∼154 nm.28


image file: d0tc00562b-f5.tif
Fig. 5 (a) Optical absorption spectra. (b) PL spectra upon 350 and 425 nm excitation of 1 and 2 at room temperature, respectively. (c) Emission intensity versus excitation power for 1 and 2 at room temperature. (d) Mechanism diagram of photoluminescence for 1 and 2. (e) CIE color coordinates of 1 and 2 in the 1931 color space chromaticity diagram.

In order to explore the emission mechanism, experiments were conducted on the dependence of emission intensity and excitation power. As shown in Fig. 5c, a linear relationship without saturation is observed, which indicates that the broadband emission does not originate from permanent defects.37 In addition, the PL spectrum of the staring material PMABr was collected under the same conditions to prove that the luminescence does not come from the emission of organic molecules. The emission band of PMABr peaks at 425 nm (Fig. S4, ESI) which is very different from the emission characteristics of compounds 1 and 2. The PL spectra of the two compounds can be attributed to singlet and triplet emissions of isolated ns2 metal centers, which are also known as STE emission. In the present work, the ground state is 1S0 (here is 6s2 or 5s2 of Bi3+ or Sb3+), and the excited states are derived from the 6s6p (Bi3+) or 5s5p (Sb3+) orbital, which splits into four levels, namely 1P1, 3P0, 3P1, and 3P2.38 According to the transition rules, the high energy emission bands of 1 and 2 peaking at 405 and 453 nm, respectively, are attributed to the 1P11S0 transition (Fig. 5d). The low energy broadband emission of compounds 1 and 2 peaking at 510 and 625 nm, respectively, is attributed to the 3Pn1S0 (n = 0, 1, and 2) transitions owing to intersystem crossing (ISC). This ISC is accelerated in heavy elements due to larger spin–orbit coupling (SOC), leading to the mixing of 1P and 3P states. These emissions are previously discussed in the compounds of Bmpip2MBr4 (Bmpip is the abbreviation of 1-butyl-1-methyl-piperidinium; M = Ge, Sn, and Pb)39 and Cs2SnCl6:Sb3+ (Sb/Sn = 0.1).40 Upon higher energy photoexcitation at 330 and 400 nm for compounds 1 and 2, respectively, the high energy emission from the singlet state becomes stronger (Fig. S5, ESI), which is in accordance with that reported by Morad, V. et al.39 The luminescence decay curves of 1 and 2 are shown in Fig. S6 (ESI). Due to the low PL intensity, only the lifetimes of the triplet broadband emissions are detected, which are 1.031 and 1.508 ns for 1 and 2, respectively. The triplet lifetime is short, which is the result of a mixture of singlet and triplet states. This can be verified with a relatively small Stokes shift of 1.11 eV of 1 and 0.94 eV of 2.39 The CIE color coordinates of 1 and 2 are (0.3118, 0.4164) and (0.4981, 0.4445), respectively (Fig. 5e).

Fig. S7 (ESI) presents the PL spectra of 1 and 2 at low temperatures, respectively. As the temperature decreases, the intensity of the emissions increases monotonically, which can be attributed to the reduction of the nonradiative decay. With increasing temperature, the band widening of compounds 1 and 2 can be observed, and the increase of the FWHM can be described by the following law:41

image file: d0tc00562b-t1.tif
where f(T) is the FWHM at different temperatures, S is the Huang–Rhys parameter, k is the Boltzmann constant, and Eph is the effective phonon energy. The fitting results are presented in Fig. S7c and d (ESI), and the FWHM shows thermal dependence. We can determine that compound 1 has S = 13.7 and Eph = 83.2 meV and that compound 2 has S = 12.5 and Eph = 56.9 meV. The S values of compounds 1 and 2 are larger than those of inorganic compounds, such as 6.4 for SrSe,42 8.4 for Y3Al5O12,43 and 9.0 for Lu3Al5O12,44 indicating that 1 and 2 have strong electron–phonon coupling. Moreover, the values of Eph are much higher than those of all-inorganic metal halides, such as Cs3Bi2I6Cl3 (Eph = 4.0 meV)45 and Cs2AgInCl6 (Eph = 17.4 meV),46 implying that the energies of the lattice vibration for 1 and 2 are relatively large and the structural rigidity is relatively weak.

Both compounds 1 and 2 exhibit relatively weak PL intensity compared with that of their In-based analogue (PMA)3InBr6, which exhibited a photoluminescence quantum yield (PLQY) of ∼35%.47 To reveal the structural effects of PL intensities in these 0D compounds, we compared our result with those for other previously reported 0D compounds with six coordinated s2 metal cations. First and foremost, the effect of the band nature should be considered. (PMA)3InBr6 is isostructural with 1 and 2, showing an efficient broadband orange emission with the mechanism of STE emission. DFT calculation indicated that (PMA)3InBr6 possesses a direct bandgap. Thus, compared with indirect-band-gap compounds 1 and 2, in light of the absorption process of (PMA)3InBr6, no phonon processes are needed, which is more efficient, resulting in higher PLQY. In addition, previous works have shown that the formation of the STE states of OIMHs with the same structure dimensionality may be related to the distortion level of the metal coordination geometry, and also related to many other factors, such as the structural disorder, organic molecules, PL quenching temperature, etc.48,49 Herein, it is of interest to discuss the PL characterization and structural relationship of (PMA)3MBr6 (M = Bi, Sb, and In). With the purpose of characterizing the difference in octahedral distortion, we defined the distortion based on the M–X bond length and X–M–X (M and X are metal cations and halogen anions) angle variance:50

image file: d0tc00562b-t2.tif

image file: d0tc00562b-t3.tif
where dn denotes six individual M–X bond lengths, d0 is the average of the M–X bond distance, and θn denotes the bond angles of each X–M–X. The results are tabulated in Table 2.

Table 2 Summary of bandgap behavior, PL emission, FWHM, bond length distortions, angle variance and PLQY for title compounds and 0D compounds with six coordinated metal cations reported in the literature
Compound Bandgap behavior PL (nm) FWHM (nm) λ oct (×10−4) σ 2 PLQY (%) Ref.
(PMA)3BiBr6 Indirect 510 153 2.1 14.57 <1 This work
(PMA)3SbBr6 Indirect 640 175 7.1 14.60 <1 This work
(C4N2H14Br)4SnBr6 Direct 570 105 45.69 18.20 95 16
(Ph4P)2SbCl5 Direct 648 136 64.5 8.20 87 51
(C4N2H14Br)4SnI6 Direct 620 118 1.57 19.45 75 16
(PMA)3InBr6 Direct 610 132 1.1 4.99 35 47
(C3N3H11O)2PbBr6·4H2O Direct 568 200 0.00438 2.57 9.3 52
(C8NH12)4Bi0.57Sb0.43Br7·H2O Direct 450, 640 400–850 5.94 8.04 4.5 19
(C8NH12)4BiBr7·H2O 450 12.34 10.62 0.7 19


In Table 2, we can find that with a direct band gap and very large structural distortion, (C4N2H14Br)4SnBr6,16 (Ph4P)2SbCl551 (Ph4P is tetraphenylphosphonium cation) and (C4N2H14Br)4SnI616 exhibit very high PL emission efficiency. However, PLQY is not linearly related to the degree of structural distortion. For example, (PMA)3InBr6 shows a less distorted structure than (C8NH12)4Bi0.57Sb0.43Br7·H2O19 but higher PLQY; (C3N3H11O)2PbBr6·4H2O52 possesses a distortion level lower than that of (C8NH12)4Bi0.57Sb0.43Br7·H2O but higher PLQY.

To compare the PL intensities of compounds 1 and 2 (Fig. S8, ESI), PL spectra were obtained under exactly the same conditions. The PL intensity of 2 is much larger than that of 1 (Fig. S8, ESI), which correlates well with the distortion level of compound 2 (λoct = 7.1 × 10−4 and σ2 = 14.60) being greater than that of compound 1 (λoct = 2.1 × 10−4 and σ2 = 14.57). Notably, the distortion degrees of 1 and 2 are larger than that of (PMA)3InBr6, which may theoretically indicate stronger emission than the latter, provided that the band gaps of 1 and 2 can be tuned to be direct ones through doping.

4. Conclusions

In summary, we prepared two new lead-free OIMHs (PMA)3BiBr6 (1) and (PMA)3SbBr6 (2), which crystallized in the monoclinic space group P21/c. The two compounds are isostructural 0D structures and are composed of discrete octahedra and organic cations linked by hydrogen bonding. The DFT calculations revealed the indirect band gap behavior of 1 and 2, and the optical band gap was experimentally determined to be ∼3.50 and 3.40 eV, respectively. Both compounds 1 and 2 exhibited good ambient stability and relatively high decomposition temperatures under N2 gas flow (>230 °C). Upon 350 and 425 nm excitation, compounds 1 and 2 exhibited large Stokes-shifted dual broad PL emission with the FWHMs of ∼153 and 175 nm, respectively. This broad emission in the 0D Bi- and Sb-based OIMHs is attributed to STE (singlet and triplet) emission. The thermal dependence of the FWHM was discussed in detail. The discovery of environmentally stable 1 and 2 and their photoluminescence properties further demonstrate the importance of lead-free low-dimensional OIMHs for optoelectronic applications.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was funded by the Beijing Natural Science Foundation (2182080) and the National Natural Science Foundation of China (51702329, 51961145101 and 51972021). This work was also supported by the Fundamental Research Funds for the Central Universities FRF-TP-18-005A1 and FRF-IDRY-19-005. S. H. and C. W. (DFT calculations) acknowledge support from the Department of Energy, Office of Science Basic Energy Sciences under Grant DE-SC0014520. Access to QUEST, the supercomputing resources facilities at Northwestern University, is also acknowledged.

Notes and references

  1. D. Shi, V. Adinolfi, R. Comin, M. Yuan, E. Alarousu, A. Buin, Y. Chen, S. Hoogland, A. Rothenberger, K. Katsiev, Y. Losovyj, X. Zhang, P. A. Dowben, O. F. Mohammed, E. H. Sargent and O. M. Bakr, Science, 2015, 347, 519 CrossRef CAS PubMed.
  2. Z. Song, J. Zhao and Q. Liu, Inorg. Chem. Front., 2019, 6, 2969–3011 RSC.
  3. M. Kim, G.-H. Kim, T. K. Lee, I. W. Choi, H. W. Choi, Y. Jo, Y. J. Yoon, J. W. Kim, J. Lee, D. Huh, H. Lee, S. K. Kwak, J. Y. Kim and D. S. Kim, Joule, 2019, 3, 2179–2192 CrossRef CAS.
  4. Q. Zhang, S. T. Ha, X. Liu, T. C. Sum and Q. Xiong, Nano Lett., 2014, 14, 5995–6001 CrossRef CAS PubMed.
  5. L. Dou, Y. M. Yang, J. You, Z. Hong, W. H. Chang, G. Li and Y. Yang, Nat. Commun., 2014, 5, 5404 CrossRef CAS PubMed.
  6. M. Leng, Z. Chen, Y. Yang, Z. Li, K. Zeng, K. Li, G. Niu, Y. He, Q. Zhou and J. Tang, Angew. Chem., Int. Ed., 2016, 55, 15012–15016 CrossRef CAS PubMed.
  7. N. Wang, L. Cheng, R. Ge, S. Zhang, Y. Miao, W. Zou, C. Yi, Y. Sun, Y. Cao, R. Yang, Y. Wei, Q. Guo, Y. Ke, M. Yu, Y. Jin, Y. Liu, Q. Ding, D. Di, L. Yang, G. Xing, H. Tian, C. Jin, F. Gao, R. H. Friend, J. Wang and W. Huang, Nat. Photonics, 2016, 10, 699–704 CrossRef CAS.
  8. F. Hao, C. C. Stoumpos, D. H. Cao, R. P. H. Chang and M. G. Kanatzidis, Nat. Photonics, 2014, 8, 489–494 CrossRef CAS.
  9. C. C. Stoumpos, L. Frazer, D. J. Clark, Y. S. Kim, S. H. Rhim, A. J. Freeman, J. B. Ketterson, J. I. Jang and M. G. Kanatzidis, J. Am. Chem. Soc., 2015, 137, 6804–6819 CrossRef CAS PubMed.
  10. X. Zhao, L. Zhang, Z. Xiao and Y. Sun, Joule, 2018, 2, 1662–1673 CrossRef CAS.
  11. P. C. Harikesh, H. K. Mulmudi, B. Ghosh, T. W. Goh, Y. T. Teng, K. Thirumal, M. Lockrey, K. Weber, T. M. Koh, S. Li, S. Mhaisalkar and N. Mathews, Chem. Mater., 2016, 28, 7496–7504 CrossRef CAS.
  12. J.-C. Hebig, I. Kühn, J. Flohre and T. Kirchartz, ACS Energy Lett., 2016, 1, 309–314 CrossRef CAS.
  13. G. G. Briand, N. Burford, M. D. Eelman, N. Aumeerally, L. Chen, T. S. Cameron and K. N. Robertson, Inorg. Chem., 2004, 43, 6495–6500 CrossRef CAS PubMed.
  14. X.-Y. Liu, J.-H. Zeng, S.-Y. Zhang, R.-B. Zheng, X.-M. Liu and Y.-T. Qian, Chem. Phys. Lett., 2003, 374, 348–352 CrossRef CAS.
  15. D. H. Cao, C. C. Stoumpos, O. K. Farha, J. T. Hupp and M. G. Kanatzidis, J. Am. Chem. Soc., 2015, 137, 7843–7850 CrossRef CAS PubMed.
  16. C. Zhou, H. Lin, Y. Tian, Z. Yuan, R. Clark, B. Chen, L. J. van de Burgt, J. C. Wang, Y. Zhou, K. Hanson, Q. J. Meisner, J. Neu, T. Besara, T. Siegrist, E. Lambers, P. Djurovich and B. Ma, Chem. Sci., 2018, 9, 586–593 RSC.
  17. S. Tomimoto, H. Nansei, S. Saito, T. Suemoto, J. Takeda and S. Kurita, Phys. Rev. Lett., 1998, 81, 417–420 CrossRef CAS.
  18. Y. Wada, U. Lemmer, E. O. Göbel, M. Yamashita and K. Toriumi, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 52, 8276–8282 CrossRef CAS PubMed.
  19. R. Zhang, X. Mao, Y. Yang, S. Yang, W. Zhao, T. Wumaier, D. Wei, W. Deng and K. Han, Angew. Chem., Int. Ed., 2019, 58, 2725–2729 CrossRef CAS PubMed.
  20. N. Elfaleh and S. Kamoun, J. Organomet. Chem., 2016, 819, 95–102 CrossRef CAS.
  21. J. Zaleski, R. Jakubas, L. Sobczyk and J. Mróz, Ferroelectrics, 1990, 103, 83–90 CrossRef CAS.
  22. W. Masmoudi, S. Kamoun, H. F. Ayedi and K. Guidara, Phys. B, 2012, 407, 2365–2371 CrossRef CAS.
  23. I. W. H. Oswald, E. M. Mozur, I. P. Moseley, H. Ahn and J. R. Neilson, Inorg. Chem., 2019, 58, 5818–5826 CrossRef CAS PubMed.
  24. Z. Aloui, V. Ferretti, S. Abid, F. Lefebvre, M. Rzaigui and C. Ben Nasr, J. Mol. Struct., 2016, 1118, 301–306 CrossRef CAS.
  25. Y. Li, Z. Xu, X. Liu, K. Tao, S. Han, Y. Wang, Y. Liu, M. Li, J. Luo and Z. Sun, Inorg. Chem., 2019, 58, 6544–6549 CrossRef CAS PubMed.
  26. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
  27. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
  28. N. Elfaleh and S. Kamoun, J. Mol. Struct., 2014, 1075, 479–485 CrossRef CAS.
  29. H. Dammak, H. Feki, H. Boughzala and Y. Abid, Spectrochim. Acta, Part A, 2015, 137, 1235–1243 CrossRef CAS PubMed.
  30. R. Shannon, Acta Crystallogr., Sect. A: Found. Adv., 1976, 32, 751–767 CrossRef.
  31. A. Ben Ahmed, H. Feki and Y. Abid, Spectrochim. Acta, Part A, 2014, 133, 357–364 CrossRef CAS PubMed.
  32. M. Leng, Y. Yang, K. Zeng, Z. Chen, Z. Tan, S. Li, J. Li, B. Xu, D. Li, M. P. Hautzinger, Y. Fu, T. Zhai, L. Xu, G. Niu, S. Jin and J. Tang, Adv. Funct. Mater., 2018, 28, 1704446 CrossRef.
  33. T. P. Fehlner, M. Wu, B. J. Meneghelli and R. W. Rudolph, Inorg. Chem., 1980, 19, 49–54 CrossRef CAS.
  34. A. Yangui, S. Pillet, E.-E. Bendeif, A. Lusson, S. Triki, Y. Abid and K. Boukheddaden, ACS Photonics, 2018, 5, 1599–1611 CrossRef CAS.
  35. A. Samet, A. B. Ahmed, A. Mlayah, H. Boughzala, E. K. Hlil and Y. Abid, J. Mol. Struct., 2010, 977, 72–77 CrossRef CAS.
  36. M. D. Smith and H. I. Karunadasa, Acc. Chem. Res., 2018, 51, 619–627 CrossRef CAS PubMed.
  37. E. R. Dohner, A. Jaffe, L. R. Bradshaw and H. I. Karunadasa, J. Am. Chem. Soc., 2014, 136, 13154–13157 CrossRef CAS PubMed.
  38. A. Vogler and H. Nikol, Comments Inorg. Chem., 1993, 14, 245–261 CrossRef CAS.
  39. V. Morad, Y. Shynkarenko, S. Yakunin, A. Brumberg, R. D. Schaller and M. V. Kovalenko, J. Am. Chem. Soc., 2019, 141, 9764–9768 CrossRef CAS PubMed.
  40. Y. Jing, Y. Liu, J. Zhao and Z. Xia, J. Phys. Chem. Lett., 2019, 10, 7439–7444 CrossRef CAS PubMed.
  41. J. Lee, E. S. Koteles and M. O. Vassell, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 5512–5516 CrossRef CAS PubMed.
  42. N. Yamashita, O. Harada and K. Nakamura, Jpn. J. Appl. Phys., 1995, 34, 5539–5545 CrossRef CAS.
  43. V. Bachmann, C. Ronda and A. Meijerink, Chem. Mater., 2009, 21, 2077–2084 CrossRef CAS.
  44. J. M. Ogiegło, A. Zych, K. V. Ivanovskikh, T. Jüstel, C. R. Ronda and A. Meijerink, J. Phys. Chem. A, 2012, 116, 8464–8474 CrossRef PubMed.
  45. K. M. McCall, C. C. Stoumpos, O. Y. Kontsevoi, G. C. B. Alexander, B. W. Wessels and M. G. Kanatzidis, Chem. Mater., 2019, 31, 2644–2650 CrossRef CAS.
  46. J. Luo, X. Wang, S. Li, J. Liu, Y. Guo, G. Niu, L. Yao, Y. Fu, L. Gao, Q. Dong, C. Zhao, M. Leng, F. Ma, W. Liang, L. Wang, S. Jin, J. Han, L. Zhang, J. Etheridge, J. Wang, Y. Yan, E. H. Sargent and J. Tang, Nature, 2018, 563, 541–545 CrossRef CAS PubMed.
  47. D. Chen, S. Hao, G. Zhou, C. Deng, Q. Liu, S. Ma, C. Wolverton, J. Zhao and Z. Xia, Inorg. Chem., 2019, 58, 15602–15609 CrossRef CAS PubMed.
  48. R. Gautier, F. Massuyeau, G. Galnon and M. Paris, Adv. Mater., 2019, 31, e1807383 CrossRef PubMed.
  49. L. Mao, P. Guo, M. Kepenekian, I. Hadar, C. Katan, J. Even, R. D. Schaller, C. C. Stoumpos and M. G. Kanatzidis, J. Am. Chem. Soc., 2018, 140, 13078–13088 CrossRef CAS PubMed.
  50. K. Robinson, G. V. Gibbs and P. H. Ribbe, Science, 1971, 172, 567 CrossRef CAS PubMed.
  51. C. Zhou, M. Worku, J. Neu, H. Lin, Y. Tian, S. Lee, Y. Zhou, D. Han, S. Chen, A. Hao, P. I. Djurovich, T. Siegrist, M.-H. Du and B. Ma, Chem. Mater., 2018, 30, 2374–2378 CrossRef CAS.
  52. B. B. Cui, Y. Han, B. Huang, Y. Zhao, X. Wu, L. Liu, G. Cao, Q. Du, N. Liu, W. Zou, M. Sun, L. Wang, X. Liu, J. Wang, H. Zhou and Q. Chen, Nat. Commun., 2019, 10, 5190 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available: Crystal structures of 1 and 2; tables of atomic coordinates, displacement parameters, anisotropic displacement parameters, and bond distances of 1 and 2 (Tables S1–S8); SEM and EDS results, N–H⋯Br hydrogen bonding of 1 and 2, XRD analysis of 1 after the first step decomposition, PLE and PL spectra of PMABr at room temperature, PL spectra excited by higher energy photoexcitation for compounds 1 and 2, PL decay curves of compound 1 and 2 polycrystalline powders at room temperature, temperature-dependent steady-state PL spectra of 1 and 2, fitting of the FWHM for compounds 1 and 2 as a function of temperature, temperature-dependent normalized steady-state PL spectra of 1 and 2, PL spectra of 1 and 2 at 350 K and 80 K, and comparison of PL spectra of 1 and 2 (Fig. S1–S8). CCDC 1964916 and 1964917. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0tc00562b

This journal is © The Royal Society of Chemistry 2020