Lead-Free Hybrid Indium Perovskites with Highly Efficient and Stable Green Light Emissions

铅(地质) 材料科学 绿灯 环境科学 工程物理 纳米技术 光电子学 物理 蓝光 地质学 地貌学
作者
Chen Sun,Jin-Peng Zang,Yuqing Liu,Qian‐Qian Zhong,Xinxin Xing,Jiapeng Li,Cheng‐Yang Yue,Xiao‐Wu Lei
出处
期刊:CCS Chemistry [Chinese Chemical Society]
卷期号:4 (9): 3106-3121 被引量:33
标识
DOI:10.31635/ccschem.021.202101092
摘要

Open AccessCCS ChemistryRESEARCH ARTICLE5 Sep 2022Lead-Free Hybrid Indium Perovskites with Highly Efficient and Stable Green Light Emissions Chen Sun, Jin-Peng Zang, Yu-Qing Liu, Qian-Qian Zhong, Xin-Xin Xing, Jia-Peng Li, Cheng-Yang Yue and Xiao-Wu Lei Chen Sun School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Jin-Peng Zang School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Yu-Qing Liu School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Qian-Qian Zhong School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Xin-Xin Xing School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Jia-Peng Li School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author , Cheng-Yang Yue *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author and Xiao-Wu Lei *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] School of Chemistry, Chemical Engineer and Materials, Jining University, Qufu, Shandong, 273155 Google Scholar More articles by this author https://doi.org/10.31635/ccschem.021.202101092 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Lead halide perovskite nanocrystals (PNCs) have emerged as new-generation light-emitting materials with important potentials in optoelectronic applications. However, the toxicity of lead metal and the instabilities of PNCs overwhelmingly hinder their device engineering; thus, it is still urgent to construct new single crystalline lead-free perovskites with high-performance luminescent properties. Herein, we initially designed two novel crystalline hybrid indium chlorides, [DAPEDA]InCl6·Cl·H2O ( 1) and [DPA]3InCl6 ( 2), containing discrete octahedral [InCl6]3− units. Remarkably, bulk crystals of 1 and 2 exhibited strong broadband green light emissions, derived from self-trapped excitons (STE) with high photoluminescence quantum yields (PLQYs) of 40% and 34%, respectively, with benefits from intrinsic quantum confinement and highly localized electrons. More excitedly, the controlled Sb3+-doping strategy significantly boosted the emission efficiency with the highest PLQY of 89.2%. In addition, these indium perovskites possessed ultrahigh structural and spectral stabilities in various extreme environments exceeding typical PNCs. Combined efficient and stable luminescent performances enable these indium perovskites as excellent down-conversion green phosphors to successfully fabricate white light-emitting diodes with high color rendering index of 93. To the best of our knowledge, this work represents the first fabricated green light-emitting hybrid indium perovskites as excellent candidates for lead PNCs. Download figure Download PowerPoint Introduction Three-dimensional (3D) lead perovskite nanocrystals (PNCs) of APbX3 (A = CH3NH3, Cs+, X = Cl, Br, and I) have currently emerged as new-generation light-emitting materials with diversified advantages of adjustable bandgaps, high defect tolerance, tunable and wide light spectra across visible to near-infrared region, narrow emission bandwidth and high photoluminescence quantum yield (PLQY) stemming from quantum confinement of surrounded organic ligands.1–3 Subsequently, various synthetic modulation strategies, including morphological regulation, surface passivation, metal doping, or halogen mixing, have been massively adopted to reduce the defect concentrations, regulate the excited energy states or enhance the emission efficiencies, which optimized or diversified their luminescent properties successfully and expanded the application fields.4–6 Despite the substantial optimization evolutions toward conventional PNCs, the colloidal PNCs remain to be confronted with inevitable particle aggregation along with severe optical instabilities and PLQY degradations, which, overwhelmingly, hinder their further optoelectronic device engineering. Assuredly, 3D perovskite bulk materials were rarely reported as excitonic luminescence species because the large band dispersion of corner-shared 3D network dissociates the excitons to charge carriers readily at room temperature.7 In addition, the inflexible 3D cubic network and the simple chemical composition reduce the feasibility of structural modulation from the molecular level to regulate the luminescent properties. Inspired by the fact that the highly efficient luminescent performance of PNC originates from the quantum confinement effect of nanoscale particle size, another structural design strategy of preserving the quantum confinement has been proposed and confirmed, that is, slicing the 3D corner-shared [PbX3]− network into an zero-dimensional (0D) unit by introducing wide-bandgap organic species as spatial cations.8–10 In these bulk crystalline organic–inorganic hybrid 0D perovskites, the discrete lead halide polyhedrons are embedded periodically in the insulating organic matrix with intrinsic quantum confinement distinguishing from the external effect of PNC. Specifically, the absence of electron interaction between spatially discrete polyhedrons resulted in a flat band structure, which firmly restrained the photoinduced excitons around each light-emitting species and significantly promoted the radiative recombination, finally leading to efficient light emission.11–13 Indeed, the versatile organic species endow the hybrid 0D perovskites with diversified structures, highly tunable emissions, and remarkable PLQYs.14,15 Moreover, the luminescence performances of 0D perovskites are almost independent of the crystal morphologies and sizes; hence, they provide higher spectral stabilities than 3D PNCs. Nevertheless, the environmental and biological toxicity of the lead metal remains an obstacle for the broad application of lead perovskites. Therefore, it is highly desirable and significant to construct new single crystalline lead-free 0D perovskites with highly efficient, versatile, and stable light-emitting performances. Under this condition, substantial lead-free metal halide perovskites based on tin, germanium, antimony, bismuth, manganese, indium, and others, were intensively explored as candidates of solid-state light-emitting materials in the large family of metal halide perovskites.16–19 Compared with instabilities of Sn2+ and Ge2+ halides, nearly fixed emission spectral ranges of Sb3+ and Mn2+-based perovskites, as well as low efficiencies of Bi3+ phases, In3+-based perovskites, are more desirable for down-converter phosphors, benefitting from the tunable light emissions, nontoxicity, and oxidation resistance abilities.20–23 In general, 0D indium perovskites display broadband light emissions with substantial potential in solid-state white light-emitting diodes (WLEDs).24–26 Previous works have unveiled that the broadband light emission originated from the intrinsic self-trapped excitons (STE) due to an excited-state structural distortion of [InX6]3− octahedrons or the strong electron-quantum coupling effect in the soft 0D crystal lattice.24–26 Regrettably, all these 0D indium perovskites exhibit strongly Stokes-shifted emission spectra with maximum peaks of over 600 nm as orange to red light emitters due to the large excited-state structural distortion level of isolated octahedrons. Up to date, none of the 0D indium-based perovskites is able to display higher-energy light emission with a maximum peak wavelength shorter than 600 nm. Besides, the emission efficiencies of 0D indium perovskites still need further improvement, considering the passable PLQYs, such as (C4H14N2)2In2Br10 (3%), Cs2InBr5·H2O (33%), and phenylmethylammine, C6H5CH2NH3 (PMA)3InBr6 (35%).24–26 In view of these shortcomings, we focus on this domain to enable the application of 0D indium perovskites for higher-energy light emitters to realize their full-color display by fine-tuning the molecular structure. Herein, we designed two new single crystalline lead-free 0D indium perovskites of [DAPEDA]InCl6·Cl·H2O (DAPEDA = 1,2-bis(3-aminopropylamino)ethane) and [DPA]3InCl6 (DPA = dipropylamine) successfully via facile wet-chemistry or a mechanical grinding solid-phase approach. Unlike the previous near red light-emitting phases, these 0D indium perovskites were highlighted with higher-energy broadband green light emissions (510–520 nm) with promising PLQYs of 34.01% and 40.40%, respectively, able to be optimized up to the highest value of 89.2% through rational Sb3+-doping strategy. The compelling luminescent performances and ultra-high stabilities enabled down-conversion green phosphors to fabricate WLEDs showcasing their potential in backlit applications. So far as we know, this work exemplifies the first green light-emitting hybrid indium perovskites with successful enrichment of the color gamut and device application field. Experimental Section Materials All the precursor materials of indium chloride tetrahydrate (InCl3·4H2O, 98%, Aladdin, Shanghai, China), DAPEDA (C8H22N4, 97%Aladdin, Shanghai, China), DPA (C6H15N, 99%, Aladdin, Shanghai, China), hydrochloric acid (HCl, 38%, SCRC, Shanghai, China), acetonitrile (99.8%, SCRC, Shanghai, China), and acetone (99.5%, SCRC, Shanghai, China) were commercially purchased and directly used without further purification. Synthesis of [DAPEDA]Cl4 Briefly, DAPEDA was added into HCl (38%) solution affording white depositions, which were filtrated and washed with ethanol several times. The final products were collected after drying in a vacuum oven. Synthesis of [DPA]Cl Briefly, DPA was added into HCl (38%) solution to form a clear solution and condensed via rotary evaporation with the formation of a white powder. The powder was washed with ethanol several times and collected after drying in a vacuum oven. Solution growth of [DAPEDA]InCl6·Cl·H2O (1) and [email protected][DAPEDA]InCl6·Cl·H2O single crystals InCl3·4H2O (0.5 mmol, 0.1466 g) and DAPEDA (0.6 mmol, 0.1046 g) were mixed and dissolved in 10 mL HCl (38%) with constant magnetic stirring and heating at about 120 °C leading to a clear solution. After slowly cooling to room temperature (25 °C), colorless layer-shaped crystals were obtained and subsequently determined as [DAPEDA]InCl6·Cl·H2O ( 1) by single-crystal X-ray diffraction (SCXRD). The crystals were then washed with ethanol, dried, and stored in a vacuum (65% yield based on In). Elemental analysis was performed and calculated: C, 17.18%; H, 5.05%; N, 10.02%. Found: C, 17.39%; H, 4.99%; N, 10.14%. Also, [email protected][DAPEDA]InCl6·Cl·H2O was prepared via a similar method to compound 1 except with the addition of SbCl3 (1 mmol, 0.2281 g). The colorless crystals of [email protected][DAPEDA]InCl6·Cl·H2O were obtained with a yield of 61%. Mechanical solid-state grinding synthesis of microscale [DPA]3InCl6 (2) A mixture of InCl3 (0.5 mmol, 0.1106 g) and [DPA]Cl (1.5 mmol, 0.2050 g) was ground together using a mortar and pestle. A white microscale powder of [DPA]3SbCl6 was obtained and used for further studies. Solution growth of [DPA]3InCl6 (2) and [email protected][DPA]3InCl6 single crystals A mixture of InCl3·4H2O (0.5 mmol, 0.1466 g) and DPA (1 mmol, 0.1012 g) was dissolved in a mixed solution of HCl (38%) (4 mL), acetonitrile (4 mL), and acetone (4 mL), and the mixture was stirred for 20 min until forming a clear solution. Large-size colorless block crystals were obtained by slow evaporation of the solution at room temperature after a week. The crystals were washed with ethanol and dried in a vacuum oven with a yield of 52% based on InCl3·4H2O. Elemental analysis performed and calculated: C, 34.09%; H, 7.63%; N, 6.63%; Found: C, 35.18%; H, 7.59%; N, 6.65%. The synthetic procedure of [email protected][DPA]3InCl6 crystal was similar to that of [DPA]3InCl6 with the addition of SbCl3 (0.8 mmol, 0.1825 g) in the mixed solution. The colorless crystals of [email protected][DPA]3InCl6 were obtained in a yield of 48%. X-ray crystallography The single-crystal data of compounds 1 and 2 were collected on the Bruker Apex II CCD diffractometer (Bruker Corporation, Billerica, Massachusetts, U.S.) with Mo-Kα radiation (λ = 0.71073 Å) at room temperature. The crystal structures were solved by direct method and refined based on F2 using the SHELXTL-2018 program. All the non-hydrogen atoms were refined with anisotropic thermal parameters, and hydrogen atoms of organic molecules were positioned geometrically and refined isotropically. Structural refinement parameters of compounds 1-2 are summarized in Supporting Information Table S3, and important bond parameters are listed in Supporting Information Tables S4–S7. Powder X-ray diffraction The powder X-ray diffraction (PXRD) analysis was performed on Bruker D8 ADVANCE powder X-ray diffractometer (Bruker Corporation, Billerica, Massachusetts, U.S.) equipped with copper Kα radiation at a voltage of 40 kV and a current of 40 mA. The diffraction patterns were scanned over an angular range of 5–60° (2θ) with a step size of 5°/min at room temperature. A simulated powder pattern was calculated employing the Mercury 3.8 software ( https://mercury1.software.informer.com/3.8/) using the crystallographic file from a SCXRD experiment. Scanning electron microscopy measurements The morphology observation and elemental mapping were conducted by scanning electron microscopy (SEM; Zeiss Merlin Compact, Carl Zeiss AG, Oberkochen, Germany). Inductively coupled plasma measurements Inductively coupled plasma optical emission spectroscopy (ICP-OES) was performed on Agilent 5110 (Agilent Technologies Inc. California, U.S.). Thermogravimetric analysis Thermogravimetric analysis (TGA) was carried out using a Q500 SDT system (TA Instruments, Pennsylvania, U.S.). The sample was heated from room temperature to 800 °C at a rate of 5 °C·min−1 under an argon flux of 40 mL·min−1. Absorption spectrum measurements The solid-state UV–vis absorption optical spectrum for powder sample was collected using a PE Lambda 900 UV/Vis spectrophotometer (Perkin-Elmer Optoelectronics, Massachusetts, U.S.) at room temperature in the wavelength range of 200–800 nm with BaSO4 serving as a reference standard. Raman measurements The Raman measurement was performed on the powder sample in the range of 0–3500 cm−1 using Horiba Scientific LabRam HR Evolution (HORIBA Scientific, Paris, France) under 532 nm excitation wavelength. Photoluminescence property characterization The photoluminescence (PL) spectrum was performed on an Edinburgh FLS980 fluorescence spectrometer (Edinburgh Instruments, Livingston, Scotland, UK). PLQY was achieved by incorporating an integrating sphere into the FLS980 spectrofluorometer, and Xe-lamp (450 W) was utilized to generate a monochromatic light source with an excitation power of 600 μW/cm2. PLQY was calculated based on the equation: ηQE = IS/(ER − ES), where IS represents the luminescence emission spectrum of the sample, ER is the spectrum of the excitation light from the empty integrated sphere (without the sample), and ES is the excitation spectrum for exciting the sample. The time-resolved decay data were carried out using the PL spectrofluorometer with a picosecond pulsed diode laser. The average lifetime was obtained by exponential fitting. Theoretical band calculation The single-crystal data were directly used to calculate the electronic band structure using the CASTEP software ( http://www.castep.org/CASTEP/GettingCASTEP). The total energies were calculated using the density functional theory (DFT) using Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation. The interactions between the ionic cores and the electrons were described by the norm-conserving pseudopotential. Hence, the C-2s22p2, N-2s22p3, H-1s1, In-5s25p2, and Cl-3s2sp5 orbitals were adopted as valence electrons. The number of plane wave included in the basis sets were determined by cut-off energy of 340 eV. Other calculating parameters and convergence criteria were set by the default values of the CASTEP code. Fabrication of WLED device The WLED was fabricated by integrating the Sb-doped compounds 1 or 2, the commercial red phosphor of K2SiF6∶Mn4+, and the blue phosphor BaMgAl10O17∶Eu2+ on an UV LED chip (365 nm). The phosphors were thoroughly mixed with epoxy resin, and the mixture obtained was coated on the LED chip. The optoelectronic properties, including electroluminescence (EL) spectrum, correlated color temperature (CCT), color rendering index (CRI), and Commission Internationale de l’Eclairage (CIE) color coordinates of the WLED were collected using an integrating sphere spectroradiometer system (PCE-2000B; EVERFINE Corporation, Hangzhou, China). Results and Discussion Colorless bulk single crystals of [DAPEDA]InCl6·Cl·H2O ( 1) and [DPA]3InCl6 ( 2) were prepared using facile wet-chemistry methods (Figure 1a). Specifically, we performed gradual evaporation of saturated HCl solution containing precursor InCl3·4H2O and corresponding organic salts. After the successful growth of single crystals and determination of crystal structure, we attempted the solid-state reaction route considering that solvent-free mechanochemical reaction is more economically applicable with multiple advantages, including low pollution, high production yield, simplified assembly technology, short reaction time, and reduced energy consumption, compared with the conventional wet-chemistry synthetic approach. As is known, conventional wet-chemistry synthetic techniques inescapably involve the massive use of hazardous organic solvents in assembly, separation, and purification processes, which are harmful to the environment and cause pollution to water bodies and soil. Based on these considerations, prepared the title compounds using a mechanochemical grinding solid-state reaction route. As indicated in Figure 1d, compound 2 was obtained through the mechanochemical grinding solid-state reaction from precursor materials within a fast time of about 60 s and 100% yield; the PXRD pattern was consistent with the simulated result from the crystal structure ( Supporting Information Figure S1). Compared with the demand precise controlling technique, tedious reaction process, and low-yields of wet-chemistry preparation avenues, this solvent-free mechanochemical grinding reaction method was highly cost-effective, simple, and time-saving. The facile and versatile preparation methods used provided sufficient assembly feasibility for optoelectronic device engineering. Figure 1 | Normal characterizations of compounds 1 and 2: (a) The solution evaporation synthetic route, (b) structure of [InCl6]3− unit, and (c) packing crystal structure of 1 along the b-axis. (d) The mechanochemical grinding solid-state synthetic route, (e) structure of [InCl6]3− unit, and (f) the crystal structure of compound 2 along the b-axis. Download figure Download PowerPoint The structures of compounds 1 and 2 were determined by using SCXRD. Compound 1 crystallizes in the monoclinic space group (P21/c), and the structure contains isolated [InCl6]3− octahedrons surrounded by [DAPEDA]4+ cations with dissociative Cl− ions as charge-balanced units (Figure 1b). The long-chain-like [DAPEDA]4+ cations feature closely paralleled stacking via weak molecular forces to form insulating organic walls to separate the isolated [InCl6]3− octahedrons with abundant N–H···Cl bonding interactions (H···Cl: 2.7–3.6 Å) (Figure 1c). Compound 2 belongs to the trigonal system (R3c), and the structure is composed of [InCl6]3− octahedrons and [DPA]+ cations (Figure 1e). The [DPA]+ cations feature a chiral column-like arrangement along the c-axis, resulting in cabined quasi-one-dimensional tunnels among them to enclose the parallel [InCl6]3− units (Figure 1f). In both compounds 1 and 2, all the [InCl6]3− octahedrons displayed parallel arrangement with closest In···In distance of 7.336 and 9.163 Å, respectively, circumventing the electron interactions between neighboring octahedrons.27 Therefore, these compounds belong to perfect 0D hybrid perovskites with [InCl6]3− units periodically embedding in the wide bandgap organic matrix to exhibit the intrinsic photophysical properties of individual species. The high purities of bulk crystals for compounds 1 and 2 are confirmed by the coincidence of measured PXRD patterns with simulated results from single-crystal data without any visible impurities ( Supporting Information Figure S2). The chemical compositions of the single crystals were checked by semi-quantitative energy-dispersive X-ray spectroscopy (EDX), and the average In:Cl mole ratios of 1:7.4 and 1:6.3 for compounds 1 and 2 are closing to the refined results obtained from the SCXRD, respectively ( Supporting Information Figures S3 and S4). At the same time, elemental mappings by SEM confirmed that all the C, N, In, and Cl elements were uniformly distributed on the surface of bulk crystals 1 and 2. These characterizations indicated high order crystallizations with unified crystal structures and chemical compositions for the bulk crystals 1 and 2. Supporting Information Figure S5 shows the wavelength-dependent solid-state UV–vis absorption spectra; the results indicate that both compounds 1 and 2 display steep absorption edges with almost identical cut-off wavelengths at 260 nm and a strong excitonic absorption peak around 330 nm. In addition, the broad absorption bands in the range of 260–310 and 350–500 nm were assigned to the electron transition within the organic salts ( Supporting Information Figure S6). Under 254 nm UV lamp irradiation, both the bulk crystals of compounds 1 and 2 displayed bright green light observed by the naked eye (Figures 2a and 2b). We further investigated the PL properties of compounds 1 and 2 using steady-state PL excitation and emission spectra, as well as time-resolved PL decay dynamics; the essential photophysical parameters are summarized in Table 1. The PL excitation spectra of both compounds show broad structured band characterizations with two dominant peaks at 262 nm, 324 nm for compound 1, and 271 nm, 324 nm accompanied by two shoulder weak excitation peaks of 284 and 340 nm for compound 2 ( Supporting Information Figure S7). The main excitation spectra were almost consistent with the absorption spectra preliminarily, indicating that the consequent PL emissions originated from the intrinsic bulk crystals and not from surface defects.28,29 Upon the 264 nm UV light excitation, compound 1 exhibited a prominent broadband green light emission centered at 520 nm with a full-width at half-maximums (FWHM) of 116 nm and a large Stokes shift of 256 nm (Figure 2c). Similarly, compound 2 gave a single broadband green emission (400–700 nm) centered at 510 nm with FWHM of 108 nm and Stokes shift of 240 nm (Figure 2d). The perfect Gaussian-shaped broadband emission spectra demonstrated the single optical center in compounds 1 and 2. The PL emission wavelengths and FWHM were close to those of low-dimensional lead halides such as [bmpy]9[ZnCl4]2[Pb3Cl11] (512 nm), [C9NH20]6Pb3Br12 (522 nm), and hybrid tetrahedral manganese halides ( Supporting Information Table S1).28,30,31 The corresponding CIE chromaticity coordinates were calculated to be (0.29, 0.47) and (0.25, 0.44) for compounds 1 and 2, respectively (Figure 2e). The slightly different PL emission linewidths of compounds 1 and 2 can be interpreted by the distinct distortion levels of [InCl6] octahedrons. In the [InCl6] unit of compound 1, the In–Cl bond distances and Cl–In–Cl bond angles fell in variable ranges of 2.489(2)–2.547(2) Å and 87.37(6)–91.73(6)° with a larger distortion of Δd = 0.7177 × 10−4 and σ2oct = 1.7807, respectively, whereas the [InCl6] unit of compound 2 belonged to a regular octahedron with entirely equivalent In–Cl distances of 2.5215(5) Å without any distortion (Figures 1b and 1e). The larger octahedral distortion level resulted in broader PL emission and larger Stokes shift of compound 1; such phenomenon has been confirmed in previously reported low-dimensional hybrid metal perovskites.32,33 The large Stokes shifts indicated the negligible self-absorption characterizations, which is of great interest for scintillators and luminescent solar concentrators.34–36 Figure 2 | The PL characterizations of compounds 1 and 2 at 300 K. (a and b) Photo images of bulk crystals of 1 (up) and 2 (down) under ambient light and UV light (254 nm); (c and d) excitation (purple) and emission (green) spectra of 1 (up) and 2 (down); (e) CIE chromaticity coordinates; (f and g) PL decay curves of 1 (up) and 2 (down). Download figure Download PowerPoint Table 1 | Summary of Photophysical Properties for All the Prepared Samples in This Work at 300 Ka λex (nm) λem (nm) CIE Stokes Shift[nm] FWHM [nm] Φ (%) τav (μs) Kr(S−1) × 106 Kn(S−1) × 106 Com- 1 264 520 0.29,0.47 256 116 40.40 3.3150 0.1219 0.1798 Sb @1 340 530 0.32,0.51 190 114 89.29 4.9495 0.1804 0.0216 Com- 2 270 510 0.25,0.44 240 108 34.01 3.1819 0.1069 0.2074 Sb @2 364 520 0.28,0.47 156 113 85.84 4.4500 0.1929 0.0318 aΦ is the PLQY and τav is the average PL lifetime, Kr and Kn are the radiation and nonradiation recombination rates, respectively. To verify the green light emission efficiencies of compounds 1 and 2, we performed the excitation wavelength-dependent PLQYs at 300 K. The results showed the varied PLQY values along with the change of excitation wavelength, in which the highest values are approximately corresponding to the highest excitation wavelength ( Supporting Information Figure S9). Significantly, the highest PLQYs of 40.40% and 34.01% for compounds 1 and 2 far exceeded those of hybrid lead halides, such as [C9NH20]6Pb3Br12 (12%), (bmpy)9(ZnBr4)2[Pb3Br11] (8.1%), and so on ( Supporting Information Figure S8 and Table S1).28,37 It is noteworthy that these PLQYs at room temperature are in the top rank of values, even when compared with some high-performance hybrid tetrahedral manganese halides such as [PP14]MnBr4 (55%), [Bu4N]2[MnBr4] (54%), and so on.16–19,38 To the best of our knowledge, compounds 1 and 2 represent the first green light-emitting hybrid indium perovskites with promising PLQYs, providing a new structural platform to explore efficient down-converting green phosphor. Considering the fact that almost identical excitation and emission spectra derive solely from the isolated [InCl6]3− optical centers, the slightly different PLQYs of compounds 1 and 2 should originate from the difference of surrounding environments around [InCl6]3− species. It has been reported that the motion freedoms of optical species play an essential role in the modulation of proportion between radiative and nonradiative transitions; specifically, suppressing the motion freedom of optical species could improve the PLQY by decreasing the nonradiative transition, especially for the low-dimensional hybrid metal perovskites.39–42 Careful examination of single-crystal structures demonstrate more abundant weak hydrogen bonding interactions between [InCl6]3− and organic species in compound 1, which rigidify the crystal lattice and restrict the molecular motion freedom ( Supporting Information Figure S10 and Tables S4 and S6). Further, we performed a qualitative analysis of the weak Cl···H interactions between anionic and cationic species by Hirshfeld surface calculation to analyze the intermolecular interactions within the crystal structure by using CrystalExplorer software.43 As shown in Supporting Information Figure S11, the 2D fingerprint plots show a more extensive color area verifying the stronger Cl···H interactions in compound 1 than 2. As a result, the rigid crystal lattice and inhibited motion freedom in compound 1 reduced the thermal vibrations of [InCl6]3− species and suppressed the nonradiative transition, further leading to higher PLQY. To reveal the photophysical process of broadband green light emissions, we performed time-resolved PL spectra, monitoring the maximum emission positions at 300 K. F
最长约 10秒,即可获得该文献文件

科研通智能强力驱动
Strongly Powered by AbleSci AI
更新
大幅提高文件上传限制,最高150M (2024-4-1)

科研通是完全免费的文献互助平台,具备全网最快的应助速度,最高的求助完成率。 对每一个文献求助,科研通都将尽心尽力,给求助人一个满意的交代。
实时播报
赖道之发布了新的文献求助10
1秒前
文右三发布了新的文献求助10
2秒前
wanci应助淡淡的大雁采纳,获得10
2秒前
xxxxxXPoVo发布了新的文献求助10
2秒前
Akim应助程同学采纳,获得10
2秒前
周期永在发布了新的文献求助10
2秒前
Mary发布了新的文献求助10
2秒前
Maestro_S应助林弋采纳,获得10
3秒前
ysr发布了新的文献求助10
4秒前
4秒前
Autor完成签到,获得积分10
5秒前
5秒前
5秒前
6秒前
一杯美事发布了新的文献求助10
6秒前
6秒前
slp完成签到,获得积分20
6秒前
cctv18应助危机的芸采纳,获得30
8秒前
HYY发布了新的文献求助10
8秒前
pK发布了新的文献求助10
10秒前
drunkprogrammer完成签到,获得积分10
10秒前
10秒前
shennan发布了新的文献求助10
10秒前
ss完成签到 ,获得积分10
11秒前
清爽的碧空完成签到,获得积分10
11秒前
我是老大应助生动元姗采纳,获得10
12秒前
CodeCraft应助小太阳采纳,获得10
13秒前
善良小夏完成签到,获得积分10
13秒前
13秒前
小王不咕噜~完成签到,获得积分10
14秒前
昭昭完成签到,获得积分20
15秒前
15秒前
曾梦完成签到,获得积分20
16秒前
deluohaida关注了科研通微信公众号
16秒前
可爱的函函应助的订单采纳,获得10
16秒前
17秒前
ykveu完成签到,获得积分10
17秒前
昭昭发布了新的文献求助10
18秒前
18秒前
222123发布了新的文献求助10
18秒前
高分求助中
Thermodynamic data for steelmaking 3000
Teaching Social and Emotional Learning in Physical Education 900
Counseling With Immigrants, Refugees, and Their Families From Social Justice Perspectives pages 800
藍からはじまる蛍光性トリプタンスリン研究 400
Cardiology: Board and Certification Review 400
[Lambert-Eaton syndrome without calcium channel autoantibodies] 340
New Words, New Worlds: Reconceptualising Social and Cultural Geography 300
热门求助领域 (近24小时)
化学 材料科学 医学 生物 有机化学 工程类 生物化学 纳米技术 物理 内科学 计算机科学 化学工程 复合材料 遗传学 基因 物理化学 催化作用 电极 光电子学 量子力学
热门帖子
关注 科研通微信公众号,转发送积分 2363533
求助须知:如何正确求助?哪些是违规求助? 2071916
关于积分的说明 5178028
捐赠科研通 1800044
什么是DOI,文献DOI怎么找? 898782
版权声明 557833
科研通“疑难数据库(出版商)”最低求助积分说明 479752