Next Article in Journal
Lower Ultra-High Frequency Non-Deployable Omnidirectional Antenna for Nanosatellite Communication System
Next Article in Special Issue
Optical Properties of Cylindrical Quantum Dots with Hyperbolic-Type Axial Potential under Applied Electric Field
Previous Article in Journal
Rapid Synthesis of Kaolinite Nanoscrolls through Microwave Processing
Previous Article in Special Issue
Substrate-Modulated Electric and Magnetic Resonances of Lithium Niobite Nanoparticles Illuminated by White Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ligand Engineering Triggered Efficiency Tunable Emission in Zero-Dimensional Manganese Hybrids for White Light-Emitting Diodes

1
Key Laboratory of Magnetic Molecules and Magnetic Information Materials (Ministry of Education), School of Chemistry and Material Science, Shanxi Normal University, Taiyuan 030031, China
2
Laboratory of Crystal Physics, Kirensky Institute of Physics, Federal Research Center KSC SB RAS, 660036 Krasnoyarsk, Russia
3
Research and Development Department, Kemerovo State University, 650000 Kemerovo, Russia
4
Department of Physics, Far Eastern State Transport University, 680021 Khabarovsk, Russia
5
Key Laboratory of Interface Science and Engineering in Advanced Material (Ministry of Education), College of Chemistry & Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(18), 3142; https://doi.org/10.3390/nano12183142
Submission received: 14 August 2022 / Revised: 26 August 2022 / Accepted: 7 September 2022 / Published: 10 September 2022
(This article belongs to the Special Issue Optical Properties of Semiconductor Nanomaterials)

Abstract

:
Zero-dimensional (0D) hybrid manganese halides have emerged as promising platforms for the white light-emitting diodes (w-LEDs) owing to their excellent optical properties. Necessary for researching on the structure-activity relationship of photoluminescence (PL), the novel manganese bromides (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are reported by screening two ligands with similar atomic arrangements but various steric configurations. It is found that (C13H14N)2MnBr4 with planar configuration tends to promote a stronger electron-phonon coupling, crystal filed effect and concentration-quenching effect than (C13H26N)2MnBr4 with chair configuration, resulting in the broadband emission (FWHM = 63 nm) to peak at 539 nm with a large Stokes shift (70 nm) and a relatively low photoluminescence quantum yield (PLQY) (46.23%), which makes for the potential application (LED-1, Ra = 82.1) in solid-state lighting. In contrast, (C13H26N)2MnBr4 exhibits a narrowband emission (FWHM = 44 nm) which peaked at 515 nm with a small Stokes shift (47 nm) and a high PLQY of 64.60%, and the as-fabricated white LED-2 reaches a wide colour gamut of 107.8% National Television Standards Committee (NTSC), thus highlighting the immeasurable application prospects in solid-state display. This work clarifies the significance of the spatial configuration of organic cations in hybrids perovskites and enriches the design ideas for function-oriented low-dimensional emitters.

1. Introduction

Low-dimensional organic-inorganic hybrid metal halides (OHMHs) have been widely recognized for their outstanding optical properties, including high-efficiency tunable emissions, near-unity photoluminescence quantum yields (PLQYs), and long decay lifetimes, etc., which are attributed to the strong quantum confinement at the molecular levels and the most convenient radiative recombination of photo-generated excitons [1,2,3,4]. However, the absorption peaks of most OHMHs are located in the ultraviolet (UV) region, which limits the luminous efficiency of white light-emitting diodes (w-LEDs) because they cannot be excited by commercial blue chips, thus hindering their industrialization in the fields of solid-state lighting and display. Remarkably, the zero-dimensional (0D) hybrid manganese halides become the preference for blue-light-excited luminescent materials since their photoluminescence excitation (PLE) bands lie in the near-ultraviolet and blue regions, while exhibiting bright photoluminescence (PL) [5,6], and they possess abundant coordination modes and tunable emissions [7,8,9,10,11]. Whether the octahedral coordinated Mn2+ ions show intense broadband orange-red emission from the [Mn3X12]6− inorganic units [12,13,14,15], or the tetrahedral coordinated Mn2+ ions exhibit narrowband green emission from the [MnX4]2− inorganic units, both of which are attributed to the d-d transition of Mn2+ and present application potential in the field of w-LEDs [16,17,18,19].
The inorganic units are embedded in organic matrix for OHMHs and are connected to the cations through various molecular interactions, prompting that their structure and PL behaviours can be tuned by engineering the organic cations [7,20]. Different organic cations induce the deformation of inorganic units in varying degrees, resulting in varying crystal field effects, nephelauxetic effects and concentration-quenching effects [21]. Numerous research has exposed that the effect of organic cations on the luminescence properties in 0D hybrid manganese halides. Seshadri et al., and Xia et al., investigated the relationship between structure and PL properties in hybrid manganese halides, and proposed that bulky, rigid, single-protonated cations are in favour of large Mn–Mn distance, thereby leading to high PLQYs [17,22]. Xiao et al., reported and highlighted the effect of organic components with different degrees of conjugation on the optical properties from the view of the band alignment types involving ground state and excited state [7]. Kovalenko et al., found that the PL decay can be accelerated by introducing heavy atoms (e.g., iodine, bromine) in the second coordination sphere of Mn2+ [6]. However, these reports ignored the role of spatial configuration of organic components on the fine-tuning PL behaviours and the correspondence between organic components and optical properties of 0D hybrid manganese halides has not been completely established.
Herein, we screen two ligands with similar atomic arrangements but different spatial configurations to synthesize 0D manganese bromides (C13H14N)2MnBr4 and (C13H26N)2MnBr4 via a simple solution-phase crystallization method. They are crystallized in different space groups and composed of [MnBr4]2− twisted tetrahedron and organic matrix with planar and chair configuration, respectively. Experimental and theoretical researches indicate that the organic cation with planar configuration tend to induce stronger electron-phonon coupling, crystal filed effect and concentration-quenching effect than chair configuration in these 0D manganese bromides, thereby finely tuning PL behaviours, including emission bandwidth, peak position, Stokes shift, lifetimes and PLQYs, as well as the luminescence mechanism was discussed by combining the Tunabe-Sugano (T-S) energy diagram. The w-LEDs devices were fabricated using the above manganese bromides as green components; it is concluded that LED-1 based on (C13H14N)2MnBr4 with a high Ra of 82.1 is suitable for solid-state lighting, while LED-2 based on (C13H26N)2MnBr4 with a wide colour gamut of 107.8% NTSC holds unprecedented promising for use in solid-state displays. Our work demonstrates that the organic cations with different spatial configurations are able to trigger tunable emission in 0D Manganese hybrids, and the design principle provides a new idea for future research on white light-emitting materials.

2. Experimental

2.1. Materials and Preparation

N-Methyldiphenylamine (C13H13N, 98%, Aladdin, Shanghai, China), N,N-Dicyclohexylmethylamine (C13H25N, 98%, Aladdin, Shanghai, China), manganese bromide tetrahydrate (MnBr2·4H2O, 98%, Aladdin, Shanghai, China), hydrobromic acid (HBr, 40%, Aladdin, Shanghai, China) and ethanol (C2H5OH, 99.7%, Guangfu, Tianjin, China). MnBr2·4H2O needs to be heated at 120 °C for 6 h to remove the crystal water for later use.
The single-crystals of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 were synthesized by a simple solution-phase crystallization method. First, C13H13N (0.3665 g, 2 mmol) or C13H25N (0.3907 g, 2 mmol) was dissolved in 1.5 mL HBr for protonation. Then MnBr2 (0.2148 g, 1 mmol) and 5 mL C2H5OH were added to the above protonated solution, and heated and stirred at 75 °C until forming a clear solution. After the solution was naturally cooled to room temperature, yellow (C13H14N)2MnBr4 and green (C13H26N)2MnBr4 bulky crystals were precipitated overnight. Finally, the crystals were washed several times with acetone and dried in a vacuum oven.

2.2. Characterization

Single-crystal X-ray diffraction (SCXRD) data of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 were collected by using an Agilent Technologies Gemini EOS (Palo Alto, CA, USA) diffractometer at 298 K using Mo Kα radiation (λ = 0.71073 Å). Powder X-ray diffraction (PXRD) data of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 for Rietveld analysis was collected at room temperature with a Bruker D8 ADVANCE powder diffractometer EOS (Karlsruhe, Germany) (Cu-Kα radiation) and linear VANTEC detector. The step size of 2θ was 0.011°, and the counting time was 2 s per step. Rietveld structure refinements were performed by using TOPAS 4.2. PL, PLE spectra, PL decay curves and PLQYs were measured by an Edinburgh FLS920 fluorescence spectrometer EOS (Edinburgh, UK) with a picosecond pulsed diode laser. Temperature-dependent emission spectra were measured on the same spectrophotometer installed with a heating apparatus as the heating source. Morphology observation and elemental mappings were conducted by a scanning electron microscope (SEM, JEOL JSM-6510, EOS, Peabody, MA, USA). UV-vis absorption curves were recorded on a TU-1901 Ultraviolet spectrometer EOS (Beijing, China) at room temperature, in which BaSO4 was used as the standard reference. CIE chromaticity coordinates were calculated using the CIE calculator software based on the emission spectra excited at 450 nm. The emission spectra, correlated colour temperature (CCT), luminous efficacy, and CIE coordinates of w-LEDs were performed on the integrating sphere spectroradiometer system (ATA-100, Everfine, EOS, Hangzhou, Cina).

2.3. Computational Methods

The electronic band structure and density of state (DOS) were calculated by CASTEP based on plane-wave pseudopotential density functional theory (DFT) [23]. Perdew-Burke-Ernzerhof (PBE) functionals in the form of general gradient approximation (GGA) were used for electronic structure calculations [24]. A kinetic energy cut off value of 450 eV and a Monkhorst-pack k-point mesh spanning less than 0.03 Å−1 in the Brillouin zone were chosen.

3. Results and Discussion

As shown in Figure 1a,d, two ligands with similar atomic arrangements but different spatial configurations were screened out to synthesize hybrid manganese bromides (C13H14N)2MnBr4 and (C13H26N)2MnBr4 with MnBr2 via a simple solution-phase crystallization method. The crystal structures resolved by SCXRD technique show that both (C13H14N)2MnBr4 and (C13H26N)2MnBr4 correspond to typical 0D “host-guest” structures. That is, central metal Mn2+ ions are coordinated with four Br anions to form twisted tetrahedral [MnBr4]2− inorganic units and all inorganic components are periodically dispersive embedded in insulating organic cations (Figure 1c,f). Different organic cations induce the distinct packing and deformation of anions. Therefore, (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are crystallized in triclinic space group P1 and monoclinic space group P21/c, respectively. As shown in Figure 1b,e, two different [MnBr4]2− inorganic units exhibit different distortion degree. The distances of four Mn–Br bonds are in the range of 2.4829(13)–2.5201(13) Å and the Br–Mn–Br bond angles ranged from 103.19(5) to 117.83(6)° for (C13H14N)2MnBr4. However, this structural information is in the range of 2.4704(9)–2.5226(9) Å and 107.07(3) to 114.21(4)° for (C13H26N)2MnBr4. The crystallographic information files (CIFs) are shown in the Supporting Information (SI), and the main bond lengths and bond angles are shown in Table S1 and Table S2, respectively. The bond length distortion (∆d) and bond angle variance (σ2) of individual [MnBr4]2− are calculated by the following formulas [25,26]:
Δ d = 1 / 4 i = 1 4 d i d 0 / d 0 2
where d 0 represents the average Mn–Br bond length and d i are four individual lengths of Mn–Br bond.
σ 2 = 1 / 5 i = 1 6 θ i 109.47 2
where θ i refer to the individual Br–Mn–Br angles. The Δ d values of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are 1.78 × 10−4 and 2.39 × 10−4, respectively. The σ 2 values are 24.64 and 6.08. It is worth noting that the difference in bond length distortion is small and can be ignored, since the large distinction in the bond angle variance maybe the underlying reason for the disparate optical properties of (C13H14N)2MnBr4 and (C13H26N)2MnBr4.
The phase purity and crystallinity of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 powders were monitored by PXRD, and the results are shown in Figure S1. All peaks were indexed by triclinic cell (P1) and monoclinic cell (P21/c) with parameters close to those obtained from a single crystal experiment, respectively. Therefore, these structures were considered as a starting model for Rietveld refinement, as shown in Figure 2a,b, which were performed using TOPAS 4.2. The refinement results were stable and gave low R-factors. The main parameters of processing and refinement of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 were listed in Table S3. The elemental mapping images (Figure 2c,d) indicate that the elements N, Br and Mn are evenly distributed on the above manganese bromides.
As shown in Figure S2b,c, the manganese bromides of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 with yellow-green body colour exhibit bright yellow-green fluorescence under a 365 nm UV lamp, and the corresponding CIE coordinates are depicted in Figure S2a. To further reveal their PL properties, the PLE and PL spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are initially investigated in Figure 3a,b. Both manganese bromides exhibit similar excitation bands attributed to the d–d transition of tetrahedrally coordinated Mn2+ centres and can be excited by blue light, but they possess distinct emission spectra. (C13H14N)2MnBr4 displays a broadband emission peaked at 539 nm with a full width at half maximum (FWHM) of 63 nm and the Stokes shift of 70 nm, which is wider than most 0D hybrid manganese halides [19,27,28]. In contrast, the emission peak of (C13H26N)2MnBr4 appears at 515 nm, with a FWHM of 44 nm and the Stokes shift of 47 nm. Meanwhile, we synthesized the powder chlorides of (C13H14N)2MnCl4 and (C13H26N)2MnCl4 using the above two ligands and they exhibit the same effect and difference on their PL properties as (C13H14N)2MnBr4 and (C13H26N)2MnBr4 (Figure S3). We consider that the essence for the PL difference may be caused by the different spatial configuration of organic components in above manganese bromides. The ligand C13H13N contains two planar benzene rings with little steric hindrance. However, the twelve carbon atoms are not in the same plane for C13H25N, so that there is a larger steric hindrance. Proceeding from (C13H14N)2MnBr4 to (C13H26N)2MnBr4, the increasing cation volume and the steric hindrance not only shrink the free space for the atom movement or the cation/anion rotation, but also inhibit the lattice distortion [29,30]. Therefore, (C13H14N)2MnBr4 produce stronger electron-phonon coupling and crystal field strength than (C13H26N)2MnBr4, resulting in a broadband emission with large Stokes shift.
Figure 3c depicts the temperature-dependent PL spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 in the range of 80–300 K. The PL intensity decreases with the increasing temperature, which is consistent with the PL quenching behaviour caused by thermally activated non-radiative recombination. Meanwhile, the emission peak positions have a blue-shift. It is speculated that the reason for this phenomenon is that the thermally induced lattice expansion weakens the crystal field strength, or reduces the energy loss due to the spin-spin coupling between localized neighbouring Mn2+ ions [16]. Furthermore, the FWHM versus temperature (Figure 3d) can reveal the origin of the PL differences in above manganese bromides. The Huang-Rhys factor S defines the degree of electron–phonon coupling; it can be solved by the following equation [18,31,32]:
FWHM = 2.36 ħ ω S   c o t h ħ ω 2 k T    
where ω is the phonon frequency, ħ ω is the maximum phonon energy, k is the Boltzmann constant, S is Huang-Rhys factor, and c o t h x = e x + e x e x e x = 1 + 2 e 2 x 1 . When ħ ω k T is small enough, e ħ ω k T 1 ħ ω k T , it can be obtained the following equation:
FWHM 2 = 5.57 ħ ω 2 S 1 + 2 e ħ ω k T 1 = 5.57 ħ ω 2 S 1 + 1 ħ ω 2 k T
Further written as:
FWHM 2 = a + b 1 2 k T
where a = 5.57 × S × ( ħ ω )2 and b = 5.57 × S × ( ħ ω ).
The obtained S factor is 2.89 and phonon energy ħ ω phonon is 44.26 meV for (C13H14N)2MnBr4, while for (C13H26N)2MnBr4, it corresponds to S = 0.77, ħ ω = 75.85 meV. The higher S factor indicates that there is a stronger electron-phonon coupling in (C13H14N)2MnBr4, which is favourable for the formation of broadband emission with large Stokes shift.
In addition, the crystal field strength plays a key role in affecting the PL properties of manganese bromides. Previous references reported that the crystal field strength is related to the polyhedral distortion, and the increasing distortion leads to strong crystal field splitting and low position of the lowest 3d excited energy levels, resulting in the red-shift of the emission band [33,34,35,36]. The σ 2 values of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are 24.64 and 6.08, respectively, as calculated above, indicating that the larger bond angle distortion contributes to a red-shift of the emission peak from 515 nm to 539 nm.
To verify the electronic transition process, we measured the lifetimes excited at 450 nm and monitored at 539 nm and 515 nm, respectively. As shown in Figure 3e,f, the PL decay curves can be fitted with a single order exponential equation:
I t = I 0 + A   exp   t / τ
where I(t) and I0 are the luminescence intensity at time t and tτ, respectively. A is a constant, and τ is the decay time for an exponential component. At 298 K, the lifetimes are on millisecond scale, and the values are determined to be 0.245 ms and 0.370 ms, respectively, which are close to the previously reported lifetimes of hybrid manganese bromides, demonstrating that these emissions belong to the d–d transition (4T16A1) of Mn2+ [17,37,38]. At 80 K, the lifetimes are prolonged, but the variation tendency of (C13H14N)2MnBr4 is greater than that of (C13H26N)2MnBr4, which may be because that (C13H14N)2MnBr4 has more vibration options. In addition, PLQY is also affected by the spatial configuration of organic cations in these manganese bromides. Figure S4 shows that the [C13H14N]+ cations containing benzene ring planar configuration generate a short average distance between the nearest [MnBr4]2− is 8.5934 Å, while the chair configuration of [C13H26N]+ cations with greater steric hindrance produce a longer Mn–Mn average distance is 9.6703 Å. Accordingly, the concentration-quenching effect of (C13H26N)2MnBr4 is weakened, resulting in PLQYs of 46.23% and 64.60% for (C13H14N)2MnBr4 and (C13H26N)2MnBr4, respectively.
To further reveal the photo-physical properties of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 composed of organic moieties with different spatial configurations. The theoretical calculations involving electronic band structure and densities of states (DOS) are conducted based on the density functional theory. As shown in Figure 4a,b, the calculated band gap of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are 2.54 eV and 2.80 eV, respectively, which are coincide with the observed optical absorption spectra (Figure S5). It is worth noting that the (C13H14N)2MnBr4 has a lower conduction band level compared to (C13H26N)2MnBr4, resulting from the higher degree of conjugation of organic component, which is also responsible for the red-shift of the emission peak of (C13H14N)2MnBr4. Both the frontier orbitals are flat and dispersion-less, indicating that the photoelectrons are localized for above manganese bromides. This is because that the isolated [MnBr4]2− tetrahedrons separated by bulky organic cations have weak interactions, resulting in spatial confinement and electronic confinement effects. Figure 4c,d show the DOS projected onto the constituent elements in (C13H14N)2MnBr4 and (C13H26N)2MnBr4. The valence band maximum (VBM) states and the conduction band minimum (CBM) states are mainly composed of Mn 3d and Br 3p orbitals, indicating that the electronic transition process related to luminescence occurs within the [MnBr4]2− inorganic units, and the VBM- and CBM-associated charge densities (Figure S6) also confirm this mechanism.
According to the electron transition process, we employ the T-S energy diagram to visualize the effect of spatial configuration of organic ligands on PL behaviours. It is well known that the Mn2+ ions possess the outermost electron configuration of 3d5, and the original five degenerate d orbitals of free Mn2+ are split into multiple energy levels under the electrostatic field effect generated by ligands. In the tetrahedral field, the ground state is denoted as 6A1, the excited states of 4F, 4P, 4D and 4G correspond to the excited energy levels in the UV (200–400 nm) and blue (400–500 nm) regions, respectively [6,7]. The electrons located at the ground state are transferred to different excited states under excitation, and then transferred to the lowest excited state energy level 4T1 through non-radiative relaxation, and finally the electrons return to the ground state with the release of photons. In the T-S energy diagram, the strength of the crystal field is represented by the abscissa (∆) and the energy difference between excited states 4T2 and 4T1. The stronger the crystal field strength, the more obvious the splitting of 4T2 and 4T1, making the larger values of E [4T2]–E[4T1]. In view of this, the luminescent mechanism is depicted in Figure 5. (C13H14N)2MnBr4 and (C13H26N)2MnBr4 are arranged in the region of strong crystal field and weak crystal field, respectively. (C13H14N)2MnBr4 occupies a wider range on the abscissa than (C13H26N)2MnBr4, due to the larger electron-phonon coupling coefficient for (C13H14N)2MnBr4, resulting in the crystal field strength varies greatly with lattice vibrations. Reflected on the vertical coordinates, it is shown that (C13H14N)2MnBr4 has a larger ∆E than (C13H26N)2MnBr4, that is, (C13H14N)2MnBr4 exhibits a broadband emission with a large Stokes shift.
To further evaluate their practical applications, the w-LEDs (LED-1 and LED-2) were encapsulated by coating (C13H14N)2MnBr4/(C13H26N)2MnBr4 and KSF:Mn4+ red phosphors on a commercial InGaN blue chip (450 nm), respectively. As shown in Figure 6, LED-1 presents the standard white-light emission with a relatively high colour-rendering index (CIR) of Ra = 82.1, and a CIE chromaticity coordinate of (0.3211, 0.3206). LED-2 presents a cool white-light emission with a wide colour gamut of 107.8% NTSC in the CIE 1931 colour space, and a CIE chromaticity coordinate of (0.3037, 0.3297). It is concluded that (C13H14N)2MnBr4 is more suitable for solid-state lighting, while (C13H26N)2MnBr4 is unprecedentedly promising as a narrow-band green emitter for solid-state displays. The emission spectra of LED-1 and LED-2 under different drive current different current (20−120 mA) in Figure S7 suggests that the w-LEDs possess excellent colour stability. Of course, we also evaluated the thermal quenching behaviour of (C13H14N)2MnBr4 and (C13H26N)2MnBr4. As shown in Figure S8, the PL intensity decreases with increasing temperature. The PL intensity of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 at 373 K remain around 79.86% and 84.84% of that at room temperature, respectively, suggesting that the above manganese bromides have relatively good thermal stability.

4. Conclusions

In summary, the two new 0D manganese bromides were developed by using organic cations with different spatial configuration. (C13H14N)2MnBr4 exhibits a broadband emission peaked at 539 nm with a FWHM of 63 nm and a Stokes shift of 70 nm, while (C13H26N)2MnBr4 shows a narrowband emission peaked at 515 nm with a FWHM of 44 nm and a Stokes shift of 47 nm, which is ascribed to that the ligand of C13H13N with planar configuration induces a stronger electron-phonon coupling (S = 2.89) and crystal field strength ( σ 2 = 24.64) than the ligand of C13H25N with chair configuration. DFT calculations reveal that (C13H14N)2MnBr4 possesses a narrower bandgap compared to (C13H26N)2MnBr4, resulting in the red-shift of the emission peak of (C13H14N)2MnBr4. Moreover, the steric effect of organic cations on PL behaviours in manganese halides is comprehensively reflected in T-S energy diagram. The as-fabricated white LED-1 based on (C13H14N)2MnBr4 with a high CIR of Ra = 82.1 is suitable for solid-state lighting, while the as-fabricated white LED-2 based on (C13H26N)2MnBr4 with a wide colour gamut of 107.8% NTSC in the CIE 1931 colour space presents an unprecedentedly promising utility for solid-state display.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12183142/s1, Figure S1: PXRD patterns and standard diffraction pattern of (C13H14N)2MnBr4 and (C13H26N)2MnBr4; Figure S2: CIE chromaticity diagrams of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 and the photographs with UV off and on; Figure S3: Normalized PLE and PL spectra and time-resolved PL decay curves of (C13H14N)2MnCl4 and (C13H26N)2MnCl4 at 298 K; Figure S4: The schematic diagrams of Mn···Mn distance; Figure S5: Absorption spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4; Figure S6: VBM- and CBM-associated charge densities of (C13H14N)2MnBr4 and (C13H26N)2MnBr4; Figure S7: PL spectra of the w-LED devices under different current (20−120 mA); Figure S8: Temperature-dependent PL spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 ranging from 298 K to 423 K; Tables S1 and S2: The bond length (Å) and the bond angles (°) of (C13H14N)2MnBr4 and (C13H26N)2MnBr4; Table S3: Main parameters of processing and refinement of (C13H14N)2MnBr4 and (C13H26N)2MnBr4.

Author Contributions

Experiment, Q.R. and Y.M.; design, Q.R. and G.Z.; data curation, Q.R.; analysis, Q.R. and J.Z. and M.S.M.; writing—review and editing, Q.R. and G.Z.; funding acquisition, G.Z. and X.-M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Shanxi Province (No. 20210302124054), National Natural Science Foundation of China (No. 21871167), Science and Technology Innovation Project of Colleges and Universities in Shanxi Province (No. 2021L262), 1331 Project of Shanxi Province and the Postgraduate Innovation Project of Shanxi Normal University (No.2021XSY040), and funded by RFBR according to the research project No. 19−52−80003.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, L.; Liao, J.F.; Kuang, D.B. An overview for zero-dimensional broadband emissive metal-halide single crystals. Adv. Opt. Mater. 2021, 9, 2100544. [Google Scholar] [CrossRef]
  2. Chu, Z.; Chu, X.; Zhao, Y.; Ye, Q.; Jiang, J.; Zhang, X.; You, J. Emerging low-dimensional crystal structure of metal halide perovskite optoelectronic materials and devices. Small Struct. 2021, 2, 2000133. [Google Scholar] [CrossRef]
  3. Zhou, C.K.; Xu, L.-J.; Lee, S.; Lin, H.R.; Ma, B.W. Recent advances in luminescent zero-dimensional organic metal halide hybrids. Adv. Opt. Mater. 2020, 9, 2001766–2001782. [Google Scholar] [CrossRef]
  4. Li, M.Z.; Xia, Z.G. Recent progress of zero-dimensional luminescent metal halides. Chem. Soc. Rev. 2020, 50, 2626–2662. [Google Scholar] [CrossRef] [PubMed]
  5. Mark, W.; David, G. Excited state decay of tetrahalomanganese(ii) complexes. Chem. Phys. 1974, 4, 295–299. [Google Scholar]
  6. Morad, V.; Cherniukh, I.; Pöttschacher, L.; Shynkarenko, Y.; Yakunin, S.; Kovalenko, M.V. Manganese(II) in tetrahedral halide environment: Factors governing bright green luminescence. Chem. Mater. 2019, 31, 10161–10169. [Google Scholar] [CrossRef]
  7. Fu, P.; Sun, Y.; Xia, Z.; Xiao, Z. Photoluminescence Behavior of Zero-dimensional manganese halide tetrahedra embedded in conjugated organic matrices. J. Phys. Chem. Lett. 2021, 12, 7394–7399. [Google Scholar] [CrossRef]
  8. Bai, X.; Zhong, H.; Chen, B.; Chen, C.; Han, J.; Zeng, R.; Zou, B. Pyridine-modulated Mn ion emission properties of C10H12N2MnBr4 and C5H6NMnBr3 single crystals. J. Phys. Chem. C 2018, 122, 3130–3137. [Google Scholar] [CrossRef]
  9. Hu, G.; Xu, B.; Wang, A.; Guo, Y.; Wu, J.; Muhammad, F.; Meng, W.; Wang, C.; Sui, S.; Liu, Y.; et al. Stable and bright pyridine manganese halides for efficient white light-emitting diodes. Adv. Funct. Mater. 2021, 31, 2011191. [Google Scholar] [CrossRef]
  10. Li, C.; Bai, X.; Guo, Y.; Zou, B. Tunable emission properties of manganese chloride small single crystals by pyridine incorporation. ACS Omega 2019, 4, 8039–8045. [Google Scholar] [CrossRef]
  11. Sun, M.E.; Li, Y.; Dong, X.Y.; Zang, S.Q. Thermoinduced structural-transformation and thermochromic luminescence in organic manganese chloride crystals. Chem. Sci. 2019, 10, 3836–3839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zou, B.; Wang, S.; Han, X.; Kou, T.; Zhou, Y.; Liang, Y.; Wu, Z.; Huang, J.; Chang, T.; Peng, C. Lead-free MnII-based Red-emitting Hybrid Halide (CH6N3)2MnCl4 toward High Performance Warm WLEDs. J. Mater. Chem. C 2021, 9, 4895–4902. [Google Scholar]
  13. Zhou, G.; Liu, Z.; Molokeev, M.S.; Xiao, Z.; Xia, Z.; Zhang, X.-M. Manipulation of Cl/Br transmutation in zero-dimensional Mn2+-based metal halides toward tunable photoluminescence and thermal quenching behaviors. J. Mater. Chem. C 2021, 9, 2047–2053. [Google Scholar] [CrossRef]
  14. Sen, A.; Swain, D.; Guru Row, T.N.; Sundaresan, A. Unprecedented 30 K hysteresis across switchable dielectric and magnetic properties in a bright luminescent organic-inorganic halide (CH6N3)2MnCl4. J. Mater. Chem. C 2019, 7, 4838–4845. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Liao, W.Q.; Fu, D.W.; Ye, H.Y.; Chen, Z.N.; Xiong, R.G. Highly efficient red-light emission in an organic-inorganic hybrid ferroelectric: (Pyrrolidinium)MnCl3. J. Am. Chem. Soc. 2015, 137, 4928–4931. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, G.; Ding, J.; Jiang, X.; Zhang, J.; Molokeev, M.S.; Ren, Q.; Zhou, J.; Li, S.; Zhang, X.-M. Coordination units of Mn2+ modulation toward tunable emission in zero-dimensional bromides for white light-emitting diodes. J. Mater. Chem. C 2022, 10, 2095–2102. [Google Scholar] [CrossRef]
  17. Zhou, G.; Liu, Z.; Huang, J.; Molokeev, M.S.; Xiao, Z.; Ma, C.; Xia, Z. Unraveling the near-unity narrow-band green emission in zero-dimensional Mn2+-based metal halides: A case study of (C10H16N)2Zn1−xMnxBr4 solid solutions. J. Phys. Chem. Lett. 2020, 11, 5956–5962. [Google Scholar] [CrossRef]
  18. Zhang, S.; Zhao, Y.; Zhou, J.; Ming, H.; Wang, C.-H.; Jing, X.; Ye, S.; Zhang, Q. Structural design enables highly-efficient green emission with preferable blue light excitation from zero-dimensional manganese (II) hybrids. Chem. Eng. J. 2021, 421, 129886. [Google Scholar] [CrossRef]
  19. Su, B.; Molokeev, M.S.; Xia, Z. Mn2+-based narrow-band green-emitting Cs3MnBr5 phosphor and the performance optimization by Zn2+ alloying. J. Mater. Chem. C 2019, 7, 11220–11226. [Google Scholar] [CrossRef]
  20. Gong, L.K.; Hu, Q.Q.; Huang, F.Q.; Zhang, Z.Z.; Shen, N.N.; Hu, B.; Song, Y.; Wang, Z.P.; Du, K.Z.; Huang, X.Y. Efficient modulation of photoluminescence by hydrogen bonding interactions between inorganic [MnBr4]2− anions and organic cations. Chem. Commun. 2019, 55, 7303–7306. [Google Scholar] [CrossRef]
  21. Han, Y.; Yue, S.; Cui, B.B. Low-dimensional metal halide perovskite crystal materials: Structure strategies and luminescence applications. Adv. Sci. (Weinh) 2021, 8, e2004805. [Google Scholar] [CrossRef] [PubMed]
  22. Mao, L.; Guo, P.; Wang, S.; Cheetham, A.K.; Seshadri, R. Design principles for enhancing photoluminescence quantum yield in hybrid manganese bromides. J. Am. Chem. Soc. 2020, 142, 13582–13589. [Google Scholar] [CrossRef] [PubMed]
  23. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Krist. Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  24. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  25. Robinson, K.; Gibbs, G.V.; Ribbe, P.H. Maintenance of resting potential in anoxic guinea pig ventricular muscle: Electrogenic sodium pumping. Science 1971, 172, 567. [Google Scholar] [CrossRef]
  26. Zhang, L.; Luo, Z.; Wei, Y.; Wang, W.; Liu, Y.; Li, C.; He, X.; Quan, Z. Zero-dimensional hybrid binuclear manganese chloride with thermally stable yellow emission. Chem. Commun. 2022, 58, 6926–6929. [Google Scholar] [CrossRef]
  27. Fu, H.; Jiang, C.; Lao, J.; Luo, C.; Lin, H.; Peng, H.; Duan, C.-G. An organic-inorganic hybrid ferroelectric with strong luminescence and high Curie temperature. CrystEngComm 2020, 22, 1436–1441. [Google Scholar] [CrossRef]
  28. Xu, L.-J.; Lin, X.; He, Q.; Worku, M.; Ma, B. Highly efficient eco-friendly X-ray scintillators based on an organic manganese halide. Nat. Commun. 2020, 11, 4329. [Google Scholar] [CrossRef]
  29. Li, D.-Y.; Cheng, Y.; Hou, Y.-H.; Song, J.-H.; Sun, C.-J.; Yue, C.-Y.; Jing, Z.-H.; Lei, X.-W. Modulating photoelectron localization degree to achieve controllable photoluminescence quenching and activation of 0D hybrid antimony perovskites. J. Mater. Chem. C 2022, 10, 3746–3755. [Google Scholar] [CrossRef]
  30. Wang, Z.; Xie, D.; Zhang, F.; Yu, J.; Chen, X.; Wong, C.P. Controlling information duration on rewritable luminescent paper based on hybrid antimony (III) chloride/small-molecule absorbates. Sci. Adv. 2020, 6, eabc2181. [Google Scholar] [CrossRef]
  31. Brik, M.G.; Avram, N.M. Electron-vibrational interaction in the 5d states of Ce3+ ions in halosulphate phosphors. Mater. Chem. Phys. 2011, 128, 326–330. [Google Scholar] [CrossRef]
  32. Zhou, X.; Geng, W.; Li, J.; Wang, Y.; Ding, J.; Wang, Y. An ultraviolet-visible and near-infrared-responded broadband NIR phosphor and its NIR spectroscopy application. Adv. Opt. Mater. 2020, 8, 1902003. [Google Scholar] [CrossRef]
  33. Leng, Z.; Li, R.; Li, L.; Xue, D.; Zhang, D.; Li, G.; Chen, X.; Zhang, Y. Preferential neighboring substitution-triggered full visible spectrum emission in single-phased Ca10.5−xMgx(PO4)7:Eu(2+) phosphors for high color-rendering white LEDs. ACS Appl. Mater. Interfaces 2018, 10, 33322–33334. [Google Scholar] [CrossRef] [PubMed]
  34. Denault, K.A.; Brgoch, J.; Gaultois, M.W.; Mikhailovsky, A.; Petry, R.; Winkler, H.; DenBaars, S.P.; Seshadri, R. Consequences of optimal bond valence on structural rigidity and improved luminescence properties in SrxBa2−xSiO4:Eu2+ orthosilicate phosphors. Chem. Mater. 2014, 26, 2275–2282. [Google Scholar] [CrossRef]
  35. Ming, Z.; Qiao, J.; Molokeev, M.S.; Zhao, J.; Swart, H.C.; Xia, Z. Multiple substitution strategies toward tunable luminescence in Lu2MgAl4SiO12:Eu(2+) phosphors. Inorg. Chem. 2020, 59, 1405–1413. [Google Scholar] [CrossRef] [PubMed]
  36. Denault, K.A.; George, N.C.; Paden, S.R.; Brinkley, S.; Mikhailovsky, A.A.; Neuefeind, J.; DenBaars, S.P.; Seshadri, R. A green-yellow emitting oxyfluoride solid solution phosphor Sr2Ba(AlO4F)1−x(SiO5)x:Ce3+ for thermally stable, high color rendition solid state white lighting. J. Mater. Chem. 2012, 22, 18204. [Google Scholar] [CrossRef]
  37. Tan, L.; Luo, Z.; Chang, X.; Wei, Y.; Tang, M.; Chen, W.; Li, Q.; Shen, P.; Quan, Z. Structure and photoluminescence transformation in hybrid manganese(II) chlorides. Inorg. Chem. 2021, 60, 6600–6606. [Google Scholar] [CrossRef]
  38. Wang, N.; Xiong, Y.; Liu, K.; He, S.; Cao, J.; Zhang, X.; Zhao, J.; Liu, Q. Efficient narrow-band green light-emitting hybrid halides for wide color gamut display. ACS Appl. Electron. Mater. 2022, 4, 4068–4076. [Google Scholar] [CrossRef]
Figure 1. (a,d) The chemical structures of organic cations [C13H14N]+ and [C13H26N]+. (b,e) Ball and stick models of inorganic units and structural information including bond length, bond angles and twist degrees. (c,f) Crystal structures of (C13H14N)2MnBr4 and (C13H26N)2MnBr4.
Figure 1. (a,d) The chemical structures of organic cations [C13H14N]+ and [C13H26N]+. (b,e) Ball and stick models of inorganic units and structural information including bond length, bond angles and twist degrees. (c,f) Crystal structures of (C13H14N)2MnBr4 and (C13H26N)2MnBr4.
Nanomaterials 12 03142 g001
Figure 2. Rietveld refinements of PXRD patterns for (a) (C13H14N)2MnBr4, (b) (C13H26N)2MnBr4. SEM images and EDS elemental mapping images showing the homogeneous distribution of N, Mn and Br in compounds for (c) (C13H14N)2MnBr4 and (d) (C13H26N)2MnBr4.
Figure 2. Rietveld refinements of PXRD patterns for (a) (C13H14N)2MnBr4, (b) (C13H26N)2MnBr4. SEM images and EDS elemental mapping images showing the homogeneous distribution of N, Mn and Br in compounds for (c) (C13H14N)2MnBr4 and (d) (C13H26N)2MnBr4.
Nanomaterials 12 03142 g002
Figure 3. Normalized PLE and PL spectra of (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. (c) Temperature-dependent PL spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 under 450 nm excitation in the range from 80 to 300 K. (d) FWHM2 fitting curves of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 as a function of 1/(2kT). Time-resolved PL decay curves of (e) (C13H14N)2MnBr4 and (f) (C13H26N)2MnBr4 at 298 K and 80 K.
Figure 3. Normalized PLE and PL spectra of (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. (c) Temperature-dependent PL spectra of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 under 450 nm excitation in the range from 80 to 300 K. (d) FWHM2 fitting curves of (C13H14N)2MnBr4 and (C13H26N)2MnBr4 as a function of 1/(2kT). Time-resolved PL decay curves of (e) (C13H14N)2MnBr4 and (f) (C13H26N)2MnBr4 at 298 K and 80 K.
Nanomaterials 12 03142 g003
Figure 4. Electronic energy band structures of (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. The total and orbital projection of partial density of states of (c) (C13H14N)2MnBr4 and (d) (C13H26N)2MnBr4.
Figure 4. Electronic energy band structures of (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. The total and orbital projection of partial density of states of (c) (C13H14N)2MnBr4 and (d) (C13H26N)2MnBr4.
Nanomaterials 12 03142 g004
Figure 5. Tanabe–Sugano energy-level diagram of Mn2+ ions in tetrahedral crystal field.
Figure 5. Tanabe–Sugano energy-level diagram of Mn2+ ions in tetrahedral crystal field.
Nanomaterials 12 03142 g005
Figure 6. Emission spectrum of w-LEDs based on (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. The insets show the photographs of w-LEDs in the daylight (top) and working at 20 mA (bottom). (c) CIE 1931 colour coordinates of fabricated LED-1 (orange dotted line), LED-2 (red dotted line) and the NTSC standard (black dotted line).
Figure 6. Emission spectrum of w-LEDs based on (a) (C13H14N)2MnBr4 and (b) (C13H26N)2MnBr4. The insets show the photographs of w-LEDs in the daylight (top) and working at 20 mA (bottom). (c) CIE 1931 colour coordinates of fabricated LED-1 (orange dotted line), LED-2 (red dotted line) and the NTSC standard (black dotted line).
Nanomaterials 12 03142 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ren, Q.; Zhang, J.; Mao, Y.; Molokeev, M.S.; Zhou, G.; Zhang, X.-M. Ligand Engineering Triggered Efficiency Tunable Emission in Zero-Dimensional Manganese Hybrids for White Light-Emitting Diodes. Nanomaterials 2022, 12, 3142. https://doi.org/10.3390/nano12183142

AMA Style

Ren Q, Zhang J, Mao Y, Molokeev MS, Zhou G, Zhang X-M. Ligand Engineering Triggered Efficiency Tunable Emission in Zero-Dimensional Manganese Hybrids for White Light-Emitting Diodes. Nanomaterials. 2022; 12(18):3142. https://doi.org/10.3390/nano12183142

Chicago/Turabian Style

Ren, Qiqiong, Jian Zhang, Yilin Mao, Maxim S. Molokeev, Guojun Zhou, and Xian-Ming Zhang. 2022. "Ligand Engineering Triggered Efficiency Tunable Emission in Zero-Dimensional Manganese Hybrids for White Light-Emitting Diodes" Nanomaterials 12, no. 18: 3142. https://doi.org/10.3390/nano12183142

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop