Next Article in Journal
Modeling of the Flow Field and Clad Geometry of a Molten Pool during Laser Cladding of CoCrCuFeNi High-Entropy Alloys
Previous Article in Journal
Preparation and Characterization of Composites Based on ABS Modified with Polysiloxane Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mn(II)-Activated Zero-Dimensional Zinc(II)-Based Metal Halide Hybrids with Near-Unity Photoluminescence Quantum Yield

1
Traffic Information Engineering Institute, Guangxi Transport Vocational and Technical College, Nanning 530004, China
2
School of Semiconductors and Physics, North University of China, Taiyuan 030051, China
3
School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(3), 562; https://doi.org/10.3390/ma17030562
Submission received: 19 December 2023 / Revised: 18 January 2024 / Accepted: 19 January 2024 / Published: 25 January 2024

Abstract

:
As derivatives of metal halide perovskite materials, low-dimensional metal halide materials have become important materials that have attracted much attention in recent years. As one branch, zinc-based metal halides have the potential for practical applications due to their lead-free, low-toxicity and high-stability characteristics. However, pure zinc-based metal halide materials are still limited by their poor optical properties and cannot achieve large-scale practical applications. Therefore, in this work, we report an organic–inorganic hybrid zero-dimensional zinc bromide, (TDMP)ZnBr4, using transition metal Mn2+ ions as dopants and incorporating them into the (TDMP)ZnBr4 lattice. The original non-emissive (TDMP)ZnBr4 exhibits bright green emission under the excitation of external UV light after the introduction of Mn2+ ions with a PL peak position located at 538 nm and a PLQY of up to 91.2%. Through the characterization of relevant photophysical properties and the results of theoretical calculations, we confirm that this green emission in Mn2+:(TDMP)ZnBr4 originates from the 4T16A1 optical transition process of Mn2+ ions in the lattice structure, and the near-unity PLQY benefits from highly localized electrons generated by the unique zero-dimensional structure of the host material (TDMP)ZnBr4. This work provides theoretical guidance and reference for expanding the family of zinc-based metal halide materials and improving and controlling their optical properties through ion doping.

1. Introduction

Lead-based metal halide perovskites with the general structural formula ABX3 have received widespread attention from the scientific community in recent years due to their excellent optoelectronic properties, and they have been used in solar cells, light-emitting diodes, photodetectors and other fields [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. Although remarkable achievements have been made in the field of optoelectronic applications, the inherent toxicity and instability of lead-based materials limit their further practical applications [15,16]. How to improve the stability of materials and reduce the environmental toxicity of Pb2+ ions while maintaining excellent optoelectronic properties has become an urgent problem to be solved.
Low-dimensional metal halide materials developed based on ABX3 metal halide perovskites are ideal materials to replace lead-based metal halide materials [17,18]. In recent years, low-dimensional metal halide luminescent materials have been widely reported; in particular, Sb(III)-based, In(III)-based and Mn(II)-based materials have become typical representatives due to their high-efficiency luminescent properties and low toxicity [19,20,21,22,23,24]. In addition, low-dimensional Zn(II)-based materials based on Zn with a d10 electron configuration have gradually attracted the attention of researchers in the field in recent years due to their variable crystal structure and low-toxicity properties [25,26]. McWhorter et al. reported a Zn-based zero-dimensional metal halide, (C7H11N2)2ZnBr4, with white light emission with a PLQY of up to 7.35% [25]. At the same time, they also used single crystals of the compound to prepare X-ray detectors and explored future practical applications of Zn-based low-dimensional halide materials. Furthermore, they reported another Zn-based organic–inorganic hybrid metal halide, (P-xd) ZnCl4, with white light emission. This compound reached the current maximum white light emission PLQY in Zn-based materials of 23.06% [27]. In addition to broadband white light emission, Zn-based low-dimensional halide materials can also produce other colors of light. For example, R2ZnCl4 (R = (E)-4-styrylpyridinium, C13H12N+) with bright green emission, [N-EPD]2ZnBr4 with broadband yellow emission and (Ph3S)2ZnCl4 with blue emission have been reported [28,29,30]. Unfortunately, the optical properties of these materials are poor, which is not conducive to future practical applications. In addition to single-mode light emission, some Zn-based materials also have the characteristic of tunable light emission. Ma et al. reported a kind of (BPP)ZnX2 (X = Cl, Br) with a tunable emission color and ultralong room-temperature phosphorescence. Combining the results of experiments and theoretical calculations, they found that the dynamic polychromatic afterglow at room temperature can be attributed to the emission of the ligand BPP and its interaction with halogen atoms [31]. In addition, Wei et al. also reported a BAPPZn2 × 8 material with ultralong room-temperature phosphorescence, whose luminescence color is adjustable with the excitation wavelength and applied to the field of optical anti-counterfeiting [32]. Li et al. used a large-sized cyclic organic molecule, 2,5-bis(4-pyridinium)thiazolo [5,4-d]thiazole, as the A-site organic ligand and combined it with zinc halide to prepare a zero-dimensional structure of (H2TTz)ZnX4·MeOH (X = Cl, Br); the emission wavelength could switch from 462 nm to 512 nm, and they used Br to replace Cl in the lattice structure. Further, they explored the application of these two materials in the field of optical waveguides and found that both compounds had excellent optical waveguide properties. The optical loss coefficient R of (H2TTz)ZnCl4·MeOH at 462 nm was 0.01577 dB·μm−1, and that of (H2TTz)·ZnBr4·MeOH at 515 nm was 0.01182 dB·μm−1 [33]. These results revealed the great potential of low-dimensional Zn-based metal halide materials in photonic technology and provided new inspiration for the practical application of low-loss optical waveguides.
In summary, not only have many Zn-based halide materials been designed, developed and reported, but also a preliminary exploration of future practical applications has been carried out. However, the optical properties of the pure Zn-based metal halides reported so far still have inherent flaws and need to be further improved. How to improve the optical properties of Zn-based materials is of great significance to promoting their future commercial applications. Ion doping is an effective strategy to improve the optical properties of materials [34,35]. By introducing specific optically active ions into a non-luminescent matrix material, efficient luminescence can be achieved without changing the crystal structure of the matrix material. Taking the most widely reported and most intensively studied all-inorganic Cs-Zn-X (Cs2ZnX4 and Cs3ZnX5, X = Cl, Br, I) system as an example, many optically active ions have been doped into Cs-Zn-X and achieved a variety of efficient luminescent colors. The incorporation of Cu+ ions into Cs-Zn-X can achieve violet-to-blue light emission [36,37,38]. Cu+ ions usually exhibit blue emission after doping in chloride and bromide, while in iodide, they exhibit violet emission. Divalent Mn2+ ions usually form tetrahedrons in Zn-based halide systems and exhibit green light emission, and Sn2+ or Sb3+ exhibit deep red to near-infrared emission [39,40,41,42]. In addition to single doping, Cs-Zn-X co-doped with different luminescent ions has also received widespread attention. Wang et al. reported the tunable emission of Mn2+/Cu+ co-doped Cs2ZnBr4 and explored the application in the field of optical anti-counterfeiting [43]. Guo et al. reported Sb3+/Mn2+ co-doped Cs2ZnCl4 with dual-band emission. This material showed different luminescence responses at different temperatures. They built tunable thermochromic luminescent materials that can be used in applications such as thermal sensing and optical encryption [44]. Zeng et al. also reported Pb2+/Mn2+ co-doped Cs2ZnBr4 microcrystals with single-matrix white light emission. White light comes from the cooperative emission of Pb2+ ions and Mn3+ ions. At the same time, this white light emission can also be tuned from cold white to warm white by controlling the doping concentration of Pb2+ ions [45].
Inspired by ion doping, we used the protonated large-size cyclic organic molecule trans-2,5-dimethylpiperazine (C6H14N2, TDMP) as the A-site cation and combined it with zinc bromide to prepare a zero-dimensional organic–inorganic hybrid zinc bromide, (TDMP)ZnBr4. Optically active Mn2+ ions were introduced into the crystal lattice of non-luminescent (TDMP)ZnBr4 as dopants, and bright green light emission (λem = 538 nm) with a PLQY of up to 91.2% under UV light excitation was achieved. Through optical property characterization and first-principles calculations, it was confirmed that the bright green emission in Mn2+:(TDMP)ZnBr4 comes from the d-d transition of Mn2+ ions, and near-unity PLQY is related to the zero-dimensional structure of (TDMP)ZnBr4. The zero-dimensional structure is conducive to the generation of a strong quantum confinement effect and dielectric confinement effect, which can have a strong binding effect on photogenerated excitons and ensure an efficient radiation recombination process, thereby obtaining an ultra-high PLQY. This work provides a certain reference for improving the optical properties of Zn-based metal halide materials and lays a theoretical foundation for their future commercial applications.

2. Results and Discussion

2.1. Structure and Morphology

The crystal structure of the host material (TDMP)ZnBr4 is shown on the left side of Figure 1a. The (TDMP)ZnBr4 crystalline in the C2/c space group is as previously reported [46]. A (TDMP)ZnBr4 unit cell consists of four [ZnBr4]2− tetrahedrons and fifteen protonated TDMP molecules. [ZnBr4]2− tetrahedrons are periodically arranged in the structural skeleton composed of protonated TDMP molecules. Adjacent [ZnBr4]2− tetrahedrons are fully separated by protonated TDMP molecules and exhibit isolated distribution at the molecular level. Therefore, there are no interactions between adjacent [ZnBr4]2− tetrahedrons, which is typical of zero-dimensional structures. There are two types of Zn-Br bonds in each [ZnBr4]2− tetrahedron, with bond lengths of 2.3964 Å and 2.4248 Å. These two bond lengths are very similar, indicating that the structural distortion of the [ZnBr4]2− tetrahedron is very small. It is worth noting that, when Mn2+ ions are incorporated into the (TDMP)ZnBr4 lattice, they replace the lattice positions of Zn2+ ions to form [MnBr4]2− tetrahedrons, which indicates that the doping of Mn2+ ions does not change the crystal structure of (TDMP)ZnBr4. As shown on the right side of Figure 1a, a unit cell of Mn2+:(TDMP)ZnBr4 contains the metal–halogen tetrahedron, with Zn and Mn occupying the site of the tetrahedron center and protonated TDMP molecule. In Mn2+:(TDMP)ZnBr4, the metal–Br bond length of the tetrahedron is the same as the undoped system. There are two types of metal–Br bonds, with bond lengths of 2.5375 Å and 2.5675 Å. The deformation of tetrahedrons is also very small in Mn2+:(TDMP)ZnBr4, which indicates that the doping of Mn2+ ions does not cause local lattice distortion of the host material. In addition, the bond length of the metal–Br bond becomes larger after Mn2+ doping, indicating that the doping of Mn2+ ions causes the expansion of the tetrahedron. This is due to the different ionic radii of Mn2+ and Zn2+. In the case of four coordination, the ionic radius of Mn2+ is 0.66 Å, and the ionic radius of Zn2+ is 0.60 Å [47]. The incorporation of larger Mn2+ ions into the crystal structure will inevitably lead to the expansion of the lattice, which will increase the bond length of the metal–Br in the tetrahedron. The XRD patterns of (TDMP)ZnBr4 samples with different Mn2+ ion feed ratios are shown in Figure 1b. The undoped (TDMP)ZnBr4 lattice structure files used for simulation calculations were obtained from the Cambridge Crystallographic Data Center (CCDC-1424478). The XRD patterns of all (TDMP)ZnBr4 samples are in good agreement with those calculated using simulations, and no redundant diffraction peaks appear, indicating that Mn2+:(TDMP)ZnBr4 were successfully synthesized and that the samples were pure and free of impurities. At the same time, the diffraction peak of the sample moves to a low angle direction as the Mn2+ ion feed ratio increases. This result indicates the successful doping of Mn2+ ions, because the incorporation of Mn2+ ions with a larger radius will cause the lattice to expand and shift the diffraction peak to a lower angle direction. The surface morphology of Mn2+:(TDMP)ZnBr4 microcrystals and their distribution of elements were observed using a scanning electron microscope (SEM) and energy-dispersive X-ray spectrometer attachment (EDS), which obtain by HITACHI SU8020 (HITACHI, Tokyo, Japen). Figure 1c shows the surface morphology of the Mn2+:(TDMP)ZnBr4 microcrystals and their corresponding element distribution. The surface of rod-shaped Mn2+:(TDMP)ZnBr4 microcrystals is smooth, indicating that the microcrystalline samples synthesized through solvothermal synthesis have good crystallinity and high crystal quality. The results of the elemental analysis show that Zn, Mn and Br elements are evenly distributed in the crystal, and the atomic ratio of the three elements is Zn:Mn:Br = 20.4:0.95:78.65 = 1:0.047:3.855 (Figure S1). The atomic ratio of Zn to Br is close to the stoichiometric ratio of 1:4 in (TDMP)ZnBr4, which further demonstrates the successful synthesis of (TDMP)ZnBr4 and the doping of Mn2+ ions. In addition, the Raman spectrum of undoped (TDMP)ZnBr4 and Mn2+:(TDMP)ZnBr4 obtained with a 633 nm continuous laser as the excitation source was also collected, as shown in Figure S2. Figure S2a shows the inorganic part of the Raman spectra. The Raman mode in the low wavenumber (<400 cm−1) range originates from the vibrations of inorganic [ZnBr4]2− and [MnBr4]2− tetrahedrons [48]. The Raman mode of the Mn2+ ion-doped sample shifts to a low wavenumber direction compared with the undoped sample, which is consistent with the XRD pattern. This result also proves that Mn2+ ions were successfully doped in the structure of (TDMP)ZnBr4, because the expansion or contraction of the crystal lattice causes a shift in the Raman peak [49]. The Raman spectrum of the organic part demonstrates that the protonated organic TDMP molecules were successfully inserted into the lattice of inorganic ZnBr2 (Figure S2b).

2.2. Optical Properties

The optical properties of Mn2+:(TDMP)ZnBr4 were characterized to explore their underlying photophysical mechanisms. First, a Lambda-750 UV-vis-NIR spectrophotometer was used to collect the absorption spectra of (TDMP)ZnBr4 with different Mn2+ ion feed ratios. As shown in Figure 2a, undoped (TDMP)ZnBr4 absorbs in the range of 250 nm to 550 nm, and the absorption intensity gradually increases as the wavelength decreases. Three new absorption bands appear at 267 nm, 285 nm and 353 nm when Mn ions are incorporated, and the intensity of these three absorption bands increases with the increase in the Mn2+ ion feed ratio. Obviously, these absorption bands come from the d-d transition absorption of Mn2+ ions. According to the elemental analysis results, the actual Mn2+ ions incorporated into the (TDMP)ZnBr4 lattice are still quite limited, even though the feed ratio of Mn2+ ions reaches 40% or even 60%. The following formula was used to calculate the bandgap value of (TDMP)ZnBr4 with different Mn2+ ion feed ratios [50]:
[F(R)]n = A(Eg)
where F(R) is the Kubelka–Munk (K-M) equation, is the photon energy, A is the constant, and Eg is the bandgap. The value of n depends on the bandgap properties of the semiconductor. n = 2 for the direct bandgap semiconductor, and n = 1/2 for the indirect bandgap semiconductor. Since (TDMP)ZnBr4 is a direct bandgap semiconductor, the bandgap values of the (TDMP)ZnBr4 samples with different Mn2+ ion feed ratios were calculated according to the above formula, and the results are shown in Figure S3. The bandgap of the undoped (TDMP)ZnBr4 sample is 3.71 eV, and the bandgap will slightly reduce while Mn2+ ions are introduced.
The photoluminescence excitation spectrum (PLE) and photoluminescence (PL) spectrum of Mn2+:(TDMP)ZnBr4 are shown in Figure 2b. Mn2+:(TDMP)ZnBr4 exhibits bright green emission under the excitation of 365 nm UV light, with the PL peak located at 538 nm; the full width of half maximum (FWHM) is 51.71 nm, and the Stokes shift is 173 nm. It is worth noting that the green emission of Mn2+:(TDMP)ZnBr4 is not completely standard green. The CIE coordinate of this green emission is (0.30, 0.67) and located at the edge of the green light area, as shown in Figure S4. The PLE spectrum of Mn2+:(TDMP)ZnBr4 was collected at the PL position of 538 nm. Obvious excitation bands can be observed at 275 nm, 291 nm, 360 nm, 375 nm, 388 nm and 462 nm, which belong to the absorption transitions of 6A1 to 4E(D), 4T2(D), 4A1(G), 4E(G), 4T2(G) and 4T1(D) of Mn2+ ions, respectively [51,52]. The excitation band in the PLE spectrum does not observe the contribution of the host material (TDMP)ZnBr4 to the 538 nm green emission. All the excitation bands come from the absorption transition of Mn2+ ions, which shows that the optical transition process of Mn2+:(TDMP)ZnBr4 is dominated by the Mn2+ ions doped in the crystal lattice. The PL spectra of (TDMP)ZnBr4 with different Mn2+ ion feed ratios are shown in Figure S5. As the feed ratio increases, the PL intensity of (TDMP)ZnBr4 first increases and reaches the maximum when the Mn2+ feed ratio is up to 20%. Then, the PL intensity of (TDMP)ZnBr4 decreases as the feed ratio of Mn2+ ions is further increased, which is caused by the concentration quenching effect under high-concentration doping. The PLQY of the Mn2+-doped sample was collected using 365 nm UV light as the excitation source. The measurement results show that Mn2+:(TDMP)ZnBr4 reaches the maximum of 91.2% when the Mn2+ ion feed ratio is 20%, as shown in Figure S6. In order to further explore the origin of green emission in Mn2+:(TDMP)ZnBr4, we collected the PL spectra under different excitation wavelengths and the PLE spectra under different emission wavelengths, as shown in Figure S7. The PLE spectra of Mn2+:(TDMP)ZnBr4 collected at different emission wavelengths were normalized. We found that the obtained PLE spectra all have the same profile as the emission wavelength increases from 520 nm to 550 nm, as shown in Figure S7a. At the same time, the PL spectra obtained at different excitation wavelengths also show the same profile (Figure S7b). This indicates that the 538 nm green emission in Mn2+:(TDMP)ZnBr4 originates from the same excited state.
The PL lifetime decay curves of (TDMP)ZnBr4 under different Mn2+ ion feed ratios were collected to confirm the source of green emission. As shown in Figure 2c, all Mn2+:(TDMP)ZnBr4 samples exhibit hundreds of microseconds PL decay lifetimes. This long lifetime of hundreds of microseconds to tens of milliseconds is a typical feature of the Mn2+ ions’ d-d transition (4T16A1) [51,53,54]. Combining the PLE spectrum and PL spectrum results, we confirmed that the 538 nm green emission in Mn2+:(TDMP)ZnBr4 originates from the 4T16A1 radiation transition of Mn2+ ions. Therefore, the photophysical processes in Mn2+:(TDMP)ZnBr4 can be summarized as follows: Optically active Mn2+ ions are introduced into non-luminescent (TDMP)ZnBr4. Mn2+ ions replace the lattice site of Zn2+ ions and combine with the surrounding four Br ions to form [MnBr4]2− tetrahedrons when they enter the (TDMP)ZnBr4 lattice. According to the crystal field theory, the crystal field intensity experienced by the four-coordinated Mn2+ ions is weak. The energy level difference between the 4T1 and 6A1 energy levels is large, leading to high-energy green emission under the excitation of UV light. According to the results of the PLE spectrum, the host material does not contribute to the emission; hence, the excitation–emission process in Mn2+:(TDMP)ZnBr4 occurs entirely at the d energy level of the Mn2+ ions. As shown in Figure 2d, under UV light excitation, electrons at the ground-state energy level 6A1 absorb the energy of UV light and transition to excited states of higher energy levels. Subsequently, the electrons in the high-energy excited state will return to the lowest energy level 4T1 excited state through non-radiative transition and, finally, transition back to the ground state through radiative recombination producing 538 nm green emission.
In order to deeply study the relationship between the crystal structure and photophysical properties, PL spectra of Mn2+:(TDMP)ZnBr4 in the temperature range of 80–400 K were collected. As shown in Figure 3a, Mn2+:(TDMP)ZnBr4 exhibits single-band emission in the entire temperature range from 80 K to 400 K, without additional emission bands appearing with temperature changes. This indicates that Mn2+:(TDMP)ZnBr4 has a high structural stability and that the crystal structure will not undergo phase changes with drastic changes in temperature, because phase changes may generate new emission bands. The PL intensity at each temperature was extracted from the temperature-dependent PL spectrum and used as the vertical axis with the reciprocal of temperature as the horizontal axis, and a scatter plot of the changing trend of PL intensity with the reciprocal of temperature was drawn. The scatter plot was fitted using the Arrhenius formula to obtain the activation energy Eb of Mn2+:(TDMP)ZnBr4 [52]:
  I ( T ) = I 0 1 + A e E b k b T
where I0 is the PL intensity at 0 K, I(T) is the PL intensity at temperature K, and kb is the Boltzmann constant. The activation energy Eb obtained by fitting is 43.75 meV, which is much higher than the room-temperature thermal energy (~26 meV) [55]. This indicates that the emission of Mn2+:(TDMP)ZnBr4 has strong resistance to thermal quenching. A larger activation energy means that the dissociation process requires more energy, which also indicates that the excitons are more stable and have stronger resistance to thermal quenching [56]. This means that, when the temperature reaches 400 K, Mn2+:(TDMP)ZnBr4 can still maintain a certain intensity of emission, as shown in the pseudo-color image of the temperature-dependent PL spectrum (Figure 3c). In addition, we also extracted the FWHM of the Mn2+:(TDMP)ZnBr4 spectrum at each temperature and fitted the FWHM–temperature change curve using the following formula [57]:
FWHM T = 2.36 S ω phonon coth ω phonon 2 k b T
where S is the Huang–Rhys factor, ωphonon is the phonon frequency, T is the temperature, FWHM is the full-width half maximum of the PL spectra, and kb is the Boltzmann constant. The fitting results are shown in Figure 3d, and Huang–Rhys factor S obtained by fitting is 6.65. Such a small S value indicates that the emission of Mn2+:(TDMP)ZnBr4 comes from the d-d transition of Mn2+ ions rather than self-trapped exciton (STE) emission caused by local lattice distortion because STE emission requires a certain intensity of the electron–phonon coupling effect, and the S value obtained by fitting is larger [58]. In addition, the measurement results of the Raman spectrum also confirm this result. Strong Raman scattering at room temperature is an important indicator of strong coupling between excited-state electrons and lattice vibrations [59]. Therefore, strong scattering interactions reflect the tendency of local fluctuations in the lattice. This propensity to form localized charges plays a role in charge transport, as free charge carriers cause lattice deformation, which moves with the carriers and forms self-trapped excitons. The Raman spectrum in Figure S2 shows a very weak Raman scattering intensity, which indicates that the coupling between its electrons and lattice vibration is too weak to produce STE emission. At the same time, cyclic organic molecules with greater structural rigidity are also an important reason for the smaller S value, because the lattice vibration produced by rigid organic molecules is smaller [60].

2.3. Theoretical Calculation

In order to study the relationship between the Mn2+:(TDMP)ZnBr4 electronic structure and luminescence properties, we performed first-principles calculations on (TDMP)ZnBr4 before and after Mn2+ ion doping. Figure 4a,c present the energy band structure of (TDMP)ZnBr4 before and after Mn2+ ion doping, respectively. The pristine (TDMP)ZnBr4 bandgap value obtained by first-principles calculations is 3.76 eV, while the bandgap value of Mn2+:(TDMP)ZnBr4 is 3.12 eV. The calculation results show that the doping of Mn2+ ions shrinks the bandgap of the host material, which is consistent with the actual bandgap change in the samples obtained using the tauc plot method. Figure 4b,d show the density of state (DOS) of (TDMP)ZnBr4 and Mn2+:(TDMP)ZnBr4, respectively. In undoped (TDMP)ZnBr4, the conduction band minimum (CBM) is formed by Br-4p orbitals and Zn-5s orbitals, while the valence band maximum (VBM) is mainly formed by Br-5p orbitals, as shown in Figure 4b. The VBM of Mn2+:(TDMP)ZnBr4 is formed by Br-4p orbitals and Mn-3d orbitals, while the CBM is mainly formed by Br-4p orbitals, as shown in Figure 4d. The theoretical calculation results indicate that the Zn-5s orbital involved in band-edge formation is replaced by the Mn-3d and Br-4p orbitals when Mn2+ ions are incorporated. This indicates that [MnBr4]2− tetrahedrons dominate the main optical transition process in Mn2+:(TDMP)ZnBr4 after doping. The UV light directly excites Mn2+ ions in the lattice structure, and all optical processes occur between the d orbital energy levels of Mn2+ ions. This also explains why the contribution of the host material to 538 nm green emission cannot be observed in the PLE spectrum. Organic cations do not contribute to the formation of frontier orbits but could affect the optical processes by producing spatial confinement effects and dielectric confinement effects on [MnBr4]2− tetrahedrons. At the same time, Mn2+:(TDMP)ZnBr4 shows a relatively flat conduction band and valence band with very small dispersion, which indicates that the electrons are highly localized (Figure 4c). This band structure is unique to the 0D structure [22]. Each [MnBr4]2− tetrahedron is separated by organic cations to form the highly localized electrons in Mn2+:(TDMP)ZnBr4, thus generating the isolated luminescent centers. This highly localized electron is a key factor in producing efficient luminescence [22,61].

2.4. Stability

The stability of a material is an important indicator of whether it has practical application value. Here, we determined the structure, luminescence and thermal stability of Mn2+:(TDMP)ZnBr4 to evaluate its potential for practical applications in the future. As shown in Figure 5a, the XRD diffraction pattern of Mn2+:(TDMP)ZnBr4 after being placed under ambient conditions for 60 days is basically consistent with the diffraction pattern of the as-synthesized sample, and no additional impurity diffraction peaks were observed, indicating that this compound can retain a stable structure without decomposition for a long time. Meanwhile, Mn2+:(TDMP)ZnBr4 also has high PL stability. The PL intensity of Mn2+:(TDMP)ZnBr4 does not decrease significantly and remains at 90–95% of the initial PL intensity after being placed under ambient conditions for 60 days (Figure 5b), and it can exhibit bright green emission under UV light excitation. Finally, the results of the thermogravimetric analysis show that Mn2+:(TDMP)ZnBr4 also has high thermal stability. As shown in Figure 5c, the decomposition temperature of Mn2+:(TDMP)ZnBr4 is about 342 °C. All experiments, including synthesis, characterization and storage, were performed under ambient conditions, with a relative humidity of 50–70% without any inert gas protection. The stability test results show that Mn2+:(TDMP)ZnBr4 has great structural, luminescence and thermal stability, and it has application potential in the field of future optoelectronic devices.

3. Conclusions

In summary, we selected transition metal Mn2+ ions as sensitizers and introduced them into a non-luminescent (TDMP)ZnBr4 lattice in this work. Under the excitation of 365 nm UV light, Mn2+:(TDMP)ZnBr4 exhibits bright green light emission (λem = 538 nm) with a PLQY of up to 91.2%. Combining optical property characterization and first-principles calculations, we confirmed that the 538 nm green light emission in Mn2+:(TDMP)ZnBr4 originates from the 4T16A1 transition of Mn2+ ions in the lattice structure. The unique 0D structure of the host material can generate a strong quantum confinement effect and dielectric confinement effect, which is very beneficial to the generation of highly localized excitons. The photogenerated excitons will be bound within the tetrahedron by the strong confinement effect of the crystal lattice once they are formed by outside excitation, and they will radiate photons outward through radiative recombination, ultimately producing efficient green light emission. Otherwise, Mn2+:(TDMP)ZnBr4 also has high structural, luminescence and thermal stability, which further enhances the competitiveness of this compound in future practical applications. This work provides theoretical guidance and references for expanding the zinc-based metal halide family and improving its optical properties via ion doping.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma17030562/s1, Figure S1: EDS spectra, Figure S2: Raman spectra, Figure S3: Bandgap value, Figure S4: CIE coordination, Figure S5: PL spectra, Figure S6: PLQY, Figure S7: PLE and PL spectra.

Author Contributions

Methodology, C.P.; Software, Q.W.; Validation, L.D.; Resources, J.W.; Writing—original draft, C.P.; Writing—review & editing, C.P. and Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study received funding from the National Natural Science Foundation of China (grant no. 62175219), Fundamental Research Program of Shanxi Province (grant no. 20210302124397) and the open research fund of the State Key Laboratory of Dynamic Testing Technology (grant no. 2022-SYSJJ-01).

Institutional Review Board Statement

Not involving humans or animals.

Informed Consent Statement

Studies not involving humans.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Ye, J.; Byranvand, M.M.; Martínez, C.O.; Hoye, R.L.; Saliba, M.; Polavarapu, L. Defect Passivation in Lead-Halide Perovskite Nanocrystals and Thin Films: Toward Efficient LEDs and Solar Cells. Angew. Chem. Int. Ed. 2021, 133, 21804–21828. [Google Scholar] [CrossRef]
  2. Safronov, K.R.; Popkova, A.A.; Markina, D.I.; Pushkarev, A.P.; Makarov, S.V.; Bessonov, V.O.; Fedyanin, A.A. Efficient Emission Outcoupling from Perovskite Lasers into Highly Directional and Long-Propagation-Length Bloch Surface Waves. Laser Photonics Rev. 2022, 16, 2100728. [Google Scholar] [CrossRef]
  3. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.-K.; Moon, C.S.; Jeon, N.J.; Correa-Baena, J.-P. Efficient Perovskite Solar Cells via Improved Carrier Management. Nature 2021, 590, 587–593. [Google Scholar] [CrossRef] [PubMed]
  4. Jokar, E.; Cai, L.; Han, J.; Nacpil, E.J.C.; Jeon, I. Emerging Opportunities in Lead-Free and Lead–Tin Perovskites for Environmentally Viable Photodetector Applications. Chem. Mater. 2023, 35, 3404–3426. [Google Scholar] [CrossRef]
  5. Moseley, O.D.; Roose, B.; Zelewski, S.J.; Stranks, S.D. Identification and Mitigation of Transient Phenomena That Complicate the Characterization of Halide Perovskite Photodetectors. ACS Appl. Energy Mater. 2023, 6, 10233–10242. [Google Scholar] [CrossRef] [PubMed]
  6. Isikgor, F.H.; Zhumagali, S.; Merino, L.V.T.; De Bastiani, M.; McCulloch, I.; De Wolf, S. Molecular Engineering of Contact Interfaces for High-Performance Perovskite Solar Cells. Nat. Rev. Mater. 2023, 8, 89–108. [Google Scholar] [CrossRef]
  7. Furasova, A.; Voroshilov, P.; Sapori, D.; Ladutenko, K.; Barettin, D.; Zakhidov, A.; Di Carlo, A.; Simovski, C.; Makarov, S. Nanophotonics for Perovskite Solar Cells. Adv. Photonics Res. 2022, 3, 2100326. [Google Scholar] [CrossRef]
  8. Fakharuddin, A.; Gangishetty, M.K.; Abdi-Jalebi, M.; Chin, S.-H.; bin Mohd Yusoff, A.R.; Congreve, D.N.; Tress, W.; Deschler, F.; Vasilopoulou, M.; Bolink, H.J. Perovskite Light-Emitting Diodes. Nat. Electron. 2022, 5, 203–216. [Google Scholar] [CrossRef]
  9. Sánchez, R.S.; Villanueva-Antolí, A.; Bou, A.; Ruiz-Murillo, M.; Mora-Seró, I.; Bisquert, J. Radiative Recombination Processes in Halide Perovskites Observed by Light Emission Voltage Modulated Spectroscopy. Adv. Mater. 2023, 35, 2207993. [Google Scholar] [CrossRef]
  10. Liao, C.-H.; Mahmud, M.A.; Ho-Baillie, A.W. Recent Progress in Layered Metal Halide Perovskites for Solar Cells, Photodetectors, and Field-Effect Transistors. Nanoscale 2023, 15, 4219–4235. [Google Scholar] [CrossRef]
  11. Dyrvik, E.G.; Warby, J.H.; McCarthy, M.M.; Ramadan, A.J.; Zaininger, K.-A.; Lauritzen, A.E.; Mahesh, S.; Taylor, R.A.; Snaith, H.J. Reducing Nonradiative Losses in Perovskite LEDs Through Atomic Layer Deposition of Al2O3 on The Hole-Injection Contact. ACS Nano 2023, 17, 3289–3300. [Google Scholar] [CrossRef] [PubMed]
  12. Moeini, A.; Martínez-Sarti, L.; Zanoni, K.P.; Sessolo, M.; Tordera, D.; Bolink, H.J. Semitransparent Near-Infrared Sn–Pb Hybrid Perovskite Photodetectors. J. Mater. Chem. C 2022, 10, 13878–13885. [Google Scholar] [CrossRef]
  13. Rahmany, S.; Etgar, L. Semitransparent Perovskite Solar Cells. ACS Energy Lett. 2020, 5, 1519–1531. [Google Scholar] [CrossRef]
  14. Bisquert, J.; Gonzales, C.; Guerrero, A. Transient On/Off Photocurrent Response of Halide Perovskite Photodetectors. J. Phys. Chem. C 2023, 127, 21338–21350. [Google Scholar] [CrossRef]
  15. Hutter, E.M.; Sangster, R.; Testerink, C.; Ehrler, B.; Gommers, C.M. Metal Halide Perovskite Toxicity Effects on Arabidopsis Thaliana Plants are Caused by Iodide Ions. Iscience 2022, 25, 103583. [Google Scholar] [CrossRef]
  16. Ikram, M.; Malik, R.; Raees, R.; Imran, M.; Wang, F.; Ali, S.; Khan, M.; Khan, Q.; Maqbool, M. Recent advancements and future insight of lead-free non-toxic perovskite solar cells for sustainable and clean energy production: A review. Sustain. Energy Technol. Assess. 2022, 53, 102433. [Google Scholar] [CrossRef]
  17. Ray, A.; De Trizio, L.; Zito, J.; Infante, I.; Manna, L.; Abdelhady, A.L. Light Emission from Low-Dimensional Pb-Free Perovskite-Related Metal Halide Nanocrystals. Adv. Opt. Mater. 2023, 11, 2202005. [Google Scholar] [CrossRef]
  18. Morana, M.; Kaiser, W.; Chiara, R.; Albini, B.; Meggiolaro, D.; Mosconi, E.; Galinetto, P.; De Angelis, F.; Malavasi, L. Origin of Broad Emission Induced by Rigid Aromatic Ditopic Cations in Low-Dimensional Metal Halide Perovskites. J. Phys. Chem. Lett. 2023, 14, 7860–7868. [Google Scholar] [CrossRef]
  19. Schäfer, T.C.; Becker, J.; Heuler, D.; Seuffert, M.T.; Sedykh, A.E.; Müller-Buschbaum, K. Coordination Polymers Based on Aluminum and Indium Halides Together with Pyrazine. Aust. J. Chem. 2022, 75, 676–683. [Google Scholar] [CrossRef]
  20. Pinchetti, V.; Moro, F.; Zhang, B.; Fanciulli, M.; De Trizio, L.; Meinardi, F.; Manna, L.; Brovelli, S. Magnetic Transitions and Energy Transfer Processes in Sb-Based Zero-Dimensional Metal Halide Nanocrystals Doped with Manganese. ACS Energy Lett. 2022, 7, 1566–1573. [Google Scholar] [CrossRef]
  21. Abdelhadi, A.B.; Gutiérrez, M.; Cohen, B.; Lezama, L.; Lachkar, M.; Douhal, A. A New Eco-Friendly and Highly Emitting Mn-Based Hybrid Perovskite toward High-Performance Green Down-Converted LEDs. J. Mater. Chem. C 2024, 12, 286–295. [Google Scholar] [CrossRef]
  22. Kshirsagar, A.S.; Singh, Y.; Kuruppu, U.M.; Donnadieu, B.; Rai, N.; Gangishetty, M.K. Pivotal Role of A-Site Cations in Tailoring the Band-Edge States, Optical Properties, and Stability of 0D Hybrid Indium Chlorides. Chem. Mater. 2022, 34, 10928–10939. [Google Scholar] [CrossRef]
  23. Dutta, S.; Yoo, J.H.; Kwon, S.B.; Yoo, H.; Yoon, D.H. Simple Water-Assisted Synthesis of Ce-Doped Lead-Free Cesium Metal Halides for Visible-Blind UV Photodetectors. ACS Sustain. Chem. Eng. 2022, 11, 92–100. [Google Scholar] [CrossRef]
  24. Biswas, A.; Bakthavatsalam, R.; Das, D.K.; Sam, J.; Mali, B.P.; Biswas, C.; Maana, N.; Thomson, S.; Raavi, S.S.K.; Kurungot, S. Synergistic Electronic Coupling/Cross-Talk between the Isolated Metal Halide Units of Zero Dimensional Heterometallic (Sb, Mn) Halide Hybrid with Enhanced Emission. J. Mater. Chem. C 2022, 10, 360–370. [Google Scholar] [CrossRef]
  25. McWhorter, T.M.; Zhang, Z.; Creason, T.D.; Thomas, L.; Du, M.H.; Saparov, B. (C7H11N2)2MBr4 (M = Cu, Zn): X-ray Sensitive 0D Hybrid Metal Halides with Tunable Broadband Emission. Eur. J. Inorg. Chem. 2022, 2022, e202100954. [Google Scholar] [CrossRef]
  26. Marayathungal, J.H.; Das, D.K.; Bakthavatsalam, R.; Sam, J.; Hathwar, V.R.; Pallepogu, R.; Dutta, S.; Kundu, J. Mn2+-Activated Zero-Dimensional Metal (Cd, Zn) Halide Hybrids with Near-Unity PLQY and Zero Thermal Quenching. J. Phys. Chem. C 2023, 127, 8618–8630. [Google Scholar] [CrossRef]
  27. Popy, D.; Evans, B.; Jiang, J.; Creason, T.; Banerjee, D.; Loftus, L.; Pachter, R.; Glatzhofer, D.; Saparov, B. Intermolecular Arrangement Facilitated Broadband Blue Emission in Group-12 Metal (Zn, Cd) Hybrid Halides and Their Applications. Mater. Today Chem. 2023, 30, 101502. [Google Scholar] [CrossRef]
  28. Luo, Z.; Liu, Y.; Liu, Y.; Li, C.; Li, Y.; Li, Q.; Wei, Y.; Zhang, L.; Xu, B.; Chang, X. Integrated Afterglow and Self-Trapped Exciton Emissions in Hybrid Metal Halides for Anti-Counterfeiting Applications. Adv. Mater. 2022, 34, 2200607. [Google Scholar] [CrossRef]
  29. Creason, T.D.; Fattal, H.; Gilley, I.W.; Evans, B.N.; Jiang, J.; Pachter, R.; Glatzhofer, D.T.; Saparov, B. Stabilized Photoemission from Organic Molecules in Zero-Dimensional Hybrid Zn and Cd Halides. Inorg. Chem. Front. 2022, 9, 6202–6210. [Google Scholar] [CrossRef]
  30. Liu, Y.-H.; Wang, W.-Q.; Zhang, B.-L.; Wang, Y.-J.; Ren, M.-P.; Jing, Z.; Yue, C.-Y. Zero-Dimensional Organic–Inorganic Hybrid Znc Halide with Broadband Yellow Light Emission. Cryst. Eng. Comm. 2023, 25, 444–449. [Google Scholar] [CrossRef]
  31. Ma, Y.-J.; Qi, Z.; Xiao, G.; Fang, X.; Yan, D. Metal-Halide Coordination Polymers with Excitation Wavelength-and Time-Dependent Ultralong Room-Temperature Phosphorescence. Inorg. Chem. 2022, 61, 16477–16483. [Google Scholar] [CrossRef] [PubMed]
  32. Wei, J.H.; Ou, W.T.; Luo, J.B.; Kuang, D.B. Zero-Dimensional Zn-Based Halides with Ultra-Long Room-Temperature Phosphorescence for Time-Resolved Anti-Counterfeiting. Angew. Chem. Int. Ed. 2022, 134, e202207985. [Google Scholar] [CrossRef]
  33. Li, K.-J.; Zhao, Y.-Y.; Sun, M.-E.; Chen, G.-S.; Zhang, C.; Liu, H.-L.; Li, H.-Y.; Zang, S.-Q.; Mak, T.C. Zero-Dimensional Zinc Halide Organic Hybrids with Excellent Optical Waveguide Properties. Cryst. Growth. Des. 2022, 22, 3295–3302. [Google Scholar] [CrossRef]
  34. Marciniak, L.; Kniec, K.; Elżbieciak-Piecka, K.; Trejgis, K.; Stefanska, J.; Dramićanin, M. Luminescence Thermometry with Transition Metal Ions: A Review. Coordin. Chem. Rev. 2022, 469, 214671. [Google Scholar] [CrossRef]
  35. Adhikari, S.D.; Echeverría-Arrondo, C.; Sánchez, R.S.; Chirvony, V.S.; Martínez-Pastor, J.P.; Agouram, S.; Muñoz-Sanjosé, V.; Mora-Seró, I. White Light Emission from Lead-Free Mixed-Cation Doped Cs2SnCl6 Nanocrystals. Nanoscale 2022, 14, 1468–1479. [Google Scholar] [CrossRef]
  36. Babu, R.; López-Fernández, I.; Prasanthkumar, S.; Polavarapu, L. All-Inorganic Lead-Free Doped-Metal Halides for Bright Solid-State Emission from Primary Colors to White Light. ACS Appl. Mater. Interfaces 2023, 15, 35206–35215. [Google Scholar] [CrossRef]
  37. Yang, Y.; Wu, J.; Zhou, T.; Wang, Y.; Zheng, J.; Liu, R.; Hou, J.; Li, X.; Wang, L.; Jiang, W. Efficient Emission in Copper-Doped Cs3ZnX5 (X = Cl, I) for White LED and X-ray Scintillator. J. Mater. Chem. C 2023, 11, 9850–9859. [Google Scholar] [CrossRef]
  38. Geng, D.; Cai, B.; Ouyang, Z.; Cheng, P.; Li, B.; Zeng, H. Heterovalent Doping Enabled Efficient and Stable Extrinsic Self-Trapped Exciton Emission in Zero-Dimensional Cesium Zinc Halides. Adv. Opt. Mater. 2022, 10, 2201061. [Google Scholar] [CrossRef]
  39. Liu, X.; Zhang, W.; Xu, R.; Tu, J.; Fang, G.; Pan, Y. Bright Tunable Luminescence of Sb3+ Doping in Zero-Dimensional Lead-Free Halide Cs3ZnCl5 Perovskite Crystals. Dalton Trans. 2022, 51, 10029–10035. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Zhou, L.; Zhang, L.; Luo, W.; Shen, W.; Li, M.; He, R. Highly Stable Metal Halides Cs2ZnX4 (X= Cl, Br) with Sn2+ as Dopants for Eefficient Deep-Red Photoluminescence. Chinese. Chem. Lett. 2023, 34, 107556. [Google Scholar] [CrossRef]
  41. 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] [PubMed]
  42. Wang, X.; Shen, Q.; Chen, Y.; Ali, N.; Ren, Z.; Bi, G.; Wu, H. Self-Trapped Exciton Emission in an Sn(II)-Doped All-Inorganic Zero-Dimensional Zinc Halide Perovskite Variant. Nanoscale 2021, 13, 15285–15291. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, M.; Yang, R.; Cheng, S.; Li, G.; Jia, M.; Chen, X.; Wu, D.; Li, X.; Shi, Z. Mn and Cu Codoped Cs2ZnBr4 Metal Halide with Multiexcitonic Emission Toward Anti-Counterfeiting. J. Phys-Condens. Mat. 2022, 34, 204009. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, Y.; Chen, J.; Chen, B.; Zheng, W.; Zhang, X.; Suo, H.; Chun, F.; Wei, X.; Wang, F. Sequential Thermochromic Switching in Zero-Dimensional Cs2ZnCl4 Metal Halides. Mater. Today Phys. 2023, 35, 101111. [Google Scholar] [CrossRef]
  45. Zeng, X.; Xu, Y.; Sun, D.; Yu, Y.; Miao, C.; Fu, Y.; Yu, L. Pb2+/Mn2+ Co-Doped Cs2ZnBr4 Microcrystals with Dual-Band Tunable White Light Emission. Ceram. Int. 2023, 50, 3958–3966. [Google Scholar] [CrossRef]
  46. Gautier, R.; Paris, M.; Massuyeau, F. Templating Effect of Trans-2, 5-Dimethylpiperazine (TDMP) on the Structural Dimensionality of Hybrid Metal Halides. Dalton Trans. 2022, 51, 10758–10762. [Google Scholar] [CrossRef]
  47. Shannon, R.D. Revised Efective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  48. Barreda-Argüeso, J.A.; Nataf, L.; Rodríguez-Lazcano, Y.; Aguado, F.; González, J.s.; Valiente, R.; Rodríguez, F.; Wilhelm, H.; Jephcoat, A.P. Bulk and Molecular Compressibilities of Organic–Inorganic Hybrids [(CH3)4N]2MnX4 (X= Cl, Br): Role of Intermolecular Interactions. Inorg. Chem. 2014, 53, 10708–10715. [Google Scholar] [CrossRef]
  49. Late, D.J.; Shirodkar, S.N.; Waghmare, U.V.; Dravid, V.P.; Rao, C.N. Thermal Expansion, Anharmonicity and Temperature-Dependent Raman Spectra of Single- and Few-Layer MoSe2 and WSe2. ChemPhysChem 2014, 15, 1592–1598. [Google Scholar] [CrossRef]
  50. Zhong, H.; Pan, F.; Yue, S.; Qin, C.; Hadjiev, V.; Tian, F.; Liu, X.; Lin, F.; Wang, Z.; Bao, J. Idealizing Tauc Plot for Accurate Bandgap Determination of Semiconductor with Ultraviolet–Visible Spectroscopy: A Case Study for Cubic Boron Arsenide. J. Phys. Chem. Lett. 2023, 14, 6702–6708. [Google Scholar] [CrossRef]
  51. Yadav, R.S.; Pandey, S.K.; Pandey, A.C. BaAl12O19:Mn2+ Green Emitting Nanophosphor for PDP Application Synthesized by Solution Combustion Method and its Vacuum Ultra-Violet Photoluminescence Characteristics. J. Lumin. 2011, 131, 1998–2003. [Google Scholar] [CrossRef]
  52. Samanta, T.; Viswanath, N.S.M.; Jang, S.W.; Min, J.W.; Cho, H.B.; Han, J.H.; Im, W.B. Thermally Stable Self-Trapped Assisted Single-Component White Light from Lead-Free Zero-Dimensional Metal Halide Nanocrystals. Adv. Opt. Mater. 2023, 11, 2202744. [Google Scholar] [CrossRef]
  53. Bortoluzzi, M.; Castro, J.; Ferraro, V. Dual Emission from Mn (II) Complexes with Carbazolyl-Substituted Phosphoramides. Inorg. Chim. Acta. 2022, 536, 120896. [Google Scholar] [CrossRef]
  54. Sreenan, B.; Lee, B.; Wan, L.; Zeng, R.; Zhao, J.; Zhu, X. Review of Mn-Doped Semiconductor Nanocrystals for Time-Resolved Luminescence Biosensing/Imaging. ACS Appl. Nano Mater. 2022, 5, 17413–17435. [Google Scholar] [CrossRef] [PubMed]
  55. Wei, J.-H.; Liao, J.-F.; Wang, X.-D.; Zhou, L.; Jiang, Y.; Kuang, D.-B. All-Inorganic Lead-Free Heterometallic Cs4MnBi2Cl12 Perovskite Single Crystal with Highly Efficient Orange Emission. Matter 2020, 3, 892–903. [Google Scholar] [CrossRef]
  56. Li, C.; Luo, Z.; Liu, Y.; Wei, Y.; He, X.; Chen, Z.; Zhang, L.; Chen, Y.; Wang, W.; Liu, Y. Self-Trapped Exciton Emission with High Thermal Stability in Antimony-Doped Hybrid Manganese Chloride. Adv. Opt. Mater. 2022, 10, 2102746. [Google Scholar] [CrossRef]
  57. Das, D.K.; Bakthavatsalam, R.; Hathwar, V.R.; Pallepogu, R.; Kundu, J. Intrinsic vs. Extrinsic STE Emission Enhancement in ns2 Ion Doped Metal (Cd, In) Halide Hybrids. J. Mater. Chem. C 2023, 11, 3855–3864. [Google Scholar] [CrossRef]
  58. Dakshinamurthy, A.C.; Sudakar, C. Sublattice Distortion Enabled Strong Interplay between Phonon Vibrations, Electron–Phonon Coupling, and Self-Trapped Excitonic Emissions in Cs2Ag1–xNaxBiCl6 Double Perovskites. J. Phys. Chem. Lett. 2022, 13, 433–439. [Google Scholar] [CrossRef]
  59. McCall, K.M.; Stoumpos, C.C.; Kostina, S.S.; Kanatzidis, M.G.; Wessels, B.W. Strong Electron–Phonon Coupling and Self-Trapped Excitons in the Defect Halide Perovskites A3M2I9 (A= Cs, Rb; M= Bi, Sb). Chem. Mater. 2017, 29, 4129–4145. [Google Scholar] [CrossRef]
  60. Gong, X.; Voznyy, O.; Jain, A.; Liu, W.; Sabatini, R.; Piontkowski, Z.; Walters, G.; Bappi, G.; Nokhrin, S.; Bushuyev, O. Electron–Phonon Interaction in Efficient Perovskite Blue Emitters. Adv. Mater. 2018, 17, 550–556. [Google Scholar] [CrossRef]
  61. Lenferink, E.J.; LaMountain, T.; Stanev, T.K.; Garvey, E.; Watanabe, K.; Taniguchi, T.; Stern, N.P. Tunable Emission from Localized Excitons Deterministically Positioned in Monolayer p–n Junctions. ACS Photonics 2022, 9, 3067–3074. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal structure of (TDMP)ZnBr4 and Mn2+:(TDMP)ZnBr4. (b) XRD patterns of (TDMP)ZnBr4 with different Mn2+ ion feed ratios. (c) SEM image and element mapping of Mn2+:(TDMP)ZnBr4.
Figure 1. (a) Crystal structure of (TDMP)ZnBr4 and Mn2+:(TDMP)ZnBr4. (b) XRD patterns of (TDMP)ZnBr4 with different Mn2+ ion feed ratios. (c) SEM image and element mapping of Mn2+:(TDMP)ZnBr4.
Materials 17 00562 g001
Figure 2. (a) The absorption spectrum of (TDMP)ZnBr4 under different Mn2+ ion feed ratios. (b) The PLE spectra and PL spectra of Mn2+:(TDMP)ZnBr4. (c) The PL decay lifetime curve of (TDMP)ZnBr4 under different Mn2+ ion feed ratios. (d) Schematic diagram of the photophysical mechanism of Mn2+:(TDMP)ZnBr4.
Figure 2. (a) The absorption spectrum of (TDMP)ZnBr4 under different Mn2+ ion feed ratios. (b) The PLE spectra and PL spectra of Mn2+:(TDMP)ZnBr4. (c) The PL decay lifetime curve of (TDMP)ZnBr4 under different Mn2+ ion feed ratios. (d) Schematic diagram of the photophysical mechanism of Mn2+:(TDMP)ZnBr4.
Materials 17 00562 g002
Figure 3. (a) The temperature-dependent PL spectra of Mn2+:(TDMP)ZnBr4. (b) The activation energy fitting results of Mn2+:(TDMP)ZnBr4. (c) The pseudo-color image of the Mn2+:(TDMP)ZnBr4 temperature-dependent PL spectrum. (d) The fitting result of FWHM and temperature.
Figure 3. (a) The temperature-dependent PL spectra of Mn2+:(TDMP)ZnBr4. (b) The activation energy fitting results of Mn2+:(TDMP)ZnBr4. (c) The pseudo-color image of the Mn2+:(TDMP)ZnBr4 temperature-dependent PL spectrum. (d) The fitting result of FWHM and temperature.
Materials 17 00562 g003
Figure 4. (a) The bandgap structure of (TDMP)ZnBr4. (b) The DOS of (TDMP)ZnBr4. (c) The bandgap structure of Mn2+:(TDMP)ZnBr4. (d) The DOS of Mn2+:(TDMP)ZnBr4.
Figure 4. (a) The bandgap structure of (TDMP)ZnBr4. (b) The DOS of (TDMP)ZnBr4. (c) The bandgap structure of Mn2+:(TDMP)ZnBr4. (d) The DOS of Mn2+:(TDMP)ZnBr4.
Materials 17 00562 g004
Figure 5. (a) The XRD patterns of Mn2+:(TDMP)ZnBr4 before and after 60 days. (b) The PL spectra of Mn2+:(TDMP)ZnBr4 before and after 60 days. (c) The thermal decomposition curve of Mn2+:(TDMP)ZnBr4.
Figure 5. (a) The XRD patterns of Mn2+:(TDMP)ZnBr4 before and after 60 days. (b) The PL spectra of Mn2+:(TDMP)ZnBr4 before and after 60 days. (c) The thermal decomposition curve of Mn2+:(TDMP)ZnBr4.
Materials 17 00562 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peng, C.; Wei, J.; Duan, L.; Tian, Y.; Wei, Q. Mn(II)-Activated Zero-Dimensional Zinc(II)-Based Metal Halide Hybrids with Near-Unity Photoluminescence Quantum Yield. Materials 2024, 17, 562. https://doi.org/10.3390/ma17030562

AMA Style

Peng C, Wei J, Duan L, Tian Y, Wei Q. Mn(II)-Activated Zero-Dimensional Zinc(II)-Based Metal Halide Hybrids with Near-Unity Photoluminescence Quantum Yield. Materials. 2024; 17(3):562. https://doi.org/10.3390/ma17030562

Chicago/Turabian Style

Peng, Chengyu, Jiazheng Wei, Lian Duan, Ye Tian, and Qilin Wei. 2024. "Mn(II)-Activated Zero-Dimensional Zinc(II)-Based Metal Halide Hybrids with Near-Unity Photoluminescence Quantum Yield" Materials 17, no. 3: 562. https://doi.org/10.3390/ma17030562

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