Next Article in Journal
Polarization Splitting at Visible Wavelengths with the Rutile TiO2 Ridge Waveguide
Previous Article in Journal
Modification of Fabrication Process for Prolonged Nitrogen Release of Lignin–Montmorillonite Biocomposite Encapsulated Urea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metal-Ion-Doped Manganese Halide Hybrids with Tunable Emission for Advanced Anti-Counterfeiting

1
Tianjin Key Laboratory of Electronic Materials and Devices, School of Electronics and Information Engineeing, Hebei University of Technology, 5340 Xiping Road, Tianjin 300401, China
2
Baotou Teachers’ College, Inner Mongolia University of Science and Technology, Baotou 014020, China
3
Zhejiang Ruico Advanced Material Co., Ltd., No. 188 Liangshan Road, Huzhou 313018, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(12), 1890; https://doi.org/10.3390/nano13121890
Submission received: 8 June 2023 / Revised: 16 June 2023 / Accepted: 17 June 2023 / Published: 20 June 2023
(This article belongs to the Special Issue Applications of Nanocrystal in LED Lighting and Display)

Abstract

:
Stimuli-responsive luminescent materials have received great attention for their potential application in anti-counterfeiting and information encryption. Manganese halide hybrids have been considered an efficient stimuli-responsive luminescent material due to their low price and adjustable photoluminescence (PL). However, the photoluminescence quantum yield (PLQY) of PEA2MnBr4 is relatively low. Herein, Zn2+- and Pb2+-doped PEA2MnBr4 samples are synthesized, and show an intense green emission and orange emission, respectively. After doping with Zn2+, the PLQY of PEA2MnBr4 is elevated from 9% to 40%. We have found that green emitting Zn2+-doped PEA2MnBr4 could transform to a pink color after being exposed to air for several seconds and the reversible transformation from pink to green was achieved by using heating treatment. Benefiting from this property, an anti-counterfeiting label is fabricated, which exhibits excellent “pink-green-pink” cycle capability. Pb2+-doped PEA2Mn0.88Zn0.12Br4 is acquired by cation exchange reaction, which shows intense orange emission with a high QY of 85%. The PL of Pb2+-doped PEA2Mn0.88Zn0.12Br4 decreases with increasing temperature. Hence, the encrypted multilayer composite film is fabricated relying on the different thermal responses of Zn2+- and Pb2+-doped PEA2MnBr4, whereby the encrypted information can be read out by thermal treatment.

Graphical Abstract

1. Introduction

Counterfeiting and forgery are an ever-growing global issue, causing serious financial losses to governments, companies and individuals. Over the last few decades, counterfeit products have been widely found in daily consumer goods, diplomas, medicines and banknotes [1,2,3]. Information security, including encryption and anticounterfeiting, is becoming an important task that needs to be urgently accomplished. The development of advanced anti-counterfeiting technologies is the only way to solve this dilemma. Due to the merit of visual identifiability and easy operation, luminescent anti-counterfeiting is considered to provide ideal security elements [4,5].
However, the current luminescent materials are easy to mimick due to their monotony response to the excitation light. Stimuli-responsive luminescent materials and multi-mode excited luminescent materials as new kinds of advanced anti-counterfeiting luminescent materials have attracted much attention. Multi-mode excited luminescent materials show multicolor emissions under multiple excitation modes (PL, upconversion luminescence, and long-lasting luminescence), which is hard to mimick [6,7]. However, multi-mode excited luminescent materials usually contain expensive rare earth elements and require different excitation light sources, which is severely restricted by non-portable testing tools. Stimuli-responsive luminescent materials will change their optical properties, including spectra and lifetime, in response to external stimuli such as light, [8] electricity [9], temperature [10,11], force [12], and gases [13]. So far, various organic and inorganic luminescent materials, including organic dyes [14], quantum dots [15,16,17], and transition metal complexes [9], have been used as stimuli-responsive PL materials, but most of these materials suffer from complicated and tedious synthesis and unsatisfactory PL properties.
The newly emerging luminescent materials lead halide perovskites have received great attention due to their fascinating properties such as low cost, high PLQY, and adjustable emission spectrum [18,19,20,21,22,23,24,25,26,27]. Benefiting from their intrinsic vulnerability toward the external environment, some studies have also shown that perovskites could be exploited in anti-counterfeiting, confidential information encryption, and decryption [15,16,17,28]. However, the toxicity of lead and low crystal stability severely impede their further application in information security. To address these issues, much effort has been devoted to developing lead-free perovskites by replacing Pb2+ with Sn2+, Ge2+, Bi2+, and so on. Although Sn2+ and Ge2+ ions have similar electronic structures to Pb2+, tin-based and germanium-based perovskites possess a relatively low PLQY and they are easily oxidized in the air [29,30]. Other lead-free perovskites including Sb-based Cs4CuSb2Cl12 [31], bi-based Cs3Bi2X9 [32], and double perovskites Cs2AgBiBr6 [33], and Cs2AgInCl6 [34], have low defect tolerance and show unsatisfactory PL performance.
Due to the low toxicity, low price, and high abundance of Mn2+, manganese halide hybrids have been considered as an efficient luminescent material [35,36,37]. Besides, Mn2+ is stable in an ambient environment. The emission spectrum of manganese halide hybrids can be tuned by the crystalline field of Mn2+ ions. The tetrahedral coordination Mn2+ emits a green color with a narrow full width at half-maximum (FWHM), whereas the octahedral coordination Mn2+ exhibits orange emission with a broad FWHM [35,36,37]. For Mn-based material, its emission can be adjusted by the ratio of reactants due to this unique luminescent property, which is conducive to a color-changing application. Zang et al. synthesized green emitting 0D organic metal halide C6N2H16MnBr4, which could be transformed into the non-emissive hydrated phase C6N2H16MnBr4(H2O)2 by uptake of water molecules [3]. The reversible color changing was due to the change in the coordination environment. Rewritable PL paper has been constructed relying on this reversible structure transformation, which showed excellent cycle capability. Chen and co-workers found that tetrahedron and trigonal bipyramid could be converted to each other as well. Specifically, green emitting [MnBr2(dppeO2)]n could transform to [MnBr2(dppeO2)(DMF)]n upon exposure to DMF vapor, and reversible conversion could be realized by heating treatment [38]. Han et al. reported a tunable pure-color red/green/blue emission in cesium manganese bromides nanocrystals (NCs) by modulating their crystal field strengths. Red-emitting CsMnBr3 NCs could transform into green-emitting Cs3MnBr5 NCs by adding isopropanol. Furthermore, after contact with water, either CsMnBr3 NCs or Cs3MnBr5 NCs could transform into blue-emitting Cs2MnBr4·2H2O NCs, while Cs2MnBr4·2H2O NCs could transform into the mixture of CsMnBr3 and Cs3MnBr5 phase during dehydration treatment. [39]. Compared to other Mn-based halide materials, PEA2MnBr4 can be regarded as a good candidate for the stimuli-responsive luminescent materials, which does not involve complicated ligands and shows complete reversal transformation. Tang and co-workers prepared a new organic-inorganic hybrid PEA2MnBr4 single-crystal, and they found that PEA2MnBr4 possessed humidity chromism characteristics (it emitted green and pink emission at the water-desorption state and water-adsorption state, respectively) [40]. Due to the visible chromism, this PEA2MnBr4 was used as a marker to check water content (0.02 and 0.05 vol%) in toluene. However, the PLQY of PEA2MnBr4 was relatively low and the cycle stability was unsatisfactory.
Ion doping has been regarded as an effective method to optimize the properties of Mn-based materials [36,37,41]. Here, we report on the synthesis of Zn2+- and Pb2+-doped PEA2MnBr4 by the direct hot-injection method and cation-exchange method, respectively. Zn2+-doped PEA2MnBr4 exhibits a strong green PL band centered at 528 nm with a PLQY of 40%, and it can transform to a pink color after being exposed to air for several seconds. The reversible color change from pink to green can be realized after the heating treatment. Benefiting from this property, the anti-counterfeiting label and trademark were fabricated, which exhibited excellent “pink-green-pink” cycle capability. The Pb2+-doped PEA2Mn0.88Zn0.12Br4 shows broad-band orange emission originating from the 4T16A1 transition of octahedrally coordinated Mn2+ ions with a high QY of 85%. It’s worth noting that different from the Zn2+-doped PEA2MnBr4, the Pb2+ counterpart is stable in air and its PL decreases with increasing temperature. Utilizing the different thermal-responsive properties of Zn2+- and Pb2+-doped PEA2MnBr4, we have fabricated the encrypted multilayer composite film containing both Zn2+- and Pb2+-doped PEA2MnBr4. The encrypted information was concealed by the brighter orange fluorescence of Pb2+-doped PEA2Mn0.88Zn0.12Br4 at the upper layer and it could be read out by thermal treatment.

2. Materials and Methods

Materials: Zinc acetate dihydrate (ZnAc2, 99.9%), lead acetate (PbAc2, 99.99%), manganese acetate (MnAc2, 98%), β-phenylethylamine (PEA, 98%), tetra octyl ammonium bromide (TOAB, 98%), dodecylbenzene sulfonic acid (DBSA, 90%), and octanoic acid (OTAc, 99%) were purchased from Aladdin, Shanghai, China. Lead bromide (PbBr2, 99.99%) was purchased from Macklin, Beijing, China. Manganese (II) bromide (MnBr2, 97%) was purchased from Strem, Newburyport, MA, USA. Bromotrimethylsilane (TMSBr, 97%) was purchased from J&K, Beijing, China. Oleic acid (OA, 85%) was purchased from TCI, Shanghai, China. Xylene (95%) and hexane (97%) were purchased from Kermel, Tianjin, China.
Synthesis of PEA2MnBr4: 0.2 mmoL of MnAc2, 80 μL PEA, 0.3 mL DBSA, 0.2 mL OTAc and 5 mL ODE was placed into a 10 mL three-necked flask. Then, the flask was subjected to vacuum-nitrogen three times and heated to 80 °C to dissolve. After that, the flask was heated to 140 °C, and 150μL TMSBr was injected into the flask. 5 s later, the solution was cooled by an ice-water bath. The crude solution was centrifugated at 5000 rpm for 3 min. After centrifugation, the precipitate was washed with xylene twice and dried under vacuum.
Synthesis of PEA2Mn0.88Zn0.12Br4: The synthetic procedure of PEA2Mn0.88Zn0.12Br4 was similar to that of pure PEA2MnBr4, except for the addition of ZnAc2. The new additions of MnAc2 and ZnAc2 are 0.0294 and 0.0066 g, respectively.
Synthesis of Pb2+-doped PEA2Mn0.88Zn0.12Br4 by hot injection method: The synthetic procedure of Pb-doped PEA2Mn0.88Zn0.12Br4 was similar to that of PEA2Mn0.88Zn0.12Br4, except the addition of PbAc2. The new additions of MnAc2, ZnAc2, and PbAc2 are 0.0277, 0.0044, and 0.0065 g, respectively.
Synthesis of Pb2+-doped PEA2Mn0.88Zn0.12Br4 by cation exchange method: 0.0037 g PbBr2 and 0.0109 g TOAB were dissolved in 0.2 mL xylene to form a lead bromide precursor. Then, the lead bromide precursor was dropped into 3 mL of PEA2Mn0.88Zn0.12Br4 solution (0.124 g PEA2Mn0.88Zn0.12Br4 in 5 mL xylene). The solution was stirred for 10 min at room temperature. Then, the solution was subjected to centrifuge at 5000 rpm for 5 min. 5 mL of Xylene was added to the precipitate, and then centrifuged at 5000 rpm for 5 min. After that, hexane was added to wash the precipitate.
Synthesis of multilayer fluorescent composite films: 5 g of EVA was dissolved in 45 g of xylene to obtain a polymer solution. 0.2 mmol PEA2Mn0.88Zn0.12Br4 was dispersed into 2 mL polymer solution. The mixture was screen-printed on the glass substrate to print the pattern. Next, the polymer solution was spun onto the glass substrate to completely cover the pattern. Then, 0.2 mmol product of Pb2+-doped PEA2Mn0.88Zn0.12Br4 was dispersed into 2 mL polymer solution and spun onto the glass substrate as the upper layer. The glass was placed on the hot plate to observe the encryption and decryption process.
Characterization: The PL and PLE spectra were conducted by a FLS920P spectrometer (Edinburgh Instruments, Livingston, UK). The absolute PLQYs and time-resolved PL decays of the samples were measured by a fluorescence spectrometer (FLS920P, Edinburgh Instruments, UK). The Fourier transform infrared (FTIR) spectrum was conducted on a Thermo-Nicole iS50 FTIR spectrometer with an attenuated total reflection detector (Bruker, Bremen, Germany). The inductively coupled plasma optical emission spectrometry (ICP-OES) measurements were measured by an ICP Optima 8300 (PerkinElmer, Waltham, MA, USA). X-ray photoelectron spectroscopy (XPS) was recorded on a Thermo Scientific K-Alpha spectrometer (Thermo, Waltham, MA, USA). The X-ray diffraction (XRD) measurements were measured on a Rigaku Smart Lab 9 kW (Rigaku Corporation, Tokyo, Japan). Scanning electron microscope (SEM) images were acquired using ZEISS Sigma 500 (ZEISS, Jena, Germany).

3. Results and Discussion

PEA2MnBr4 micro-sized powders were prepared by the hot-injection method (see details in the Materials and Methods Section). We found that the carboxylic acid ligands have an important influence on the composition of products. The X-ray diffraction (XRD) patterns of products prepared by different carboxylic acid ligands are shown in Figure 1. When oleic acid (OA) is used, the impurity phase not belonging to PEA2MnBr4 appears. Instead, most of the diffraction peaks of the product synthesized by dodecylbenzene sulfonic acid (DBSA) are consistent with the main peaks in ICSD 13,856, demonstrating that DBSA facilitates the acquisition of the PEA2MnBr4 phase. This may be because DBSA ligands have an aromatic ring structure and they can stabilize the PEA2MnBr4 phase by π-π stacking interaction with β-phenylethylamine (PEA) molecules. The strong acidity of DBSA may have a tight connection with Mn2+, which may further strengthen the stability of the PEA2MnBr4 phase. However, the DBSA possesses large steric hindrance, which offers incomplete protection of PEA2MnBr4, causing low PLQY and a small amount of impurity at 32.2°, 37.1°, and 38.5°. In order to elevate the PLQY and acquire the pure phase of the PEA2MnBr4, a small amount of octanoic acid (OTAc) with a short carbon-chain was added to further passivate the surface of PEA2MnBr4. As shown in Figure 1 and Figure S1, the product prepared by DBSA and OTAc possesses a pure phase structure, and the PLQY is elevated from 3% to 9%.
For further elevating the PLQY, Zn2+-doped PEA2MnBr4 was prepared by following the procedure of PEA2MnBr4 except adding ZnAc2 into the solution. Here, the feed ratios of Zn2+ ([Zn]/([Zn] + [Mn]) mass ratio) were 5%, 15%, 25%, 35%, and 50%. After being calibrated by inductively coupled plasma optical emission spectrometer (ICP-OES) elemental analysis, the actual ratios of Zn2+ were 4.6%, 12.3%, 23.6%, 32.9%, and 45.7%, respectively (Table S1). The PL spectra of PEA2MnBr4 with different Zn2+ ratios are shown in Figure 2a. The PL intensity of Zn2+-doped PEA2MnBr4 first increases with increasing Zn2+ concentration and reaches the maximum value at the Zn2+ ratio of 15% (PEA2Mn0.88Zn0.12Br4). Further increasing the substitution ratio results in a decline of PL intensity. As shown in Figure 2b,c, the PEA2Mn0.88Zn0.12Br4 shows much more intense green emission compared with the undoped PEA2MnBr4. The PEA2Mn0.88Zn0.12Br4 possesses a high PLQY of 40% compared to 9% for the pristine PEA2MnBr4. The improvement of PLQY is caused by the mitigation of concentration quenching. Because Zn2+ ions prefer to form a tetragonal coordination, Zn2+ ions can replace Mn2+ ions to inhibit the concentration quenching of Mn2+ ions, thus leading to high QY.
XRD patterns of the undoped PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4 are displayed in Figure S2. All the diffraction peaks of PEA2Mn0.88Zn0.12Br4 move to a larger angle compared to that of undoped PEA2MnBr4, demonstrating that the interplanar crystal spacing is reduced. This is because the ionic radius of Mn2+ (0.80 Å) is slightly larger than that of Zn2+ (0.74 Å). Replacing Mn2+ with smaller Zn2+ leads to lattice contraction.
X-ray photoelectron spectroscopy (XPS) was carried out to analyze the electron density around the ions in PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4 (Figure S3). Both the PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4 show strong characteristic Mn 2p and Br 3d peaks. Only the PEA2Mn0.88Zn0.12Br4 shows the Zn 2p peak located at 1022.6 eV, proving the presence of Zn2+ in PEA2Mn0.88Zn0.12Br4. Besides, the binding energies of Mn 2p are reduced from 641.5 eV and 653.7 eV to 641.2 eV and 653.4 eV after Zn2+ doping. Introducing Zn2+ ions leads to the lattice contraction, which increases the electron density around the Mn2+ ions, thus resulting in a decrease in Mn2+ ion binding energy [42]. All the above characterizations suggest that Zn2+ ions have been successfully doped into the PEA2MnBr4.
In addition, scanning electron microscope SEM measurement was also performed to investigate the effect of Zn2+ doping on the morphology. As shown in Figure S4, both the PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4 show irregular morphology. However, after doping with Zn2+, the particle size becomes smaller and the particles are more uniform than the pristine one, demonstrating that Zn2+ doping can retard the reaction speed and homogenize particle size distribution.
To understand the origin of emissions in PEA2Mn0.88Zn0.12Br4, excitation power dependent PL measurements were performed. As shown in Figure S5, the PL intensity is linearly increased with the excitation power density. The origin of the emission can be estimated using the power-law equation, [43]. which is defined as IPL = nLk, where IPL represents the PL intensity; L is the excitation power; coefficient k relates to the recombination mechanism; and n represents the emission efficiency. Here, the fitted k value is 1.1, which represents a free exciton mechanism. We have also measured the PL spectra under different excitation wavelengths. As shown in Figure S5b, all the PL spectra were fixed at the peak of 528 nm as the excitation wavelength changed from 350 to 490 nm. The independence of excitation wavelength demonstrates that the source of this green emission is attributed to the same excited state-Mn2+ emission center (4T16A1).
Interestingly, green-emitting PEA2Mn0.88Zn0.12Br4 becomes pink after being placed in air for several seconds and the pink color can convert back to the green color by thermal treatment or vacuum treatment (Figure 3a). This color-changing phenomenon was associated with water adsorption and desorption, which can be confirmed by the Fourier transform infrared (FTIR) spectra results (Figure 3b). The stretching vibration peak belonging to water can be found at 3400 cm−1. A similar phenomenon has also been observed in PEA2MnBr4 single crystal [40]. Tang et al. believed that green and red emissions originated from tetradentate Mn2+ and trigonal bipyramid Mn2+, respectively [40]. However, detailed structure characterization was not given. Color-changing phenomenon upon contact with water was also found in C6N2H16MnBr4. Zang et al. attributed the state of water molecules adsorption to octahedron coordination geometry [3]. In order to further verify the change in the structure, we have performed XRD measurements at high temperature (120 °C) and room temperature (25 °C) (Figure 3c). Strangely, the high temperature XRD pattern of the sample was similar to the room temperature XRD pattern, except that several diffraction peaks (20.9°, 22.4° and 27.9°) in the high temperature XRD are stronger than that in room temperature XRD. Because no new diffraction peak appears, the presence of red emission cannot be accurately determined. The absence of red-emitting sample diffraction peaks may be caused by the following two reasons: first, the amount of the red-emitting sample is low; second, the crystallinity of the red-emitting sample is poor or the diffraction peaks of the red-emitting sample are much like those of the green-emitting sample. Han et al. also found that it was hard to distinguish the red-emitting CsMnBr3 phase by powder XRD when a small amount of red-emitting CsMnBr3 mixed with green-emitting Cs3MnBr5 NCs [39].
We can deduce that the red emission originates from octahedron coordination geometry based on the following results of Pb2+ doping. To get further insights into the PL mechanism, photoluminescence excitation (PLE) spectra were carried out. As shown in Figure 4, the PLE spectrum monitored at 528 nm is in coincidence with that monitored at 640 nm, demonstrating that energy transfer takes place from the tetrahedron Mn2+ to the octahedron Mn2+. A similar energy transfer process of Mn2+ has also been observed in [Mn(dppeO2)3] − [MnBr4]. [38]. The crystal field strengths around Mn2+ can be calculated according to the PLE spectrum. Here, the Racah parameter (B and C) and crystal field strength (Dq) of Zn2+-doped and undoped samples are shown in Table S2. After doping with Zn2+, the crystal field strength increases, which is due to the lattice contraction.
We further monitored the PL evolution of hydrated PEA2Mn0.88Zn0.12Br4 at high temperature (140 °C) and dehydrated PEA2Mn0.88Zn0.12Br4 in different humidity environments. Upon heating, the intensity of the green PL peak increases significantly and the red emission drops as time goes on (Figure S6a). After heating for 12 s, the red emission completely disappears. Further prolonging the time to 20 s, the green emission drops as well. Hence, the thermal treatment time should be 12–20 s. For the cooling process, the green-emitting PEA2Mn0.88Zn0.12Br4 was exposed to air with different humidity. The evolution of PL spectra during the cooling process was contrary to the heating process. The corresponding response time is about 100 s at a relative humidity of 38% RH. It is worth noting that the response time is related to the humidity, which is extended at relatively lower humidity. The corresponding color coordinates of color changing during the heating and cooling process are labeled and shown in Figure S6c,d.
The heating-cooling cycle stability was further tested. As shown in Figure S7, the emission peak, intensity, and FWHM are well maintained after 60 heating-cooling cycles, which is much better than the PEA2MnBr4 single crystal [40]. To further verify the practicability of the PEA2Mn0.88Zn0.12Br4 film, a series of images were captured at different time intervals. Here, the green channel and red channel of the images were extracted via Image J software to study the change of color (Figure 5). The color index (It) can be calculated by the intensity of the red channel (It(R)) and green channel (It(G)) (It = It(G)/It(R) + It(G)). Also, a time-dependence curve could be drawn according to the following equation y = −ln[(1 − It)/(1 − It0)] = kt. Correspondingly, the change rates of the heating process and cooling process can be obtained, which are 0.025 min−1 and −0.006 min−1, respectively. For the anti-countering application, the letters “H”, “B”, and “T” were successfully coated on the paper, which exhibited reversible changes in green and pink emissions during the heating and cooling process. As shown in Figure 6, the “H”, “B”, and “T” still exhibit a bright green color at high temperature after 60 cycles, demonstrating excellent stability in the anti-counterfeiting application.
Based on the above analysis, the variable structure of PEA2MnBr4 can be attributed to its unique 0D structure. The isolated [MnBr4]2− tetrahedron is vulnerable to the external environment. According to the hard and soft acids and bases (HSAB) principle, the O belongs to a hard base while Mn2+ is a hard acid. Hence, Mn2+ ions prefer to bind to H2O to form the octahedron configuration. Because Pb-Br tends to form the octahedron structure, [PbBr6]4− has a strong inducement effect on the transformation of [MnBr4]2−. In order to acquire highly emitting and stable octahedron PEA2MnBr4, Pb2+ doping was introduced. Firstly, Pb2+-doped PEA2Mn0.88Zn0.12Br4 was synthesized in colloidal solution by introducing PbAc2 into an Mn precursor. Although a bright orange emitting product can be acquired, the product is not uniform (Figure S8a). XRD measurement was performed to identify the impurity. As shown in Figure S8b, these strong diffraction peaks at 14.0°, 20.9°, and 27.9° can be indexed to PEA2ZnBr4 (CCDC 258591), suggesting that phase separation takes place due to the different activity of the Pb2+, Mn2+ and Zn2+ ions. In order to address the activity issue, cation exchange reaction was performed. We have tried several different combinations (PEA2MnBr4 and PEA2PbBr4, PEA2PbBr4 and MnAc2, PEA2PbBr4 and MnBr2, PEA2MnBr4 and PbBr2) (Figure S9) and found that the PLQYs were not high except for the combination of PEA2Mn0.88Zn0.12Br4 and PbBr2 (Figure S10). Hence, we have chosen the reaction of PEA2Mn0.88Zn0.12Br4 and PbBr2 to prepare PEA2Mn0.88Zn0.12Br4 with different Pb-doping concentrations. The actual concentrations of Pb2+ in the final products were determined by ICP and shown in Table S3. The PL spectra of PEA2Mn0.88Zn0.12Br4 with different Pb2+-doping ratios are shown in Figure 7. The highest PL intensity is acquired at the Pb2+ ratio of 9.1%. Actually, the PLQY of the product is as high as 85%. With the Pb2+ ratio changed from 6.3% to 16.7%, the lifetime becomes longer, which is consistent with the PLQY result (Figure S11).
PLE spectra were carried out to investigate the PL mechanism. As shown in Figure 7d, the energy levels of 4T1(P) and 4E(D) maintain at 6% substitution while energy levels of low energy excited states (4A1, 4E(D), 4T2(G), 4T1(G)) disappear. As the Pb2+ concentration changes to 9.1%, the discrete energy levels of 4T1(P) and 4E(D) merge together and another peak located at 414 nm appears. According to previous studies [36,44], this newly appeared peak is assigned to the lead-halide units, indicating that the electronic structure of Pb2+-doped PEA2Mn0.88Zn0.12Br4 is related to the interaction between [MnBr6]4− units and [PbBr6]4− units. Further increasing the Pb2+ doping concentration, the intensity of the PLE peak decreases.
In order to investigate the effect of Pb2+ doping on the structure of the products, XRD was carried out (Figure 8). At low-doping concentration (6.3–16.7%), all of the peaks are still in coincidence with the patterns of PEA2MnBr4 and the diffraction peak positions shift to a smaller angle with the increasing doping concentration. Because the ionic radius of the Pb2+ ion is larger than that of Mn2+, doping Pb2+ ions leads to lattice expansion. As we discussed, the preservation of the PEA2MnBr4 XRD pattern indicates that the content of the octahedron Mn2+ is small. In addition, the intensities of several peaks (20.9°, 22.4° and 27.9°) decrease while the intensity of peaks located at 13.9°, 20.8°, and 27.8° is nearly unchanged as the doping concentration increases. Combined with the above high-temperature XRD analysis, these unchanged peaks may belong to the octahedron Mn2+ and these decreased peaks come from tetrahedron Mn2+. Further increasing to 66.7% substitution, PEA2PbBr4 appears and pure PEA2PbBr4 is acquired at 88.9% substitution. For PEA2PbBr4, the highest PLQY (3.2%) is obtained at 95.2% substitution (Figure 7a), which is much less than the Pb2+-doped PEA2Mn0.88Zn0.12Br4. FTIR was performed to study the influence of the Pb2+ doping effect. As shown in Figure 8b, the peak belonging to the stretching vibration of H2O shifts from 3400 cm−1 to 3450 cm−1 after doping with Pb2+, demonstrating that its coordination environment changes. When PEA2Mn0.88Zn0.12Br4 is exposed to air, H2O molecules can coordinate with Mn2+, changing the configuration of Mn2+ from tetrahedron to octahedron. Due to this coordination, the electron cloud density of O reduces, leading to the decrease in the H-O bond strength. However, when Pb2+ is introduced into the lattice, these unstable tetrahedrons Mn2+ will transform to octahedrons Mn2+ due to the inducive effect. In other words, the H2O molecules are no longer involved in the coordination of octahedron Mn2+. The chemical composition and electronic states of Pb2+-doped PEA2Mn0.88Zn0.12Br4 were further studied by XPS. As shown in Figure S12, after doping with Pb2+, two Pb 4f peaks at 142.3 and 138.4 eV appear, which is attributed to Pb 4f5/2, and Pb 4f7/2. Two Mn 2p peaks at 653.4 and 641.2 eV move 0.3 eV toward lower binding energy to 653.1 and 640.9 eV, and two Zn 2p peaks at 1045.6 and 1022.6 eV both move 0.5 eV toward lower binding energy to 1045.1 and 1022.1 eV. The shift of binding energy can be attributed to the Pb2+ incorporation that prevents the coordination of H2O and Mn2+ and changes the coordination environment of Mn2+. Compared to Br, the O2− possesses stronger electronegativity. After doping with Pb2+, Mn-O bonds transform to Mn-Br octahedrons, causing the increase in the electron cloud density around Mn2+.
In order to explain the influence of the Pb2+ doping effect, the photophysical process of Pb2+-doped PEA2Mn0.88Zn0.12Br4 is depicted in Figure 9. The high PLQY of Pb2+-doped PEA2Mn0.88Zn0.12Br4 is mainly caused by the following three aspects: 1. energy transfer paths are optimized from tetrahedron Mn2+ to Pb2+ then to octahedron Mn2+. 2. the removement of H2O reduces the nonradiative recombination, forming perfect [MnBr6]4−. 3. Pb2+ doping inhibits the concentration quenching of Mn2+. As we discussed, the content of the [MnBr6]4− is low in Pb2+-doped PEA2Mn0.88Zn0.12Br4 when the Pb2+ doping ratio is below 20%, but intense orange emission belonging to octahedron Mn2+ is observed. This can be understood by the PLE results. Both the red emission and green emission possess the same PLE spectrum, demonstrating that energy transfer occurs from tetrahedron Mn2+ to octahedron Mn2+. Interestingly, the discrete energy levels of 4T1(P) and 4E(D) become merged together as the Pb2+ concentration increases. According to the previous study [44], a quasi-continuous conduction band can be formed and spin-forbidden transition is further broken due to the merge of high energy states, which have less constraint on electron transitions, leading to increasement of the density of excitons in excited states. Besides, the transition of lead-halide units is also found in the PLE spectrum, which is placed under the energy levels of 4T1(P) and 4E(D). This newly appeared energy level of lead-halide precludes the energy absorption and transfer of the low energy states (4A1, 4E(D), 4T2(G), 4T1(G)) and directly transfers the energy to the 4T1(G). It is well known that H2O could easily introduce de-excitation channels for the excited state, causing severe PL decay. The removal of H2O can reduce the nonradiative recombination process, leading to effective radiative recombination. In addition, Pb/Mn alloying also dilutes the Mn2+ concentration, inhibiting the concentration quenching of Mn2+. All these factors optimize the energy transfer path and increase the density of excitons, thus leading to high PLQY.
We have also studied the stability of the Pb2+-doped PEA2Mn0.88Zn0.12Br4. To our surprise, Pb2+-doped PEA2Mn0.88Zn0.12Br4 show excellent stability under the ambient environment, which still possesses 80% of the original PL intensity after 25 days. However, the thermal stability is totally different from the PEA2Mn0.88Zn0.12Br4. As shown in Figure S13, the PL intensity decreases sharply as the temperature increases, which is common in luminescent semiconductors. For example, the PL intensity only retains 20% of the original after heating the sample to 160 °C. The different fluorescent temperature-dependencies of PEA2Mn0.88Zn0.12Br4 and Pb2+-doped PEA2Mn0.88Zn0.12Br4 offer great opportunities for information encryption. Multilayer fluorescent composite films were fabricated for the application of information encryption. As shown in Figure 10, patterns of bird, rabbit, and butterfly composed of PEA2Mn0.88Zn0.12Br4 layer are screen printed on a substrate. After that, the Pb2+-doped PEA2Mn0.88Zn0.12Br4 solution is spin-coated on the surface of the PEA2Mn0.88Zn0.12Br4 layer to form a uniform film. The composite film exhibits strong orange emission under UV light at room temperature. When the temperature increases from 20 to 160 °C, the orange emission upper film gradually decreases and the patterns of the bird, rabbit, and butterfly with green emission appear. Finally, when the composite film is heated at 160 °C for 5 s, clear patterns of bird, rabbit, and butterfly are observed. Therefore, in this multilayer composite, the encrypted information can be read out by using a simple heating treatment. Besides, such color-changing PL behaviors can be repeated more than 10 times (Figure S14), demonstrating excellent cycle performance. This indicates that these luminescent materials can be used in document security. At room temperature the information is concealed. When we want to read the information, we can simply heat the document using a hair dryer. After the document cools to room temperature, the information can be hidden again.

4. Conclusions

In summary, we have fabricated Zn2+-doped PEA2MnBr4 and Pb2+-doped PEA2Mn0.88Zn0.12Br4, which possess different PL properties. For PEA2Mn0.88Zn0.12Br4, the reversible structural transformation was responsible for the color-changing phenomenon, which was due to the desorption and absorption of coordinating water. Anti-counterfeiting labels and trademarks have been constructed by taking advantage of this property, which exhibited excellent “pink-green-pink” cycle capability. Pb2+-doped PEA2Mn0.88Zn0.12Br4 was synthesized by the cation exchange reaction, which possessed high-efficiency orange emission. Compared to PEA2Mn0.88Zn0.12Br4, the Pb2+-doped PEA2Mn0.88Zn0.12Br4 was stable in the air, but its PL intensity decreased with the increase in temperature. Multilayer fluorescent composite films were fabricated by spin-coating of Pb2+-doped PEA2Mn0.88Zn0.12Br4 with strong orange emission onto the PEA2Mn0.88Zn0.12Br4 with weak pink emission, where the information was recorded onto the PEA2Mn0.88Zn0.12Br4 film. The encrypted information could be read out by using a simple heating treatment. In addition, the composite films showed excellent cycle stability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13121890/s1. Figure S1: (a)The photograph of PEA2MnBr4 synthesized with DBSA. (b) The photograph of PEA2MnBr4 synthesized with DBSA and OTAc. Table S1: Concentration ratio of Zn2+ in the products by ICP-OES. Figure S2: XRD patterns of PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4. Figure S3: (a) XPS survey spectra of PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4. High-resolutions XPS spectra of Zn2p (b) Mn2p (c) and Br3d (d) in PEA2Mn0.88Zn0.12Br4. HR-XPS spectra of Mn2p (e) and Br3d (f) in PEA2MnBr4. Figure S4: SEM images of PEA2MnBr4 (a) and PEA2Mn0.88Zn0.12Br4 (b). Figure S5: (a) Excitation power-dependent PL spectra for an excitation wavelength of 365 nm. (b) Integrated PL intensities for the emission centered at 528 nm of the PEA2Mn0.88Zn0.12Br4 in relation to the excitation power (0.08–82 μW). (c) Excitation wavelength-dependent PL spectra of PEA2Mn0.88Zn0.12Br4. Table S2: Excitation and crystal field parameters for PEA2MnBr4 and PEA2Mn0.88Zn0.12Br4. Figure S6: PL spectra of PEA2Mn0.88Zn0.12Br4 during heating process (a) and cooling process (b). The transformation of PEA2Mn0.88Zn0.12Br4 in CIE color coordinates during heating process (c) and cooling process (d). Figure S7: (a) FWHM, PL intensity and emission peaks of PEA2Mn0.88Zn0.12Br4 during 60 heating-cooling cycles. PL spectra of the PEA2Mn0.88Zn0.12Br4 at 140 °C (b) and 25 °C (c) at different heating and cooling cycles. Figure S8: (a) PL and PLE spectrum (b) XRD pattern of PEA2Pb0.1Mn0.8Zn0.1Br4. Figure S9: PL spectra of PEA2PbBr4 with different amounts of (a) MnAc2 and (b) MnBr2. (c) PL spectra of PEA2PbBr4 with different amounts of PEA2MnBr4. (d) PL spectra of PEA2MnBr4 with different amounts of PbBr2. Figure S10: (a) PL and (b) PLE spectra of PEA2MnBr4 with PbBr2 dissolved by different ligands. Table S3: Concentration ratio of Mn2+, Zn2+ and Pb2+ in the products by ICP-OES. Figure S11: Decay curves of PEA2Mn0.88Zn0.12Br4 doping Pb2+ in different levels. Figure S12: (a) XPS survey spectra of PEA2Mn0.79Zn0.12Pb0.09Br4 and PEA2Mn0.88Zn0.12Br4. High-resolutions XPS spectra of Zn2p (b) Mn2p (c) Pb4f (d) in PEA2Mn0.79Zn0.12Pb0.09Br4. Figure S13: Air stability (a) and thermal stability (b) of PEA2Mn0.79Zn0.12Pb0.09Br4. Figure S14: PL intensity and emission peaks of the multilayer composite fluorescent film during 10 heating-cooling cycles.

Author Contributions

Conceptualization, H.Z.; data curation, H.Z. and K.L.; formal analysis, Z.D.; funding acquisition, X.L.; investigation, C.F.; methodology, C.S. and H.Z.; resources, X.L.; software, Y.Z.; validation, M.L.; writing—original draft, C.S.; writing—review & editing, C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (52002110), Baotou Teachers’ College Research Launch Fund (01108022/035), S&T Program of Hebei (216Z0601G), and Natural Science Foundation of Hebei Province (A2020202013).

Data Availability Statement

The data related with this work are not publicly available but can be obtained upon reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Feng, P.; Yang, X.; Feng, X.; Zhao, G.; Li, X.; Cao, J.; Tang, Y.; Yan, C.-H. Highly Stable Perovskite Quantum Dots Modified by Europium Complex for Dual-Responsive Optical Encoding. ACS Nano 2021, 15, 6266–6275. [Google Scholar] [CrossRef]
  2. Feng, X.; Sheng, Y.; Ma, K.; Xing, F.; Liu, C.; Yang, X.; Qian, H.; Zhang, S.; Di, Y.; Liu, Y.; et al. Multi-Level Anti-Counterfeiting and Optical Information Storage Based on Luminescence of Mn-Doped Perovskite Quantum Dots. Adv. Opt. Mater. 2022, 10, 2200706. [Google Scholar] [CrossRef]
  3. Liu, H.-L.; Ru, H.-Y.; Sun, M.-E.; Wang, Z.-Y.; Zang, S.-Q. Organic−Inorganic Manganese Bromide Hybrids with Water-Triggered Luminescence for Rewritable Paper. Adv. Opt. Mater. 2022, 10, 2101700. [Google Scholar] [CrossRef]
  4. Hong, W.; Yuan, Z.; Chen, X. Structural Color Materials for Optical Anticounterfeiting. Small 2020, 16, 1907626. [Google Scholar] [CrossRef]
  5. Guillou, O.; Daiguebonne, C.; Calvez, G.; Bernot, K. A Long Journey in Lanthanide Chemistry: From Fundamental Crystallogenesis Studies to Commercial Anticounterfeiting Taggants. Acc. Chem. Res. 2016, 49, 844–856. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, J.-C.; Pan, C.; Zhu, Y.-F.; Zhao, L.-Z.; He, H.-W.; Liu, X.; Qiu, J. Achieving Thermo-Mechano-Opto-Responsive Bitemporal Colorful Luminescence via Multiplexing of Dual Lanthanides in Piezoelectric Particles and its Multidimensional Anticounterfeiting. Adv. Mater. 2018, 30, 1804644. [Google Scholar] [CrossRef]
  7. Li, Y.; Chen, C.; Jin, M.; Xiang, J.; Tang, J.; Zhao, X.; Zheng, J.; Guo, C. Multi-mode excited Cs2NaBiCl6 Based Double Perovskite Phosphor for Anti-counterfeiting. J. Lumin. 2022, 247, 118915. [Google Scholar] [CrossRef]
  8. Kim, J.; Yun, H.; Lee, Y.J.; Lee, J.; Kim, S.-H.; Ku, K.H.; Kim, B.J. Photoswitchable Surfactant-Driven Reversible Shape- and Color-Changing Block Copolymer Particles. J. Am. Chem. Soc. 2021, 143, 13333–13341. [Google Scholar] [CrossRef]
  9. Sun, H.; Liu, S.; Lin, W.; Zhang, K.Y.; Lv, W.; Huang, X.; Huo, F.; Yang, H.; Jenkins, G.; Zhao, Q.; et al. Smart responsive phosphorescent materials for data recording and security protection. Nat. Commun. 2014, 5, 3601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Jiang, K.; Wang, Y.; Cai, C.; Lin, H. Conversion of Carbon Dots from Fluorescence to Ultralong Room-Temperature Phosphorescence by Heating for Security Applications. Adv. Mater. 2018, 30, 1800783. [Google Scholar] [CrossRef]
  11. Wang, Z.; Zhang, Z.; Tao, L.; Shen, N.; Hu, B.; Gong, L.; Li, J.; Chen, X.; Huang, X. Hybrid Chloroantimonates(III): Thermally Induced Triple-Mode Reversible Luminescent Switching and Laser-Printable Rewritable Luminescent Paper. Angew. Chem. Int. Ed. 2019, 58, 9974–9978. [Google Scholar] [CrossRef]
  12. Sagara, Y.; Kato, T. Brightly Tricolored Mechanochromic Luminescence from a Single-Luminophore Liquid Crystal: Reversible Writing and Erasing of Images. Angew. Chem. Int. Ed. 2011, 50, 9128–9132. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Yang, H.; Ma, H.; Bian, G.; Zang, Q.; Sun, J.; Zhang, C.; An, Z.; Wong, W.-Y. Excitation Wavelength Dependent Fluorescence of an ESIPT Triazole Derivative for Amine Sensing and Anti-Counterfeiting Applications. Angew. Chem. Int. Ed. 2019, 58, 8773–8778. [Google Scholar] [CrossRef] [PubMed]
  14. Yagai, S.; Okamura, S.; Nakano, Y.; Yamauchi, M.; Kishikawa, K.; Karatsu, T.; Kitamura, A.; Ueno, A.; Kuzuhara, D.; Yamada, H.; et al. Design amphiphilic dipolar π-systems for stimuli-responsive luminescent materials using metastable states. Nat. Commun. 2014, 5, 4013. [Google Scholar] [CrossRef] [Green Version]
  15. Sun, C.; Gao, Z.; Liu, H.; Wang, L.; Deng, Y.; Li, P.; Li, H.; Zhang, Z.-H.; Fan, C.; Bi, W. One Stone, Two Birds: High-Efficiency Blue-Emitting Perovskite Nanocrystals for LED and Security Ink Applications. Chem. Mater. 2019, 31, 5116–5123. [Google Scholar] [CrossRef]
  16. Sun, C.; Su, S.; Gao, Z.; Liu, H.; Wu, H.; Shen, X.; Bi, W. Stimuli-Responsive Inks Based on Perovskite Quantum Dots for Advanced Full-Color Information Encryption and Decryption. ACS Appl. Mat. Interfaces 2019, 11, 8210–8216. [Google Scholar] [CrossRef] [PubMed]
  17. Tao, J.q.; Zhang, H.; Bi, W.; Liu, X.; Fan, C.; Sun, C. Facile synthesis of Mn2+ doped ultrathin (n = 2) NPLs and their application in anti-counterfeiting. Dalton Trans. 2022, 51, 11021–11028. [Google Scholar] [CrossRef]
  18. Su, S.; Tao, J.; Sun, C.; Xu, D.; Zhang, H.; Wei, T.; Zhang, Z.-H.; Wang, Z.; Fan, C.; Bi, W. Stable and highly efficient blue-emitting CsPbBr3 perovskite nanomaterials via kinetic-controlled growth. Chem. Eng. J. 2021, 419, 129612. [Google Scholar] [CrossRef]
  19. Sun, C.; Gao, Z.; Deng, Y.; Liu, H.; Wang, L.; Su, S.; Li, P.; Li, H.; Zhang, Z.; Bi, W. Orange to Red, Emission-Tunable Mn-Doped Two-Dimensional Perovskites with High Luminescence and Stability. ACS Appl. Mat. Interfaces 2019, 11, 34109–34116. [Google Scholar] [CrossRef]
  20. Lu, M.; Zhang, X.; Zhang, Y.; Guo, J.; Shen, X.; Yu, W.W.; Rogach, A.L. Simultaneous Strontium Doping and Chlorine Surface Passivation Improve Luminescence Intensity and Stability of CsPbI3 Nanocrystals Enabling Efficient Light-Emitting Devices. Adv. Mater. 2018, 30, 1804691. [Google Scholar] [CrossRef]
  21. Zhang, X.; Lu, M.; Zhang, Y.; Wu, H.; Shen, X.; Zhang, W.; Zheng, W.; Colvin, V.L.; Yu, W.W. PbS Capped CsPbI3 Nanocrystals for Efficient and Stable Light-Emitting Devices Using pin Structures. ACS Cent. Sci. 2018, 4, 1352–1359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wei, T.; Wang, L.; Sun, C.; Xu, D.; Tao, J.; Zhang, H.; Han, J.; Fan, C.; Zhang, Z.; Bi, W. Eco-Friendly and Efficient Luminescent Solar Concentrators Based on a Copper(I)-Halide Composite. ACS Appl. Mat. Interfaces 2021, 13, 56348–56357. [Google Scholar] [CrossRef] [PubMed]
  23. Nedelcu, G.; Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Grotevent, M.J.; Kovalenko, M.V. Fast Anion-Exchange in Highly Luminescent Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X. = Cl, Br, I). Nano Lett. 2015, 15, 5635–5640. [Google Scholar] [CrossRef]
  24. Li, X.; Wu, Y.; Zhang, S.; Cai, B.; Gu, Y.; Song, J.; Zeng, H. CsPbX3 Quantum Dots for Lighting and Displays: Room-Temperature Synthesis, Photoluminescence Superiorities, Underlying Origins and White Light-Emitting Diodes. Adv. Funct. Mater. 2016, 26, 2435–2445. [Google Scholar] [CrossRef]
  25. Wei, T.; Lian, K.; Tao, J.; Zhang, H.; Xu, D.; Han, J.; Fan, C.; Zhang, Z.; Bi, W.; Sun, C. Mn-Doped Multiple Quantum Well Perovskites for Efficient Large-Area Luminescent Solar Concentrators. ACS Appl. Mat. Interfaces 2022, 14, 44572–44580. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, F.; Zhong, H.; Chen, C.; Wu, X.-g.; Hu, X.; Huang, H.; Han, J.; Zou, B.; Dong, Y. Brightly Luminescent and Color-Tunable Colloidal CH3NH3PbX3 (X = Br, I, Cl) Quantum Dots: Potential Alternatives for Display Technology. ACS Nano 2015, 9, 4533–4542. [Google Scholar] [CrossRef]
  27. Tao, J.; Sun, C.; Zhang, H.; Wei, T.; Xu, D.; Han, J.; Fan, C.; Zhang, Z.-H.; Bi, W. Perovskite energy funnels for efficient white emission. J. Colloid Interface Sci. 2022, 608, 1202–1211. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, C.; Wang, B.; Li, W.; Huang, S.; Kong, L.; Li, Z.; Li, L. Conversion of invisible metal-organic frameworks to luminescent perovskite nanocrystals for confidential information encryption and decryption. Nat. Commun. 2017, 8, 1138. [Google Scholar] [CrossRef] [Green Version]
  29. Jellicoe, T.C.; Richter, J.M.; Glass, H.F.J.; Tabachnyk, M.; Brady, R.; Dutton, S.E.; Rao, A.; Friend, R.H.; Credgington, D.; Greenham, N.C.; et al. Synthesis and Optical Properties of Lead-Free Cesium Tin Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2016, 138, 2941–2944. [Google Scholar] [CrossRef] [Green Version]
  30. Wu, X.; Song, W.; Li, Q.; Zhao, X.; He, D.; Quan, Z. Synthesis of Lead-free CsGeI3 Perovskite Colloidal Nanocrystals and Electron Beam-induced Transformations. Chem. Asian J. 2018, 13, 1654–1659. [Google Scholar] [CrossRef]
  31. Cai, T.; Shi, W.; Hwang, S.; Kobbekaduwa, K.; Nagaoka, Y.; Yang, H.; Hills-Kimball, K.; Zhu, H.; Wang, J.; Wang, Z.; et al. Lead-Free Cs4CuSb2Cl12 Layered Double Perovskite Nanocrystals. J. Am. Chem. Soc. 2020, 142, 11927–11936. [Google Scholar] [CrossRef] [PubMed]
  32. Ali, R.F.; Andreu, I.; Gates, B.D. Green solvent assisted synthesis of cesium bismuth halide perovskite nanocrystals and the influences of slow and fast anion exchange rates. Nanoscale Adv. 2019, 1, 4442–4449. [Google Scholar] [CrossRef] [Green Version]
  33. Dey, A.; Richter, A.F.; Debnath, T.; Huang, H.; Polavarapu, L.; Feldmann, J. Transfer of Direct to Indirect Bound Excitons by Electron Intervalley Scattering in Cs2AgBiBr6 Double Perovskite Nanocrystals. ACS Nano 2020, 14, 5855–5861. [Google Scholar] [CrossRef]
  34. Liao, Q.; Chen, J.; Zhou, L.; Wei, T.; Zhang, L.; Chen, D.; Huang, F.; Pang, Q.; Zhang, J.Z. Bandgap Engineering of Lead-Free Double Perovskite Cs2AgInCl6 Nanocrystals via Cu2+-Doping. J. Phys. Chem. Lett. 2020, 11, 8392–8398. [Google Scholar] [CrossRef]
  35. Guan, L.-Q.; Shi, S.; Niu, X.-W.; Guo, S.-C.; Zhao, J.; Ji, T.-M.; Dong, H.; Jia, F.-Y.; Xiao, J.-W.; Sun, L.-D.; et al. All-Inorganic Manganese-Based CsMnCl3 Nanocrystals for X-ray Imaging. Adv. Sci. 2022, 9, 2201354. [Google Scholar] [CrossRef]
  36. Meng, Q.; Zhou, L.; Pang, Q.; He, X.; Wei, T.; Zhang, J.Z. Enhanced Photoluminescence of All-Inorganic Manganese Halide Perovskite-Analogue Nanocrystals by Lead Ion Incorporation. J. Phys. Chem. Lett. 2021, 12, 10204–10211. [Google Scholar] [CrossRef]
  37. Hao, X.; Liu, H.; Ding, W.; Zhang, F.; Li, X.; Wang, S. Zn2+-Doped Lead-Free CsMnCl3 Nanocrystals Enable Efficient Red Emission with a High Photoluminescence Quantum Yield. J. Phys. Chem. Lett. 2022, 13, 4688–4694. [Google Scholar] [CrossRef]
  38. Wu, Y.; Zhang, X.; Xu, L.-J.; Yang, M.; Chen, Z.-N. Luminescent Vapochromism Due to a Change of the Ligand Field in a One-Dimensional Manganese(II) Coordination Polymer. Inorg. Chem. 2018, 57, 9175–9181. [Google Scholar] [CrossRef] [PubMed]
  39. Kong, Q.; Yang, B.; Chen, J.; Zhang, R.; Liu, S.; Zheng, D.; Zhang, H.; Liu, Q.; Wang, Y.; Han, K. Phase Engineering of Cesium Manganese Bromides Nanocrystals with Color-Tunable Emission. Angew. Chem. Int. Ed. 2021, 60, 19653–19659. [Google Scholar] [CrossRef] [PubMed]
  40. Gao, W.; Leng, M.; Hu, Z.; Li, J.; Li, D.; Liu, H.; Gao, L.; Niu, G.; Tang, J. Reversible luminescent humidity chromism of organic–inorganic hybrid PEA2MnBr4 single crystals. Dalton Trans. 2020, 49, 5662–5668. [Google Scholar] [CrossRef] [PubMed]
  41. Yang, F.; Yang, M.; Liu, X.; Yang, J.; Wu, Z.; Liang, X.; Lv, C.; Xiang, W. High Quantum Efficiency Zn2+ Doped Lead Manganese Halide @PET Film Used for Wide Color Gamut Backlit Display. Adv. Funct. Mater. 2023, 2303340. [Google Scholar] [CrossRef]
  42. Chen, W.; He, Q.; He, Z.; Wang, Q.; Ding, J.; Huang, Q.; Liang, X.; Chen, Z.; Xiang, W. Environmental-Friendly Cs3MnxZn(1−x)Br5@PS Films Used for Wide Color Gamut Backlight Displays. ACS Sustain. Chem. Eng. 2022, 10, 5333–5340. [Google Scholar] [CrossRef]
  43. Yin, J.; Naphade, R.; Gutiérrez Arzaluz, L.; Brédas, J.-L.; Bakr, O.M.; Mohammed, O.F. Modulation of Broadband Emissions in Two-Dimensional 〈100〉-Oriented Ruddlesden–Popper Hybrid Perovskites. ACS Energy Lett. 2020, 5, 2149–2155. [Google Scholar] [CrossRef]
  44. Li, Y.; Wang, C.; Xu, G.; Luo, G.; Deng, Z. Luminescence Enhancement of CsMnBr3 Nanocrystals through Heterometallic Doping. J. Phys. Chem. Lett. 2023, 14, 2006–2011. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of PEA2MnBr4 synthesized with different ligands.
Figure 1. XRD patterns of PEA2MnBr4 synthesized with different ligands.
Nanomaterials 13 01890 g001
Figure 2. (a) PL spectra of PEA2Mn1−xZnxBr4 (x = 0 − 0.5) excited at 365 nm. The photographs of PEA2MnBr4 (b) and PEA2Mn0.88Zn0.12Br4 (c) excited at 365 nm.
Figure 2. (a) PL spectra of PEA2Mn1−xZnxBr4 (x = 0 − 0.5) excited at 365 nm. The photographs of PEA2MnBr4 (b) and PEA2Mn0.88Zn0.12Br4 (c) excited at 365 nm.
Nanomaterials 13 01890 g002
Figure 3. (a) The photographs of PEA2Mn0.88Zn0.12Br4 under natural light and UV light of 365 nm. (b) FTIR spectrum of PEA2Mn0.88Zn0.12Br4. (c) XRD patterns of PEA2Mn0.88Zn0.12Br4 at 25 °C and 120 °C.
Figure 3. (a) The photographs of PEA2Mn0.88Zn0.12Br4 under natural light and UV light of 365 nm. (b) FTIR spectrum of PEA2Mn0.88Zn0.12Br4. (c) XRD patterns of PEA2Mn0.88Zn0.12Br4 at 25 °C and 120 °C.
Nanomaterials 13 01890 g003
Figure 4. PLE and PL spectra of PEA2Mn0.88Zn0.12Br4 under water-desorption (a) and water-adsorption (b).
Figure 4. PLE and PL spectra of PEA2Mn0.88Zn0.12Br4 under water-desorption (a) and water-adsorption (b).
Nanomaterials 13 01890 g004
Figure 5. The photographs of PEA2Mn0.88Zn0.12Br4 film under UV light during the heating process (a) and cooling process (b), the corresponding red channel and green channel extracted from the photographs are shown in the middle and bottom. (c) The general calculation procedure of the R value (It(R)) and G (It(G)) value from photographs. (d) The calculation of the change rate of patterns during the heating process and cooling process by linear fittings.
Figure 5. The photographs of PEA2Mn0.88Zn0.12Br4 film under UV light during the heating process (a) and cooling process (b), the corresponding red channel and green channel extracted from the photographs are shown in the middle and bottom. (c) The general calculation procedure of the R value (It(R)) and G (It(G)) value from photographs. (d) The calculation of the change rate of patterns during the heating process and cooling process by linear fittings.
Nanomaterials 13 01890 g005
Figure 6. The photographs of the “HBT” letter pattern in 60 heating-cooling cycles.
Figure 6. The photographs of the “HBT” letter pattern in 60 heating-cooling cycles.
Nanomaterials 13 01890 g006
Figure 7. (a) PL spectra, (b) PLE spectra and (c) PLQY of PEA2Mn0.88Zn0.12Br4 with different Pb content. (d) PL and PLE spectrum of PEA2Mn0.79Zn0.12Pb0.09Br4.
Figure 7. (a) PL spectra, (b) PLE spectra and (c) PLQY of PEA2Mn0.88Zn0.12Br4 with different Pb content. (d) PL and PLE spectrum of PEA2Mn0.79Zn0.12Pb0.09Br4.
Nanomaterials 13 01890 g007
Figure 8. (a) XRD patterns of PEA2Mn0.88Zn0.12Br4 with different Pb content (Red asterisk represent the PEA2PbBr4 phase and black asterisk represent PEA2MnBr4 phase). (b) FTIR spectra of PEA2Mn0.88Zn0.12Br4 and PEA2Mn0.79Zn0.12Pb0.09Br4.
Figure 8. (a) XRD patterns of PEA2Mn0.88Zn0.12Br4 with different Pb content (Red asterisk represent the PEA2PbBr4 phase and black asterisk represent PEA2MnBr4 phase). (b) FTIR spectra of PEA2Mn0.88Zn0.12Br4 and PEA2Mn0.79Zn0.12Pb0.09Br4.
Nanomaterials 13 01890 g008
Figure 9. Energy states splitting and optical transitions in PEA2Mn0.88Zn0.12Br4 (a) and PEA2Mn0.79Zn0.12Pb0.09Br4 (b).
Figure 9. Energy states splitting and optical transitions in PEA2Mn0.88Zn0.12Br4 (a) and PEA2Mn0.79Zn0.12Pb0.09Br4 (b).
Nanomaterials 13 01890 g009
Figure 10. (a) Schematic illustration of multilayer fluorescent composite films made by screen printing technology. (b) The photographs of the multilayer composite film obtained at different temperatures under UV irradiation.
Figure 10. (a) Schematic illustration of multilayer fluorescent composite films made by screen printing technology. (b) The photographs of the multilayer composite film obtained at different temperatures under UV irradiation.
Nanomaterials 13 01890 g010
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

Sun, C.; Zhang, H.; Deng, Z.; Fan, C.; Liu, X.; Luo, M.; Zhao, Y.; Lian, K. Metal-Ion-Doped Manganese Halide Hybrids with Tunable Emission for Advanced Anti-Counterfeiting. Nanomaterials 2023, 13, 1890. https://doi.org/10.3390/nano13121890

AMA Style

Sun C, Zhang H, Deng Z, Fan C, Liu X, Luo M, Zhao Y, Lian K. Metal-Ion-Doped Manganese Halide Hybrids with Tunable Emission for Advanced Anti-Counterfeiting. Nanomaterials. 2023; 13(12):1890. https://doi.org/10.3390/nano13121890

Chicago/Turabian Style

Sun, Chun, Hu Zhang, Zhihui Deng, Chao Fan, Xiaohui Liu, Mingming Luo, Yiwei Zhao, and Kai Lian. 2023. "Metal-Ion-Doped Manganese Halide Hybrids with Tunable Emission for Advanced Anti-Counterfeiting" Nanomaterials 13, no. 12: 1890. https://doi.org/10.3390/nano13121890

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