Next Article in Journal
Research Progress and Future Perspectives on Photonic and Optoelectronic Devices Based on p-Type Boron-Doped Diamond/n-Type Titanium Dioxide Heterojunctions: A Mini Review
Previous Article in Journal
Enhancing the Interfacial Adhesion of a Ductile Gold Electrode with PDMS Using an Interlocking Structure for Applications in Temperature Sensors
Previous Article in Special Issue
Photothermal and Hydrophobic Surfaces with Nano-Micro Structure: Fabrication and Their Anti-Icing Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modulating the Afterglow Time of Mn2+ Doped Metal Halides and Applications in Advanced Optical Information Encryption

1
School of Chemistry and Chemical Engineering, Nantong University, Nantong 226019, China
2
School of Public Health, Nantong University, Nantong 226019, China
3
School of Intelligent Manufacturing and Electronic Engineering, Wenzhou University of Technology, Wenzhou 325035, China
4
School of Computer and Electronic Information/School of Artificial Intelligence, Nanjing Normal University, Nanjing 210023, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(13), 1002; https://doi.org/10.3390/nano15131002
Submission received: 26 May 2025 / Revised: 25 June 2025 / Accepted: 25 June 2025 / Published: 28 June 2025
(This article belongs to the Special Issue Photofunctional Nanomaterials and Nanostructure, Second Edition)

Abstract

Mn2+ doped metal halide that can be grown by a facile solution reaction is a promising low-cost afterglow material. However, the afterglow mechanism is still elusive. Using a facile method to modulate afterglow time is still to be explored. In this work, we reveal that the afterglow of Cs2Na0.2Ag0.8InCl6:y%Mn can be significantly modulated by Mn2+ concentration. We propose that replacing Ag+ with Mn2+ leads to the appearance of interstitial Ag+, which temporally store the photogenerated electrons ( A g + + e A g ). After the removal of excitation, the gradual recombination between residual holes and stored electrons [ h + + A g + + e h ν + A g + ] explains the afterglow. However, excessive Mn2+ doping at interstitial sites does not bring about more interstitial Ag+ but instead introduces nonradiative traps. Therefore, as the Mn2+ concentration increases, the afterglow time increases from 350 s to 530 s and then decreases to 230 s, reaching a maximum at y = 40. Thus, a dynamic optical information storage and encryption application is demonstrated based on the modulated afterglow time.

1. Introduction

Afterglow phosphors have exhibited a wide range of applications in the fields of advanced anti-counterfeiting, optical information storage, optical imaging, biological detection, lighting etc. Very recently, Lv et al. reported an appealing deep-trap ultraviolet storage phosphor, ScBO3:Bi3+, which exhibited an ultra-narrowband light emission centered at 299 nm [1]. Based on the unique spectral features and trap distribution, controllable optical information read-out is demonstrated via external light or heat manipulation, highlighting the great potential for advanced optical storage application in bright environments. Sheng et al. proposed the co-multiplexing spectral and temporal dimensions based on photoluminescence (PL) and corresponding afterglow at four different wavelengths [2]. Each emission color shows four possible emissive modes (44), representing the maximum encoding capacity of 8 bits at each pixel. Furthermore, photonic synapses that emulate the human visual system are attracting research interest [3]. Afterglow also shows application potential in photonic synapses [4,5]. Lu et al. proposed an optically stimulated optical-response artificial synapse based on the afterglow material (SrAl2O4:Eu2+, Dy3+) [4]. The output intensity gradually decayed to initial state, suggesting a photo-memory characteristic, which was used to imitate the typical excitatory postsynaptic behavior of a biological synapse. Zhang et al. developed a frosted luminescent solar concentrator (FLSC) by incorporating the Si5.9Al0.1O0.1N7.9:Eu2+ phosphor and the afterglow SrAl2O4:Eu2+, Dy3+ phosphor into a thiol-ene polymer [6]. In addition to photovoltaic modules, the FLSC also acts as a zero-energy nightscape lighting device due to the persistent luminescence [6].
Until now, several afterglow phosphors have already been commercialized, such as white emitting (Li,Na)8Al6Si6O24(Cl,S)2:Ti and green emitting SrAl2O4:Eu2+. Although the oxides are highly stable, their large lattice formation energy suggests a high synthesis temperature and huge energy consumption. Besides, the available emission wavelengths are also relatively limited. Thus, it is still necessary to develop more afterglow phosphors. The ability of carbon dots to emit afterglow has attracted research interest. However, their afterglow efficiency and lifetime need further improvement [7]. Owing to the high absorption coefficient, tunable bandgap, defect-tolerant nature, high photoluminescence quantum yield (PLQY), high charge-carrier mobility, and simple solution synthesis, inorganic lead halide perovskites have attracted attention in the field of afterglow [8,9,10,11]. Zhang et al. replaced Pb2+ with appropriate lanthanide ions (Ln3+) in CsPbBr3 nanocrystals, which were embedded inside an amorphous transparent medium [12]. The CsPbBr3:Ln3+ nanocrystals emitted a green afterglow lasting 1800 s. However, the instability and the toxicity of lead ions have limited their practical applications.
During the past several years, lead-free metal halides such as Cs2AgInX6, Cs2AgBiX6, and Cs2NaInCl6 and their low dimensional derivatives have been widely investigated due to their lower toxicity and better stability compared to CsPbX3 [13,14,15,16,17,18]. Among them, Cs2AgInCl6 is one of the most noticeable metal halides due to its direct bandgap. Zheng et al. synthesized Cs2NaxAg1-xInCl6:Mn via a facile hydrothermal reaction at 180 °C, which exhibited a red afterglow [19]. Yang et al. reported the synthesis of Mn2+/Yb3+/Er3+ tri-doped Cs2Ag0.8Na0.2InCl6, which exhibited both visible and NIR afterglow due to the release of long-lived trapped electrons to Mn2+ and Er3+ ions [20]. Multi-modal emissions of the tri-doped metal halides enable a five-level anti-counterfeiting strategy. Therefore, the Mn2+ doped metal halides provide a promising platform for the development of afterglow phosphors. However, the nature of the long-lived trap (LLT) is still elusive. A facile method to modulate their afterglow time is still to be explored.
In this work, Cs2Na0.2Ag0.8InCl6:y%Mn crystals were synthesized by a solvothermal method, which showed afterglow emission at about 600 nm. As the y value increased from 20 to 60, the afterglow time increased first and then decreased, with a maximum of 530 s at y = 40. To understand the Mn2+ concentration dependent afterglow time, steady-state photoluminescence (PL) spectra and time resolved PL, the ultraviolet–visible (UV–vis) diffuse reflectance spectra of Cs2Na0.2Ag0.8InCl6:y%Mn crystals were measured and analyzed. Their crystal structures and band structures were analyzed based on X-ray diffraction (XRD) patterns, high resolution transmission electron microscopy (HRTEM) images, and ultraviolet photoelectron spectra (UPS). On the one hand, the substitution of Ag+ by Mn2+ leads to the appearance of interstitial Ag+, which temporally store the photogenerated electrons due to the photoreduction reaction. The release of carriers results in afterglow. On the other hand, the Mn2+ doping introduces nonradiative traps near the band edges, leading to the monotonic descending of both the PL and afterglow intensity with increasing Mn2+ concentration. In particular, excessive Mn2+ doping at interstitial sites does not bring about more interstitial Ag+ but rather introduces nonradiative traps. Thus, the afterglow time first increases then decreases as the Mn2+ doping concentration increases. Based on the modulated afterglow time, an optical information storage and encryption application is proposed.

2. Materials and Methods

2.1. Materials

Cesium chloride (CsCl, 99.99%), silver chloride (AgCl, 99.5%), indium chloride (InCl3, 99.99%), and manganese chloride (MnCl2, 99%) were obtained from Shanghai Aladdin Ltd. (Shanghai, China) Ethylenediaminetetraacetic acid disodium salt (EDTA-2Na·2H2O, 99%), hydrochloric acid (HCl, 12 M), polymethyl methacrylate (PMMA, M.W. 35,000), and N, N-dimethylformamide (DMF, 99.7%) were purchased from Sinopharm Chemical Reagent Co, Ltd. (Shanghai, China). All chemicals were used as received without further purification.

2.2. Synthesis of Cs2Na0.2Ag0.8InCl6:y%Mn

To synthesize Cs2Na0.2Ag0.8InCl6:y%Mn, 2 mmol CsCl, 0.8 mmol AgCl, 1 mmol InCl3, y mmol MnCl2, and 0.1 mmol EDTA-2Na·2H2O were loaded in a 25 mL polytetrafluoroethylene vessel, followed by the addition of 12 mL of concentrated hydrochloric acid. The vessel was sealed in a stainless-steel autoclave, which was then kept in an oven at 180 °C for 12 h. The crystallization was started by slow-cooling to 30 °C at a rate of ~3 °C/h. The bottom crystals were rinsed three times with isopropanol and then dried on filter paper at room temperature. The crystals with Mn2+ doping were lightly crushed and sieved with 50 mesh to obtain powder crystals and then stored in hooded vials for further characterizations.

2.3. Structural and Optical Characterizations

Powder X-ray diffraction (XRD) patterns were collected using an X-ray diffractometer (Bruker D8, Billerica, MA, USA) coupled with Cu-K radiation. The scanning rate was set at 8°/min with a step size of 0.02°. High resolution transmission electron microscopy (HRTEM) images were obtained using a transmission electron microscope (TEM, JEOL, JEM-2100PLUS, Tokyo, Japan). In the HRTEM measurements, we precisely adjusted the electron beam direction at 0.1°/step to align the crystal plane axis. Ultraviolet–visible (UV–vis) diffuse reflectance spectra in the range of 200–800 nm were recorded by a spectrophotometer (Cary 5000, Varian, Palo Alto, CA, USA). X-ray and ultraviolet photoelectron spectroscopy were measured by Thermo ESCALAB 250Xi (Sydney, Australia) with an Al Ka source. To determine VBM accurately, a linear tangent was used to fit the steepest part of the leading edge of UPS, where the intensity starts to increase. Then, this tangent was extrapolated to intersect the baseline. The BE at this intersection is the VBM. The measurements of the PL and afterglow fluorescence were conducted in a spectrofluorometer (FLS-1000 spectrometer, Edinburgh Instrument, Edinburgh, Scotland). The excitation wavelength was set at 365 nm. For measurements of the PL lifetime, a μs flash lamp was used as the excitation source. Afterglow time was obtained with a phosphorophotometer (Pr-305, HangZhou ZheJiang University Sensing Instruments Co., Ltd., Hangzhou, China). All of the afterglow measurements were conducted in ambient air. The afterglow time is defined as the period between the removal of UV excitation and when the afterglow intensity decreases to background noise level.

3. Results and Discussion

The actual Mn2+ concentrations of Cs2Na0.2Ag0.8InCl6:y%Mn (y = 0, 20, 30, 40, 50, 60) crystals are determined by X-ray photoelectron spectroscopy (Figures S1 and S2 and Table S1 in Supplementary Materials). The Cs2Na0.2Ag0.8InCl6:y%Mn were continuously irradiated by a 365 nm lamp for 10 min. Luminescence and afterglow photos of them are shown in Figure 1a. As expected, Cs2Na0.2Ag0.8InCl6 without Mn2+ doping does not exhibit any afterglow. As the Mn2+ concentration increases, the visible afterglow time first increases then decreases, reaching a maximum for nominal y = 40. The afterglow spectra captured after removing the excitation are shown in Figure 1b. All of the spectra mainly show an afterglow emission peak at ca. 600 nm owing to the Mn2+ 4T16A1 transition [19,20,21]. The transition of the d-orbital electrons associated with the Mn2+ ions is spin-forbidden [22]. Only when the Mn2+ ion is doped in a crystal is the transition allowed, due to the coupling with the host crystal’s field [23]. The finite hybridization with the host electronic states gives rise to an energy spread of Mn d states. The observation of broadband emissions at ca. 600 nm verifies the doping of Mn2+ into the metal halide crystal. As the Mn2+ concentration increases, the afterglow intensity decreases monotonically (Figure 1c). The afterglow decay traces are plotted in Figure 1d. With increasing Mn2+ doping, the afterglow time first increases from 305 s to 530 s, then decreases to 230 s (inset in Figure 1d). To obtain a more accurate understanding of the relationship between afterglow time and y value, we also prepared samples with y = 10, 35, and 45. Their afterglow decay traces are presented in Figure S3. These results manifest that both the afterglow intensity and time can be regulated by Mn2+ doping content.
In order to gain more insight into the Mn2+ doping concentration dependent afterglow, the steady-state PL spectra and time-resolved PL decay curves were measured. As shown in Figure 2a, the overall PL intensity decreases monotonically with the increasing of Mn2+ doping concentration, which is consistent with variation of afterglow intensity, indicating the Mn2+ doping introduces nonradiative traps. The ultrabroad band involves disparate transitions obviously. To visually show the change of each peak, a single PL spectrum is Gaussian fitted into three sub-bands at about 466 nm, 518 nm, and 590 nm (Figure 2b and Figure S4), which are labeled as blue, green, and red peak, respectively. According to previous reports [21,24,25], the blue and green peaks with large full widths at half maximum (FWHM) and big Stokes-shift are attributed to self-trapped excitons (STE) caused by Jahn–Teller deformation of different metal chloride octahedra, i.e., [InCl6] and [AgCl6]. The red peak is due to the 4T1 to 6A1 transition of Mn2+ [19,20,21]. As shown in Figure 2c, the intensity of blue and green peak relative to red peak increase then decrease as the Mn2+ concentration increases. Apart from the change in PL intensity, the red peak undergoes a gradual shift from 581.8 nm to 589.1 nm with the increasing Mn2+ concentration.
The time-resolved PL spectra were measured at three emission wavelengths, which were fitted by tri-exponential decay functions (Figure 2d–f and S5–S7 and Tables S2–S4). The average PL lifetimes of blue and green emissions are about 5 to 10 micro-seconds, matching well with the typical lifetime of STEs [24,25,26]. The spin-forbidden Mn2+ d−d transition is partially relaxed due to the strong coupling with the host crystal [11,23,27], which usually occurs at time scale of milliseconds. The average PL lifetime of the red emission is about 7 ms, further confirming that the 590 nm-emission is related to the Mn2+ d–d transition [11,28]. As shown in Figure 2g–i, the average PL lifetimes of all three emissions first increase then decrease with the increasing of the Mn2+ concentration, which is consistent with the dependence of afterglow time on doping concentration.
To understand the above observations in afterglow and PL, crystal structures of the Cs2Na0.2Ag0.8InCl6:y%Mn were investigated. As shown in Figure 3a, XRD patterns of the Cs2Na0.2Ag0.8InCl6:y%Mn show diffraction peaks similar to those of the undoped Cs2AgInCl6 with a space group Fm 3 ¯ m, indicating the crystal is a face-centered cubic structure. It is worth noting that the characteristic peak corresponding to diffraction of (220) plane shows significant deviations with different amounts of Mn2+ doping (Figure 3b), suggesting different degrees of lattice distortion. This characteristic peak undergoes a shift to smaller angles first and then returns, implying the lattice of co-doped crystal first contracts and then expands with the increasing of the Mn2+ concentration. By analyzing the crystal structure, the Ag and In are found to be possible substitution sites [29,30]. The radius of Mn2+ (0.90 Å) is smaller than that of Ag+ (1.15 Å). The replacement of Ag+ by Mn2+ leads to the lattice shrinkage. Substitution of In3+ (0.8 Å) by Mn2+ (0.90 Å) leads to an expansion of the lattice. Excessive doping brings interstitial Mn2+, also leading to an expansion of the lattice. The variation of lattice spacing was further verified by HRTEM results (Figure 3c,d).
The Mn2+ doping-induced lattice distortion leads to the change of band structure. The UV–vis diffuse reflectance spectra were measured to estimate the optical band gaps (Figure 4a). The absorbance at 497 nm caused by Mn2+ transition to monotonously increase with y% value (Figure S8). As well established previously, this type of metal halides is a direct bandgap semiconductor [31,32]. As shown in Figure 4b, the direct bandgaps of Cs2Na0.2Ag0.8InCl6:y% with y = 20, 30, 40, 50, 60 are calculated to be 3.463 eV, 3.388 eV, 3.367 eV, 3.357 eV, 3.349 eV, 3.329 eV, respectively, as calculated by the Tauc method [33]. The bandgap decreases with increasing Mn2+ doping (Figure S9), explaining the redshift of the red peak. To gain more insight into the band structures, the valence band positions were measured by ultraviolet photoelectron spectroscopy (UPS). As shown in Figure 4c, the valence band positions show a trend of first increasing and then decreasing, reaching a maximum at y = 40. According to the bandgap values and valence band positions, the band structures of Cs2Na0.2Ag0.8InCl6:y%Mn are schematically shown in Figure 4d.
Understanding the complex role of Mn2+ is the key to explaining the PL and afterglow. On the one hand, the Mn2+ doping leads to lattice deformation. The lattice contraction increases the bandgap while the lattice extension decreases the band gap. On the other hand, replacing monovalent Ag+ or trivalent In3+ with divalent Mn2+ necessarily disrupts local charge neutrality. This charge imbalance must be offset by compensatory defects such as vacancies or antisite defects. The Mn2+ doping, in particular excessive doping, introduces nonradiative traps (NTs) near the band edges, which reduce the bandgap. The synthetical factors result in monotonous narrowing of the bandgap. It is worth noting that the lattice distortion not only changes the bandgap but also affects the formation of STEs in the soft perovskite matrix. The appropriate lattice distortion is favorable to Jahn–Teller deformation and thus encourages STE formation [34,35]. However, excessive doping causes rapid nonradiative transition. Thereby, the relative PL intensity and lifetime corresponding to STE formation first increases, then decreases. Moreover, the Mn2+ doping introduces not only NTs but also LLTs that are responsible for the afterglow. The replacement of Ag+ with Mn2+ leads to the appearance of interstitial Ag+, which plays the role of LLTs. Under UV excitation, the interstitial Ag+ captures the photogenerated electrons, forming Ag, as demonstrated by the XPS results (Figure S10). In other words, the photogenerated electrons are temporally stored. After the removal of UV excitation, the residual holes recombine with the stored electrons, emitting the afterglow. At y = 40, the lattice shrinkage reaches its maximum, implying the highest concentration of interstitial Ag+, which agrees well with the longest afterglow time among the five samples. Excessive Mn2+ doping at interstitial sites does not bring about more LTTs, but rather introduces NTs. Thus, the afterglow time decreases when y exceeds 40.
We would like to point out that the PL and the afterglow are two different photophysical processes. For the PL process, the Mn2+ is excited by energy transfer from the Cs2Na0.2Ag0.8InCl6 without the participation of LLTs. By contrast, the afterglow is caused by the release of electrons trapped by of LLTs to Mn2+. The afterglow time essentially depends on the electron storage by LLTs, which lasts hundreds of seconds. However, they have an inherent connection. The relatively long PL lifetime of Mn2+ (ms scale) is crucial to the appearance of afterglow. If the radiative recombination is ns scale, the excited states return to ground state via radiative transition before the storage by LLTs. Thus, afterglow is absent. Afterglow is usually observed in Mn/rare earth ion doped crystals, since the radiative transitions of Mn/rare earth ions are relatively slow, allowing the storage of excited states before the radiative transition [2,3,4,5,6].
Based on the afterglow time modulated by the Mn2+ doping concentration, we demonstrated an optical information storage and encryption application. Seven disks were fabricated by incorporating the Cs2Na0.2Ag0.8InCl6, Cs2Na0.2Ag0.8InCl6:40%Mn, and Cs2Na0.2Ag0.8InCl6:50%Mn into polydimethylsiloxane (PDMS). As shown in Figure 5, all of the disks showed similar colors under sunlight and similar orange-red fluorescence under 365 nm UV excitation. The information is hidden. After removing the UV excitation, the Cs2Na0.2Ag0.8InCl6 based disk becomes completely dark immediately, while other ones exhibit similar afterglow. By assigning the red disk as “1” and dark disks as “0”, the seven disks represent “1101101”, which are decoded to be “m” in the ASCII table. Furthermore, when the UV off lasts for 100 s, the afterglow of Cs2Na0.2Ag0.8InCl6:50%Mn based disks almost disappears. Consequently, the code turns into “0100100”, representing “$”. Supposing “$” is the only correct information, it is protected by the interference information “m”. Therefore, this dynamic strategy processes a high security for information encryption.

4. Conclusions

In summary, Cs2Na0.2Ag0.8InCl6:y%Mn crystals with afterglow emission at about 600 nm were synthesized by a solvothermal method. The afterglow mechanism is elucidated by establishing a correlation between crystal structure, band structure, and afterglow time. As the nominal y increased from 20 to 60, the afterglow time increased from 350 s to 530 s and then decreased to 230 s, reaching a maximum at y = 40. The steady-state PL spectra, time resolved PL, diffuse reflectance spectra, and UPS of the Cs2Na0.2Ag0.8InCl6:y%Mn crystals of similar size were measured and analyzed. On the one hand, Mn2+ doping leads to lattice deformation, which not only changes the bandgap but also affects the formation of STEs. On the other hand, Mn2+ doping introduces nonradiative traps near the band edges. Thus, both the overall PL intensity and the afterglow intensity monotonically decrease with increasing Mn2+ concentration. Furthermore, the replacement of Ag+ with Mn2+ leads to the appearance of interstitial Ag+, which temporally stores the photogenerated electrons by changing valence. However, excessive Mn2+ doping at interstitial sites does not bring about more interstitial Ag+ but rather introduces nonradiative traps. Thus, the afterglow time first increases then decreases as the Mn2+ doping concentration increases. We propose an optical information storage and encryption application based on the modulated afterglow time.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15131002/s1, Figure S1. XPS survey spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. Figure S2. Cs 3d (a), Ag 3d (b), Na 1s (c), In 3d (d), Cl 2p (e), and Mn 2p (f) core level XPS spectra of Cs2Na0.2Ag0.8InCl6:40%Mn. Table S1. Actual Mn concentration measured by XPS. Figure S3. (a) Afterglow decay traces of Cs2Na0.2Ag0.8InCl6:y%Mn, y = 10, 35, 45. (b) The afterglow lifetime versus y value. Figure S4. Gaussian fittings of the PL spectra of Cs2Na0.2Ag0.8InCl6:y%Mn. Figure S5. Time resolved PL decay curves of the Cs2Na0.2Ag0.8InCl6:y%Mn monitored at 466 nm. Figure S6. Time resolved PL decay curves of the Cs2Na0.2Ag0.8InCl6:y%Mn monitored at 518 nm. Figure S7. Time resolved PL decay curves of the Cs2Na0.2Ag0.8InCl6:y%Mn monitored at 590 nm. Table S2. Fitting parameters for the PL lifetime curves monitored at 466 nm. Table S3. Fitting parameters for the PL lifetime curves monitored at 518 nm. Table S4. Fitting parameters for the PL lifetime curves monitored at 590 nm. Figure S8. Absorbance at 497 nm versus the y value. Figure S9. Band gaps determined by Tauc method versus y value. Error bars represent the standard deviations of repeated measurements. Figure S10. Ag 3d core level XPS spectra of Cs2Na0.2Ag0.8InCl6:40%Mn before and after UV illumination for 20 min.

Author Contributions

Methodology, J.-Q.X.; Investigation, Y.-L.H., Y.-L.Z. and S.-Y.G.; Resources, C.-G.S.; Writing—original draft, Y.-L.H.; Writing—review & editing, Z.-X.G. and C.-G.S.; Project administration, C.-G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Shandong Province (ZR2021YQ32), the Taishan Scholars Program of Shandong Province (tsqn201909117), Scientific research project of Zhejiang Provincial Department of Education (Grant No. Y202250552), and Wenzhou Scientific Research Project (ZG2022008). Partial support was also from the Opening Foundation of Hubei Key Laboratory of Photoelectric Materials and Devices (PMD202401), and Special Fund for Science and Technology Innovation Teams of Shanxi Province.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lv, X.; Liang, Y.; Zhang, Y.; Chen, D.; Shan, X.; Wang, X.-J. Deep-trap ultraviolet persistent phosphor for advanced optical storage application in bright environments. Light Sci. Appl. 2024, 13, 253. [Google Scholar] [CrossRef] [PubMed]
  2. Sheng, Y.; Zhang, Y.; Xing, F.; Liu, C.; Di, Y.; Yang, X.; Wei, S.; Zhang, X.; Liu, Y.; Gan, Z. Co-multiplexing spectral and temporal dimensions based on luminescent materials. Opt. Express 2023, 31, 24667–24677. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Y.; Zha, Y.; Bao, C.; Hu, F.; Di, Y.; Liu, C.; Xing, F.; Xu, X.; Wen, X.; Gan, Z.; et al. Monolithic 2D perovskites enabled artificial photonic synapses for neuromorphic vision sensors. Adv. Mater. 2024, 36, 2311524. [Google Scholar] [CrossRef] [PubMed]
  4. Lu, W.; Chen, Q.; Zeng, H.; Wang, H.; Liu, L.; Guo, T.; Chen, H.; Wang, R. All optical artificial synapses based on long-afterglow material for optical neural network. Nano Res. 2023, 16, 10004–10010. [Google Scholar] [CrossRef]
  5. Li, R.; Yue, Z.; Luan, H.; Dong, Y.; Chen, X.; Gu, M. Multimodal artificial synapses for neuromorphic application. Research 2024, 7, 427. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Zheng, Z.; Cao, X.; Gu, G.; Gan, Z.; Huang, R.; Guo, Y.; Hou, D.; Zhang, X. Highly efficient multifunctional frosted luminescent solar concentrators with zero-energy nightscape lighting. J. Mater. Chem. A 2022, 10, 22145–22154. [Google Scholar] [CrossRef]
  7. Peng, C.; Chen, X.; Chen, M.; Lu, S.; Wang, Y.; Wu, S.; Liu, X.; Huang, W. Afterglow carbon dots: From fundamentals to applications. Research 2021, 2021, 6098925. [Google Scholar] [CrossRef]
  8. Wang, G.; Yang, B.; Mei, S.; Wang, B.; Zhang, Z.; Mao, Y.; Guo, J.; Ma, G.; Guo, R.; Xing, G. Persistent charging of CsPbBr3 perovskite nanocrystals confined in glass matrix. Small 2024, 20, 2307785. [Google Scholar] [CrossRef] [PubMed]
  9. Xiong, P.; Yang, X.; Shen, J. Confinement reaction synthesis of CsPbBr3 quantum dots within long afterglow mesoporous materials as lighting equipment Mater. Today Commun. 2024, 39, 108590. [Google Scholar]
  10. Hai, O.; Pei, M.; Yang, E.; Ren, Q.; Wu, X.; Zhao, Y.; Du, L. Tunable luminescent color from green to blue on long afterglow materials using CsPbBr3 quantum dots. Appl. Surf. Sci. 2021, 568, 150941. [Google Scholar] [CrossRef]
  11. Yang, R.; Ji, H.; Zhao, D.; Zhang, F.; Ji, X.; Wang, M.; Zhang, M.; Jia, M.; Chen, X.; Liu, Y.; et al. Modulation of trap distribution by optimizing Mn2+ doed in CsCdCl3 crystals toward enhanced afterglow performance. Appl. Phys. Lett. 2024, 124, 091904. [Google Scholar] [CrossRef]
  12. Zhang, H.; Yang, Z.; Zhao, L.; Cao, J.; Yu, X.; Yang, Y.M.; Yu, S.; Qiu, J.; Xu, X. Long persistent luminescence from all-inorganic perovskite nanocrystals. Adv. Opt. Mater. 2020, 8, 2000585. [Google Scholar] [CrossRef]
  13. Sheng, Y.; Gui, Q.; Zhang, Y.; Yang, X.; Xing, F.; Liu, C.; Di, Y.; Wei, S.; Cao, G.; Yang, X.; et al. Lead-free perovskites with photochromism and reversed thermochromism for repeatable information writing and erasing. Adv. Funct. Mater. 2024, 34, 2406995. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Yang, X.; Yang, L.; Xing, F.; Liu, C.; Di, Y.; Cao, G.; Wei, S.; Yang, X.; Zhang, X.; et al. Advanced optical information encryption enabled by polychromatic and stimuli-responsive luminescence of Sb-doped double perovskites. Adv. Sci. 2024, 11, 2308390. [Google Scholar] [CrossRef]
  15. Han, Y.; Fang, X.; Shi, Z. Advances in chalcogenide perovskites: Fundamentals and applications. Appl. Phys. Rev. 2024, 11, 021338. [Google Scholar] [CrossRef]
  16. Fu, X.; Li, H.; Yue, H.; Li, Z.; Feng, J.; Zhang, H. Cr3+/Yb3+ codoped Cs2NaInCl6 double perovskites for near-infrared light-emitting diodes. Inorg. Chem. 2025, 64, 8782–8791. [Google Scholar] [CrossRef]
  17. Karmakar, A.; Bernard, G.M.; Meldrum, A.; Oliynyk, A.O.; Michaelis, V.K. Tailorable indirect to direct band-gap double perovskites with bright white-light emission: Decoding chemical structure using solid-state NMR. J. Am. Chem. Soc. 2020, 142, 10780. [Google Scholar] [CrossRef]
  18. Zhao, J.-Q.; Wang, D.-Y.; Yan, T.-Y.; Wu, Y.-F.; Gong, Z.-L.; Chen, Z.-W.; Yue, C.-Y.; Yan, D.; Lei, X.-W. Synchronously improved multiple afterglow and phosphorescence efficiencies in 0D hybrid zinc halides with ultrahigh anti-water stabilities. Angew Chem. 2024, 63, e202412350. [Google Scholar] [CrossRef]
  19. Zheng, W.; Li, X.; Liu, N.; Yan, S.; Wang, X.; Zhang, X.; Liu, Y.; Liang, Y.; Zhang, Y.; Liu, H. Solution-grown chloride perovskite crystal of red afterglow. Angew. Chem. 2021, 60, 24450. [Google Scholar] [CrossRef]
  20. Yang, X.; Sheng, Y.; Zhang, L.; Yang, L.; Xing, F.; Di, Y.; Liu, C.; Hu, F.; Yang, X.; Yang, G.; et al. Five-level anti-counterfeiting based on versatile luminescence of tri-doped double perovskites. Nano Res. 2024, 17, 9971–9979. [Google Scholar] [CrossRef]
  21. Li, X.; Xu, S.; Liu, F.; Qu, J.; Shao, H.; Wang, Z.; Cui, Y.; Ban, D.; Wang, C. Bi and Sb codoped Cs2Ag0.1Na0.9InCl6 double perovskite with excitation-wavelength-dependent dual-emission for anti-counterfeiting application. ACS Appl. Mater. Interfaces 2021, 13, 31031. [Google Scholar] [CrossRef] [PubMed]
  22. Ye, S.; Zhang, J.; Zhang, X.; Wang, X. Mn2+ activated red long persistent phosphors in BaMg2Si2O7. J. Lumin. 2007, 122–123, 914–916. [Google Scholar] [CrossRef]
  23. Hazarika, A.; Layek, A.; De, S.; Nag, A.; Debnath, S.; Mahadevan, P.; Chowdhury, A.; Sarma, D.D. Ultranarrow and widely tunable -induced photoluminescence from single Mn-doped nanocrystals of ZnS-CdS alloys. Phys. Rev. Lett. 2013, 110, 267401. [Google Scholar] [CrossRef]
  24. Zhou, B.; Liu, Z.; Fang, S.; Nie, J.; Zhong, H.; Hu, H.; Li, H.; Shi, Y. Emission mechanism of self-trapped excitons in Sb3+-doped all-inorganic metal halide perovskites. J. Phys. Chem. Lett. 2022, 13, 9140. [Google Scholar] [CrossRef]
  25. Chen, B.; Guo, Y.; Wang, Y.; Liu, Z.; Wei, Q.; Wang, S.; Rogach, A.L.; Xing, G.; Shi, P.; Wang, F. Multiexcitonic emission in zero-dimensional Cs2ZrCl6:Sb3+ perovskite crystals. J. Am. Chem. Soc. 2021, 143, 17599–17606. [Google Scholar] [CrossRef]
  26. Sun, C.; Guo, Y.-H.; Han, S.-S.; Li, J.-Z.; Jiang, K.; Dong, L.-F.; Liu, Q.-L.; Yue, C.-Y.; Lei, X.-W. Three-dimensional cuprous lead bromide framework with highly efficient and stable blue photoluminescence emission. Angew. Chem. 2020, 59, 16465. [Google Scholar] [CrossRef]
  27. Yadav, P.; Khurana, S.; Sapra, S. Doping Mn2+ in hybrid Ruddlesden–Popper phase of layered double perovskite (BA)4AgBiBr8. Nanotechnology 2022, 33, 415706. [Google Scholar] [CrossRef]
  28. Yang, J.; Yuan, X.; Fan, L.; Zheng, Y.Z.; Ma, F.S.; Li, H.B.; Zhao, J.L.; Liu, H.L. Enhancing Mn emission of CsPbCl3 perovskite nanocrystals via incorporation of rubidium ions. Mater. Res. Bull. 2021, 133, 111080. [Google Scholar] [CrossRef]
  29. Hu, M.; Luo, J.; Li, S.; Liu, J.; Li, J.; Tan, Z.; Niu, G.; Wang, Z.; Tang, J. Broadband emission of double perovskite Cs2Na0.4Ag0.6In0.995Bi0.005Cl6:Mn2+ for single-phosphor white-light-emitting diodes. Opt. Lett. 2019, 44, 4757. [Google Scholar] [CrossRef]
  30. Locardi, F.; Cirignano, M.; Baranov, D.; Dang, Z.; Prato, M.; Drago, F.; Ferretti, M.; Pinchetti, V.; Fanciulli, M.; Brovelli, S.; et al. Colloidal synthesis of double perovskite Cs2AgInCl6 and Mn-doped Cs2AgInCl6 nanocrystals. J. Am. Chem. Soc. 2018, 140, 12989–12995. [Google Scholar] [CrossRef]
  31. Nila, N.K.; Angshuman, N. Synthesis and luminescence of Mn-doped Cs2AgInCl6 double perovskites. Chem. Commun. 2018, 54, 5205–5208. [Google Scholar]
  32. Luo, J.; Wang, X.; Li, S.; Liu, J.; Guo, Y.; Niu, G.; Yao, L.; Fu, Y.; Gao, L.; Dong, Q.; et al. Efficient and stable emission of warm-white light from lead-free halide double perovskites. Nature 2018, 563, 541. [Google Scholar] [CrossRef] [PubMed]
  33. Makuła, P.; Pacia, M.; Macyk, W. How to correctly determine the band gap energy of modified semiconductor photocatalysts based on UV–Vis spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [PubMed]
  34. Singh, A.; Chaurasiya, R.; Bheemaraju, A.; Chen, J.-S.; Satapathi, S. Strain-induced band-edge modulation in lead-free antimony-based double perovskite for visible-light absorption. ACS Appl. Energy Mater. 2022, 5, 3926–3932. [Google Scholar] [CrossRef]
  35. Zhang, L.; Fang, Y.; Sui, L.; Yan, J.; Wang, K.; Yuan, K.; Mao, W.L.; Zou, B. Tuning emission and electron–phonon coupling in lead-free halide double perovskite Cs2AgBiCl6 under pressure. ACS Energy Lett. 2019, 4, 2975–2982. [Google Scholar] [CrossRef]
Figure 1. (a) Fluorescence and afterglow photos of the Cs2Na0.2Ag0.8InCl6:y%Mn crystals. (b) Afterglow spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. The samples were excited by 365 nm lamp for 10 min. (c) The afterglow intensity versus y value. (d) Afterglow decay traces of the Cs2Na0.2Ag0.8InCl6:y%Mn. Inset: the afterglow lifetime versus y value.
Figure 1. (a) Fluorescence and afterglow photos of the Cs2Na0.2Ag0.8InCl6:y%Mn crystals. (b) Afterglow spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. The samples were excited by 365 nm lamp for 10 min. (c) The afterglow intensity versus y value. (d) Afterglow decay traces of the Cs2Na0.2Ag0.8InCl6:y%Mn. Inset: the afterglow lifetime versus y value.
Nanomaterials 15 01002 g001
Figure 2. (a) Steady-state PL spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (b) PL spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn after normalizing the red band. Inset: Gaussian fitting of a selected PL spectrum of Cs2Na0.2Ag0.8InCl6:40% Mn. (c) Relative PL intensity and peak position of the red band versus Mn2+ doping concentration. (df) Time resolved PL spectra monitored at 466 nm (d), 518 nm (e), 590 nm (f). (gi) The relationships between PL lifetimes and y value.
Figure 2. (a) Steady-state PL spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (b) PL spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn after normalizing the red band. Inset: Gaussian fitting of a selected PL spectrum of Cs2Na0.2Ag0.8InCl6:40% Mn. (c) Relative PL intensity and peak position of the red band versus Mn2+ doping concentration. (df) Time resolved PL spectra monitored at 466 nm (d), 518 nm (e), 590 nm (f). (gi) The relationships between PL lifetimes and y value.
Nanomaterials 15 01002 g002
Figure 3. (a) The XRD patterns of the Cs2Na0.2Ag0.8InCl6:y%Mn and Cs2Na0.2Ag0.8InCl6. (b) The enlarged diffraction peak at 23.9°. The vertical dash line is used to highlight the peak shift. (c) The relationships between lattices spacing and y value according to HRTEM results. (d) HRTEM images of the Cs2Na0.2Ag0.8InCl6:y%Mn, I–V: the Mn2+ doping concentration in ascending order.
Figure 3. (a) The XRD patterns of the Cs2Na0.2Ag0.8InCl6:y%Mn and Cs2Na0.2Ag0.8InCl6. (b) The enlarged diffraction peak at 23.9°. The vertical dash line is used to highlight the peak shift. (c) The relationships between lattices spacing and y value according to HRTEM results. (d) HRTEM images of the Cs2Na0.2Ag0.8InCl6:y%Mn, I–V: the Mn2+ doping concentration in ascending order.
Nanomaterials 15 01002 g003
Figure 4. (a) UV–vis absorption spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (b) Tauc plots of (αhν)2 versus hν. (c) VB XPS spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (d) Scheme illustrating the band structures of the Cs2Na0.2Ag0.8InCl6:y%Mn.
Figure 4. (a) UV–vis absorption spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (b) Tauc plots of (αhν)2 versus hν. (c) VB XPS spectra of the Cs2Na0.2Ag0.8InCl6:y%Mn. (d) Scheme illustrating the band structures of the Cs2Na0.2Ag0.8InCl6:y%Mn.
Nanomaterials 15 01002 g004
Figure 5. Applications of the Cs2Na0.2Ag0.8InCl6, Cs2Na0.2Ag0.8InCl6:40%Mn, and Cs2Na0.2Ag0.8InCl6:50%Mn in optical information storage and encryption.
Figure 5. Applications of the Cs2Na0.2Ag0.8InCl6, Cs2Na0.2Ag0.8InCl6:40%Mn, and Cs2Na0.2Ag0.8InCl6:50%Mn in optical information storage and encryption.
Nanomaterials 15 01002 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

Hu, Y.-L.; Zhu, Y.-L.; Gu, S.-Y.; Xu, J.-Q.; Gan, Z.-X.; Shi, C.-G. Modulating the Afterglow Time of Mn2+ Doped Metal Halides and Applications in Advanced Optical Information Encryption. Nanomaterials 2025, 15, 1002. https://doi.org/10.3390/nano15131002

AMA Style

Hu Y-L, Zhu Y-L, Gu S-Y, Xu J-Q, Gan Z-X, Shi C-G. Modulating the Afterglow Time of Mn2+ Doped Metal Halides and Applications in Advanced Optical Information Encryption. Nanomaterials. 2025; 15(13):1002. https://doi.org/10.3390/nano15131002

Chicago/Turabian Style

Hu, Yu-Lin, Yi-Lin Zhu, Shi-Ying Gu, Jia-Qing Xu, Zhi-Xing Gan, and Chuan-Guo Shi. 2025. "Modulating the Afterglow Time of Mn2+ Doped Metal Halides and Applications in Advanced Optical Information Encryption" Nanomaterials 15, no. 13: 1002. https://doi.org/10.3390/nano15131002

APA Style

Hu, Y.-L., Zhu, Y.-L., Gu, S.-Y., Xu, J.-Q., Gan, Z.-X., & Shi, C.-G. (2025). Modulating the Afterglow Time of Mn2+ Doped Metal Halides and Applications in Advanced Optical Information Encryption. Nanomaterials, 15(13), 1002. https://doi.org/10.3390/nano15131002

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