DOI:
10.1039/D5QI00184F
(Research Article)
Inorg. Chem. Front., 2025,
12, 3118-3125
Long-lasting far-UVC persistent luminescence for solar-blind optical tagging†
Received
18th January 2025
, Accepted 6th March 2025
First published on 7th March 2025
Abstract
Far ultraviolet-C (far-UVC; 200–230 nm) luminescent materials have garnered significant interest in recent years, driven by the growing demands for applications such as disinfection and solar-blind imaging due to their distinct wavelength features. However, the research and development of far-UVC persistent phosphors are lacking. Here, we report the realization of far-UVC persistent luminescence in CaSO4:Pr3+ and CaSO4:Pb2+ phosphors, which show emissions peaking at 220 nm and 230 nm with a long persistence time of >24 h after ceasing X-ray excitation. This is by far the shortest UVC afterglow emission to the best of our knowledge. The far-UVC afterglow from the charged phosphors can be readily detected using a solar-blind UV camera in both indoor-lighting and outdoor environments owing to the absence of background noise from ambient light. The continuous photostimulation of indoor white LED light and outdoor sunlight has different impacts on the far-UVC afterglow performance of CaSO4:Pr3+ and CaSO4:Pb2+ phosphors, which is elucidated by the decay time-dependent thermoluminescence (TL) curves under different light conditions. This study expands the field of persistent luminescence to the far-UVC spectral region and will inspire the discovery of more excellent far-UVC persistent phosphors.
Introduction
Ultraviolet-C (UVC) light with a wavelength range from 200 to 280 nm has been instrumental in a variety of important applications such as disinfection, phototherapy, confidential communication, optical information storage, and solar-blind optical tagging.1–6 For example, UVC light can effectively eradicate bacteria, viruses, and other pathogens by disrupting their nucleic acids, thereby halting their reproduction.7,8 In the realm of solar-blind optical tagging, the lack of UVC light on the Earth's surface results in zero background interference, enabling the UVC signal to be monitored and imaged with high contrast.9 Currently, the market for UVC radiation sources is dominated by mercury discharge lamps with an emission wavelength of 254 nm. However, these gaseous light sources have the drawbacks of being bulky, fragile, and inefficient.10,11 Furthermore, these lamps are designed for operation exclusively in areas inaccessible to humans, as their emission spectra pose a risk to human eyes and skin.12,13 The recently developed solid-state deep-UV LEDs can replace gaseous lamps in some ways, but there are still challenging issues with these AlGaN-based UV LEDs, such as expensive production costs, complex preparation procedures, and low external quantum efficiency.14,15 On the other hand, extensive research efforts have focused on lanthanide-based inorganic luminescent materials for UVC luminescence through upconversion or downshifting processes, opening up new avenues for UVC technology.16–19 Among them, UVC persistent phosphors, characterized by their unique delayed luminescence properties, have drawn an unprecedented amount of research and technological attention.20
For UVC persistent luminescence, Pr3+, Bi3+ and Pb2+ stand out as promising luminescence centers in solids because of their efficient 4f5d → 4f inter-configurational transition or 3P1 → 1S0 transition.21–26 Recently, some Pr3+-doped and Bi3+-doped UVC persistent phosphors have been reported.27–31 For example, Yang et al. reported a Cs2NaYF6:Pr3+ UVC persistent phosphor with an emission peak at 250 nm and an afterglow time of ∼2 h after X-ray irradiation.32 Wang and coworkers reported a series of Pr3+-doped silicate-based UVC persistent phosphors, which can be effectively charged using a standard 254 nm mercury lamp and exhibit UVC persistent luminescence peaking at 265–270 nm.9 Lv et al. reported an X-ray-activated YBO3:Pr3+ UVC persistent phosphor with an emission peak at 261 nm, which exhibits a remarkable enhancement of UVC luminescence intensity in bright environments.33 However, the majority of existing UVC persistent phosphors show dominant emission bands above 250 nm,8,34–36 while persistent phosphors that emit in the far-UVC spectral range (200–230 nm) are lacking. Notably, the far-UVC persistent phosphors can potentially revolutionize the way that UVC light is used; for instance, these far-UVC materials can act as self-sustained glowing sources for air, surface, and water disinfection application in occupied public locations because of their stronger absorbance than regular 254 nm UVC radiation and limited penetration into human tissues.37–40
In this study, we report two new far-UVC persistent phosphors, CaSO4:Pr3+ and CaSO4:Pb2+, which can emit far-UVC persistent luminescence with a peak wavelength at 220 nm and 230 nm for >24 h after the cessation of X-ray excitation. To our knowledge, this represents the shortest UVC afterglow emission observed thus far. Because of zero UVC background on the surface of the Earth, the UVC emission from the charged persistent phosphors can be clearly detected using a solar-blind UV camera for more than 1 h in an indoor lighting or outdoor sunlight environment. This work enriches the persistent phosphors emitting in the far-UVC spectral range, paving the way for their versatile applications such as solar-blind optical tagging, disinfection, and photodynamic therapy.
Experimental
Synthesis
CaSO4:x%Pr3+ (x = 0.1, 0.2, 0.5, and 1) phosphors were prepared by the high-temperature solid-state reaction method. Stoichiometric amounts of CaSO4·2H2O (99%, Aladdin), Pr6O11 (99.99%, Aladdin), and (NH4)2SO4 (99.99%, Macklin) powders were mixed and ground homogeneously in an agate mortar. Then the obtained powders were pre-sintered at 200 °C in air for 2 h. The pre-fired powders were ground again and pressed into 11 mm diameter discs using a 30 T hydraulic press. Finally, the discs were sintered at 1100 °C in air for 4 h to form the final CaSO4:Pr3+ solid ceramic discs. The processes for determining the sintering conditions of the CaSO4:Pr3+ phosphors are shown in Fig. S1–S4 in the ESI.† The CaSO4:Pb2+ phosphors in this work were fabricated following the same procedure with the only variation being the use of PbO (99.999%, Aladdin) as the dopant source powder.
Characterization
The crystal structure and phase composition were checked using a PANalytical X'Pert PRO powder X-ray diffractometer with Cu Kα1 radiation (λ = 1.5406 Å). Structure refinement was processed using the general structure analysis system (GSAS) program. The SEM image and elemental maps were recorded using a JSM-7800F field-emission SEM. The spectral properties, including radioluminescence and persistent luminescence properties, were measured using an FLS1000 spectrofluorometer (Edinburgh Instruments) equipped with a photomultiplier tube detector (PMT, 200–900 nm). Thermoluminescence spectra were recorded using an SL18 thermoluminescence setup (Guangzhou Rongfan Science and Technology Co., Ltd; heating rate, 4 °C per s). Before all the measurements, the samples were thermally cleaned at 450 °C to empty the traps thoroughly, followed by irradiation with an X-ray tube (Moxtek, tungsten target). UVC afterglow images from the persistent phosphors were taken using a solar-blind ultraviolet camera (WNZW-01, Suzhou Micro-Nano Laser & Photonic Technology Co., Ltd) with dual UVC (240–280 nm)-visible channels, which superimpose the UVC images onto the visible images and show the invisible UVC signal as an area of red color. UVC persistent luminescence power intensities were obtained with a Newport 1936-R optical power meter and a Newport 918D-UV-OD3R UV-enhanced silicon photodetector.
First-principles calculations
The initial atomic positions and symmetry information of the host crystal were taken from the Inorganic Crystal Structure Database and the primitive cell of CaSO4 containing 96 atoms (16 Ca atoms, 16 S atoms, and 64 O atoms) was used in the simulation. Using the Vienna Ab initio simulation package (VASP),41,42 theoretical simulations were conducted based on density functional theory (DFT), the generalized gradient approximation (GGA)-Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional was used for the description of the exchange and correlation energy of the electrons and the PBE+U method was utilized to calculate Pr-4f electrons.43 Ca (4s2), S (3s23p4), O (2s22p4), and Pr (4f36s2) were treated as valence electrons, and their interactions with the respective cores were described by the projected augmented wave (PAW) method. The equilibrium structures were acquired by optimizing the atomic positions until the energy change was less than 10–6 eV and the Hellmann−Feynman forces on atoms were less than 0.01 eV Å−1. The plane-wave cut-off energy was set at 550 eV. For k-point integration within the first Brillouin zone, a 3 × 3 × 3 Monkhorst–Pack grid was chosen for self-consistent calculation, while the high symmetry k-paths were generated for band structure calculation.
Results and discussion
Crystal structure and phase identification
The CaSO4 crystal belongs to the Bmmb (63) space group with an orthorhombic structure, which has the lattice parameters of a = 6.992 Å, b = 6.999 Å, c = 6.240 Å, α = β = γ = 90°, and V = 305.37 Å3. The simulated crystal structure of CaSO4 is shown in Fig. 1a. In the CaSO4 lattice, the Ca2+ cations are bound by eight oxygen ions, whereas the S6+ ion is encircled by a tetrahedron of four oxygen ions. Under consideration of a similar ionic radius and valence state, the Pr3+ ion (1.126 Å, CN = 8) and Pb2+ ion (1.29 Å, CN = 8) most probably occupy the Ca2+ site (1.12 Å, CN = 8) rather than the S6+ site (0.12 Å, CN = 4) in the CaSO4 lattice. Fig. 1b presents the XRD patterns of the as-synthesized CaSO4:x%Pr3+ (x = 0.1, 0.2, 0.5, and 1) phosphors. All the diffraction peaks are consistent with the standard XRD pattern of the CaSO4 crystal (JCPDS No. 70-0909), suggesting that introducing Pr3+ will not affect the crystal structure of the CaSO4 compound. The calculated and experimental results as well as the differences in the Rietveld refinement of the CaSO4:0.2%Pr3+ phosphor are shown in Fig. 1d. The detailed refined results are collected in Table S1.† The orthorhombic CaSO4 crystal system was used as the initial model for refinement. The detected and calculated XRD patterns match well with each other and the reliability parameters of the refinement are Rwp = 9.75%, Rp = 6.72%, and χ2 = 2.65, which further manifests that the as-synthesized CaSO4:Pr3+ phosphor is a pure CaSO4 phase. XRD patterns of the as-obtained CaSO4:x%Pb2+ (x = 0.5–5) phosphors are displayed in Fig. 1c, and the diffraction peaks of all samples can also be well indexed to the standard card of CaSO4. The results demonstrate that the pure phase is formed independent of the Pb2+ concentration. The Rietveld refinement results (Fig. 1e) also suggest that the CaSO4:Pb2+ phosphors were synthesized with good crystallinity in a pure phase through satisfactory refinement values of Rwp, Rp, and χ2 (Table S1†). In addition, to study the microstructure of the prepared CaSO4:Pr3+ phosphor, SEM and EDS mapping images were also recorded in Fig. 1f. The phosphor particles show irregular shapes and the sizes of these particles are in the micrometer range. The EDS elemental mapping results demonstrate that Ca, S, O, and Pr elements are homogeneously distributed throughout a randomly selected phosphor particle, further confirming that the CaSO4:Pr3+ phosphor has been successfully prepared.
 |
| Fig. 1 (a) Crystal structure diagram of CaSO4 and the coordination environment of the Ca and S atoms. XRD patterns of (b) CaSO4:x%Pr3+ (x = 0.1–1) phosphors and (c) CaSO4:x%Pb2+ (x = 0.5–5) phosphors. Rietveld refinement of (d) CaSO4:0.2%Pr3+ and (e) CaSO4:3%Pb2+ phosphors. (f) SEM and EDS mapping images of the CaSO4:0.2%Pr3+ phosphor. The inset shows the EDS spectrum of the CaSO4:0.2%Pr3+ phosphor. | |
UVC photoluminescence properties
Taking into account the fact that the luminescence properties of Pr3+ are sensitive to the host material, the band structure and density of states (DOS) of the CaSO4 host and Pr3+-doped CaSO4 were calculated using density functional theory (DFT). As shown in Fig. S5a,† the valence band (VB) maximum and the conduction band (CB) minimum are located at the same point, which suggests that the CaSO4 host has a direct bandgap (Eg). Based on the calculation, the bandgap of CaSO4 is 5.96 eV, which allows for the possibility of deep UV and even far-UVC luminescence. Besides, the total and partial DOSs of the CaSO4 host in Fig. S5b† show that the d orbital of Ca atoms mainly forms the DOS of the CB, while the VB is mostly derived from the p orbital of O atoms. Fig. 2b presents the total and partial DOSs of Pr3+-doped CaSO4. After doping Pr3+ into the CaSO4 host, some new energy levels overlapping with the CB are generated, primarily resulting from the d orbital of the Pr atoms. This is consistent with the changes in the band structure of Pr3+-doped CaSO4 in Fig. 2a compared with that of the CaSO4 host.
 |
| Fig. 2 (a) Band structure and (b) total DOS and partial DOS of Pr-doped CaSO4 (Pr replaces Ca). (c) Energy level scheme of the Pr3+ 4f2 → 4f15d1 configuration and the possible optical transitions. The radioluminescence spectra of (d) CaSO4:0.2%Pr3+ and (e) CaSO4:3%Pb2+ phosphors. | |
The UVC emission of the Pr3+-activated phosphors is ascribed to parity allowed 4f5d → 4f2 inter-configurational transitions. Two general requirements must be met for Pr3+ 4f5d → 4f2 transitions to occur in a solid: an appropriate placement for the 4f5d level, meaning that it is energetically below the 1S0 level and suitably separated from the closest 3PJ level, and a very little Stokes shift (less than 3000 cm−1),44 as depicted in Fig. 2c. The radioluminescence spectrum of the CaSO4:0.2%Pr3+ phosphor upon high-energy X-ray excitation is presented in Fig. 2d. It includes the UV emission dominated by four distinct UVC emission peaks at 220 nm, 231 nm, 247 nm, and 255 nm, which are attributed to the inter-configurational transitions from the lowest 4f5d state to different 4f2 states (3H4, 3H5, 3H6, and 3F2) of Pr3+, and the visible and infrared emission bands originating from the 4f → 4f intra-configurational transitions. The effective inter-configurational and intra-configurational transitions corresponding to different emission peaks were also labeled in Fig. 2d. Fig. 2e shows the radioluminescence spectrum of the CaSO4:3%Pb2+ phosphor upon X-ray excitation. The inset is the energy level structure of the Pb2+ ion. When excited with X-rays, the CaSO4:Pb2+ phosphor exhibits an ultra-narrow emission band in the UVC spectral range with a maximum at 230 nm and a relatively weak broad UVA emission band ranging from 300 to 400 nm, which can be ascribed to the 3P0,1 → 1S0 transition of Pb2+ and the radiative emission of Pb2+ pairs and/or clusters formed in the CaSO4 host lattice, respectively.
Far-UVC persistent luminescence properties
In addition to UVC radioluminescence, irradiation with high-energy X-rays can also induce long-lasting UVC persistent luminescence in CaSO4:Pr3+ and CaSO4:Pb2+ phosphors. As shown in Fig. S6 and S7,† the UVC persistent luminescence decay curves, the corresponding persistent luminescence emission spectra, and the TL curves of the CaSO4:Pr3+ phosphors with different Pr3+ doping concentrations were recorded in detail after irradiation with an X-ray beam, which suggests that the optimal doping concentration of Pr3+ is 0.2%. Moreover, to investigate the correlation between the irradiation time and UVC persistent luminescence performance, the persistent luminescence decay curves and the corresponding persistent luminescence emission spectra of the CaSO4:0.2%Pr3+ phosphor were analyzed by varying the excitation durations with the same X-ray irradiation dose (208.20 mGy s−1), as illustrated in Fig. S8.† The results indicate that the CaSO4:0.2%Pr3+ phosphor can achieve full charging with 15 min of X-ray excitation. Additionally, the TL curves of the CaSO4:0.2%Pr3+ phosphor with varying irradiation durations also show that 15 min of 208.20 mGy s−1 X-ray exposure is sufficient for complete charging (Fig. S9†). Likewise, by varying Pb2+ doping concentrations, the UVC afterglow performance and TL spectra of the CaSO4:Pb2+ phosphors were also examined (Fig. S10 and S11†), showing that 3% is the ideal Pb2+ doping concentration. Furthermore, we also recorded the afterglow decay curves, afterglow emission spectra, and TL curves of the CaSO4:3%Pb2+ phosphor with different irradiation times (Fig. S12 and S13†), which show that 15 min of 290.30 mGy s−1 X-ray excitation can fully charge this phosphor.
Fig. 3a depicts the far-UVC persistent luminescence decay curve of the CaSO4:0.2%Pr3+ phosphor monitored at 220 nm in the dark at room temperature after X-ray irradiation for 15 min. The data were collected by tracking the afterglow intensity at 220 nm over time for a duration of 24 h. The far-UVC afterglow intensity experiences a rapid decrease within the first few hours (0–6 h), followed by a gradual decline until the end of the observation period. Even after 24 h of natural decay in darkness, the far-UVC afterglow intensity remains significantly higher than the background, indicating that the far-UVC afterglow of the CaSO4:0.2%Pr3+ phosphor can persist for over 24 h. The persistent luminescence emission spectrum of the CaSO4:0.2%Pr3+ phosphor in the upper inset of Fig. 3a shares an identical emission profile with the radioluminescence spectrum in the UVC region, suggesting that the persistent UVC emission originates from the 4f5d → 4f2 transition of Pr3+. Fig. 3b presents the afterglow emission spectra of the CaSO4:0.2%Pr3+ phosphor acquired at different decay times within 6 h after ceasing X-ray excitation. Upon prolonging the decay time, the relative emission intensity of persistent luminescence decreases, while the spectral profiles including the peak position and spectral shape remain constant, which further demonstrates that the UVC afterglow is attributed to the Pr3+ emitter. As given in Fig. S14a and Table S2,† the far-UVC persistent luminescence power density of the CaSO4:0.2%Pr3+ phosphor at 30 s after the stoppage of X-ray irradiation was determined to be ∼12.43 mW m−2, and the recording lasts for 30 min. In addition, we also measured the far-UVC afterglow decay curve monitored at 230 nm and persistent luminescence emission spectra at different decay times of the CaSO4:3%Pb2+ phosphor after irradiation with an X-ray tube, as shown in Fig. 3c and d. The results suggest that the far-UVC afterglow of the CaSO4:3%Pb2+ phosphor can also last for >24 h and does stem from the Pb2+ emitters. Moreover, the afterglow power density of the CaSO4:3%Pb2+ phosphor at a decay time of 30 s was found to be ∼23.60 mW m−2 using an optical power meter (Fig. S14b and Table S2†), which is about twice as high as that of the CaSO4:0.2%Pr3+ phosphor. Additionally, the afterglow power intensities of these two phosphors are comparable with other reported UVC persistent phosphors, as shown in Table S3.†
 |
| Fig. 3 (a) Far-UVC persistent luminescence decay curve of the CaSO4:0.2%Pr3+ phosphor monitored at 220 nm in the dark at room temperature after X-ray irradiation. The upper inset shows the persistent luminescence emission spectrum acquired at 10 min decay after the stoppage of the irradiation. (b) Persistent luminescence emission spectra of the pre-irradiated CaSO4:0.2%Pr3+ phosphor at different decay times. (c) Far-UVC afterglow decay curve of the CaSO4:3%Pb2+ phosphor monitored at 230 nm in the dark at room temperature after X-ray irradiation. The upper inset displays the afterglow emission spectrum measured at 10 min decay after ceasing X-ray irradiation. (d) Afterglow emission spectra of the pre-irradiated CaSO4:3%Pb2+ phosphor at different decay times. (e) UVC persistent luminescence images of the charged CaSO4:0.2%Pr3+ phosphor disc at different decay times in the dark, in an indoor LED lighting environment (300 lux) and in an outdoor sunlight environment (∼20 000 lux), respectively. The phosphor disc was pre-irradiated with X-rays for 15 min. | |
By taking advantage of the zero solar UVC background on the Earth's surface, we verify the unique capability of the CaSO4:Pr3+ and CaSO4:Pb2+ UVC persistent phosphors as solar-blind tags for high-contrast in bright indoor and outdoor environments using a solar-blind UV camera. Fig. 3e displays the UVC persistent luminescence images of the charged CaSO4:0.2%Pr3+ phosphor disc at various decay times under different ambient light conditions. Each image combines a UVC emission depiction with a visible one. Note that the detected invisible UVC emission is shown by the red pattern, whose area is proportional to the UVC emission intensity. It is found that the UVC persistent luminescence signal from the CaSO4:0.2%Pr3+ phosphor can be detected and imaged for more than 1 h in the dark, in an indoor LED lighting environment (300 lux) or in an outdoor sunlight environment (∼20
000 lux), and both indoor ambient light and outdoor sunlight have no significant effect on the UVC afterglow intensity and detectable imaging time. However, the UVC persistent luminescence performance of the CaSO4:3%Pb2+ phosphor under the photostimulation of ambient light is very different. As shown in Fig. S15,† the UVC afterglow photographs of the CaSO4:3%Pb2+ phosphor in darkness and in bright indoor and outdoor environments were also captured after the cessation of X-ray excitation. Compared to that in the dark, the UVC afterglow intensity obviously increases within 1 h in bright indoor ambient light, but it exhibits a more pronounced enhancement in the initial stage and decays at a much faster rate in outdoor sunlight, which can be attributed to the fact that natural sunlight with such a high illuminance value (∼10
000 lux) can dramatically accelerate the release of the stored energy in the CaSO4:3%Pb2+ phosphor.
Thermoluminescence properties
To investigate the de-trapping processes of CaSO4:Pr3+ and CaSO4:Pb2+ phosphors in the dark and bright environments and figure out the reason why the far-UVC afterglow performance of these two phosphors is affected differently by the photostimulation of ambient light, detailed TL measurements were carried out. Fig. 4a depicts the time-dependent TL curves of the decaying CaSO4:0.2%Pr3+ phosphor in darkness, which reflect the room-temperature thermal emptying of the stored energy in energy traps. For the case of 30 s short decay in the dark, the TL curve covers almost the entire test temperature range from 300 to 600 K, with a dominant peak at 377 K (trap I), a shoulder at 447 K (trap II), and a weak glow peak at 572 K (trap III). With the increase of decay time from 30 s to 24 h, the integral TL intensity gradually diminishes, accompanied by a shift of the TL peaks for traps I, II, and III towards higher temperatures. The TL intensity of trap I drops most quickly, which is followed by a slow decline of TL intensities of trap II and trap III, indicating a close relationship between room-temperature thermal stimulation and trap depth in the CaSO4:Pr3+ phosphor. The TL curves of a decaying CaSO4:0.2%Pr3+ phosphor under the illumination of 300 lux white LED light are shown in Fig. 4b. Compared with the TL spectra in Fig. 4a, there is no obvious difference in the variation of the TL intensity of trap I, while the TL intensities of trap II and trap III exhibit a faster and more pronounced decrease trend with prolonged exposure to white light. Notably, trap I remains dominant among all traps in the phosphor, and the release of energy stored in trap I is almost unaffected by indoor LED lighting illumination. This perverse finding aligns with the results in Fig. 3e, which suggests that the continuous photo-stimulation from ambient light has minimal impact on the UVC persistent luminescence of the CaSO4:0.2%Pr3+ phosphor. It makes sense that ambient light illumination can substantially accelerate the release of the stored energy in deeper traps (trap II and trap III). However, the reason why it does not notably affect the release of the stored energy in shallow traps (trap I) is vague and worth investigating.
 |
| Fig. 4 TL curves of the CaSO4:0.2%Pr3+ phosphor recorded at different decay times in the (a) dark and (b) indoor white LED lighting environments (300 lux) after irradiation with X-rays. TL curves of the CaSO4:3%Pb2+ phosphor measured at different decay times in the (c) dark and (d) indoor white LED lighting environments (300 lux) after X-ray irradiation. | |
Based on the time-dependent TL curves of the CaSO4:0.2%Pr3+ phosphor in dark and bright (300 lux) indoor environments, we estimated the depths of the shallowest occupied trap corresponding to various decay times using an initial rise analysis method, as depicted in Fig. S16a–c.† As the decay time is prolonged from 30 s to 24 h, the obtained trap depth in darkness increases continuously from 0.76 to 1.07 eV, which is smaller than that calculated using the improved peak position method.45 However, the calculated trap depth in the bright environment is lower than that in darkness at the same decay time, and first increases from 0.74 to 0.98 eV and then declines to 0.91 eV with the extension of the illumination time of white light, which is because ambient light stimulation can trigger trap redistribution within the phosphor. Specifically, partially stored electrons can be promoted from these deep traps to shallow traps directly or deep traps can release trapped electrons into the conduction band when exposed to polychromic white LED light. In the latter case, some of these released electrons can be re-captured again by the depleted shallow traps. According to the TL results, the electron transfer from deep traps to shallow traps is almost in dynamic equilibrium with the process of releasing electrons stored in these shallow traps for UVC luminescence upon continuous photostimulation of ambient light; thus it seems that photo-stimulation has no obvious effect on the UVC persistent luminescence performance of the CaSO4:Pr3+ phosphor.
To find out why the influence of ambient light stimulation on the UVC afterglow performance of the CaSO4:Pb2+ phosphor is different from that of the CaSO4:Pr3+ phosphor, the TL spectra of the charged CaSO4:3%Pb2+ phosphor at different decay times were also recorded in dark and bright (300 lux) indoor environments, as given in Fig. 4c and d. Comparing the TL curve of the CaSO4:Pb2+ phosphor after 30 s short decay in the dark with that of the CaSO4:Pr3+ phosphor, it can be observed that the TL peak for trap I moves slightly towards a higher temperature and its distribution is narrower. The peak position of the TL peak corresponding to trap II hardly changes, but the TL peak for trap III disappears when Pb2+ replaces Pr3+ as the dopant ion, indicating that trap I and trap II should stem from the intrinsic defects of the CaSO4 host, and trap III is believed to result from the Pr-related extrinsic defects due to the charge imbalance between Pr3+ and Ca2+. Fig. 4d shows the time-dependent TL spectra of the pre-irradiated CaSO4:Pb2+ phosphor under the illumination of 300 lux white LED light. Compared to that of the CaSO4:Pb2+ phosphor in the dark, the TL intensities of trap I and trap II drop much faster as the illumination time of white light is extended, which is different from the response of the CaSO4:Pr3+ phosphor to ambient light stimulation. The trap depths of the CaSO4:Pb2+ phosphor at various decay instants were also calculated using the initial rising method, as displayed in Fig. S16d–f.† As shown in Fig. S16f,† the process of re-capturing electrons released from the deep traps in the depleted shallow traps is also evident upon ambient light illumination. However, for these shallow traps in the CaSO4:Pb2+ phosphor, the predominant mechanism involves the release of stored electrons in them for UVC luminescence, outweighing the electron transfer from deep traps to shallow traps induced by photostimulation. This dominance is attributed to the absence of deep traps (trap III) in this phosphor, according to the results in Fig. 4c and d.
Solar-blind optical tagging application
Considering the long-lasting UVC persistent luminescence of CaSO4:Pr3+ and CaSO4:Pb2+ phosphors and the absence of UVC light on the Earth's surface, the CaSO4:Pr3+ and CaSO4:Pb2+ phosphors can function as solar-blind optical taggants for identifying and tracking objects in bright environments. We illustrated their covert optical tagging application in bright environments by using the CaSO4:Pr3+ phosphor. Taking advantage of the invisible UVC luminescence signal from the charged CaSO4:Pr3+ phosphor, the information encryption application using the binary code was first designed and realized, as shown in Fig. 5a. The charged CaSO4:Pr3+ phosphor discs and other non-glowing ceramic discs (same size but without UVC afterglow) were arranged in a specific order of three rows and eight columns, and represent the binary codes “1” and “0”, respectively. We can only see three rows of dots with the naked eye or a standard digital camera. With the help of a solar-blind UV camera, the two kinds of ceramic discs were distinguished, and the encrypted information “UVC” in the binary code was decoded based on the ASCII table. As displayed in Fig. 5b, by virtue of the strong and long-lasting UVC afterglow of the CaSO4:Pr3+ phosphor, the dynamic tracking application has also been demonstrated in a bright outdoor environment. A pre-irradiated CaSO4:Pr3+ phosphor disc was affixed to a cartoon car in the outdoor square. The movement of the cartoon car can be tracked by utilizing a solar-blind UV camera to capture the UVC luminescence signal. The self-sustaining UVC luminescence of CaSO4:Pr3+ and CaSO4:Pb2+ phosphors presents exciting prospects for information encryption and long-lasting optical tagging applications in bright environments.
 |
| Fig. 5 (a) Visible and UVC images of letters “U”, “V”, and “C” encrypted using the binary code. (b) Schematic diagram of dynamic tracking application of a pre-irradiated CaSO4:Pr3+ phosphor disc affixed to a cartoon car in the outdoor square. A solar-blind UV camera was used to capture the location of the cartoon car in a two-dimensional plane, with a potted plant as a reference. | |
Conclusions
In summary, we report two new far-UVC persistent phosphors, CaSO4:Pr3+ and CaSO4:Pb2+, capable of emitting intense far-UVC afterglow at 220 nm and 230 nm for >24 h after ceasing X-ray excitation. To our knowledge, this represents the shortest UVC persistent luminescence emission observed thus far. Benefitting from the strong persistent UVC light emission and zero-background noise from ambient light, the afterglow signal from the charged phosphors can be clearly monitored and imaged using a solar-blind UV camera in bright indoor and outdoor environments. Moreover, we found that both indoor ambient light and outdoor sunlight exert a negligible influence on the UVC afterglow intensity and detectable imaging duration of the CaSO4:Pr3+ phosphor, while the UVC afterglow performance of the CaSO4:Pb2+ phosphor has a rather different profile. This phenomenon is elucidated by the decay time-dependent TL spectra in darkness and under indoor white LED light exposure, and the different trap distribution caused by the substitution of Pr3+ and Pb2+ for Ca2+ in the CaSO4 host is responsible for the distinct UVC afterglow performance of the CaSO4:Pr3+ and CaSO4:Pb2+ phosphors. The far-UVC persistent phosphors discovered here are expected to bring new solutions to some important applications where far-UVC light is needed, such as sterilization, water purification, solar-blind optical tagging, and photodynamic therapy.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (Grant Nos. 12474407 and 51902184), Key Research and Development Program of Shandong Province (Major Scientific and Technological Innovation Project) (Grant No. 2021CXGC011101), the State Key Laboratory of Infrared Physics, Shanghai Institute of Technical Physics, Chinese Academy of Sciences (Grant No. SITP-NLIST-YB-2022-10), and the “Qi-Lu Young Scholar Fund” of Shandong University.
References
- X. Fan, R. Huang and H. Chen, Application of ultraviolet C technology for surface decontamination of fresh produce, Trends Food Sci. Technol., 2017, 70, 9–19 CrossRef CAS
.
- G. Matafonova and V. Batoev, Recent advances in application of UV light-emitting diodes for degrading organic pollutants in water through advanced oxidation processes: A review, Water Res., 2018, 132, 177–189 CrossRef CAS PubMed
.
- E. L. Cates, M. Cho and J.-H. Kim, Converting Visible Light into UVC: Microbial Inactivation by Pr3+-Activated Upconversion Materials, Environ. Sci. Technol., 2011, 45, 3680–3686 CrossRef CAS PubMed
.
- S. Nayak, S.-E. O'Donnell, C. M. Sales and R. V. Tikekar, Fructose Accelerates UV-C Induced Photochemical Degradation of Pentachlorophenol in Low and High Salinity Water, J. Agric. Food Chem., 2016, 64, 4214–4219 CrossRef CAS PubMed
.
- X. He, E. Xie, M. S. Islim, A. A. Purwita, J. J. D. McKendry, E. Gu, H. Haas and M. D. Dawson, 1 Gbps free-space deep-ultraviolet communications based on III-nitride micro-LEDs emitting at 262 nm, Photon. Res., 2019, 7, B41–B47 CrossRef CAS
.
- X. Chen, F. Ren, S. Gu and J. Ye, Review of gallium-oxide-based solar-blind ultraviolet photodetectors, Photon. Res., 2019, 7, 381–415 CrossRef CAS
.
- M. Raeiszadeh and B. Adeli, A Critical Review on Ultraviolet Disinfection Systems against COVID-19 Outbreak: Applicability, Validation, and Safety Considerations, ACS Photonics, 2020, 7, 2941–2951 CrossRef CAS PubMed
.
- X. Zhou, J. Qiao, Y. Zhao, K. Han and Z. Xia, Multi-responsive deep-ultraviolet emission in praseodymium-doped phosphors for microbial sterilization, Sci. China Mater., 2022, 65, 1103–1111 CrossRef CAS PubMed
.
- X. Wang, Y. Chen, F. Liu and Z. Pan, Solar-blind ultraviolet-C persistent luminescence phosphors, Nat. Commun., 2020, 11, 2040 CrossRef CAS PubMed
.
- M. M. Afandi, H. Kang, T. Kang, J. Park and J. Kim, AC-Driven Ultraviolet-C Electroluminescence from an All-Solution-Processed CaSiO3:Pr3+ Thin Film Based on a Metal-Oxide-Semiconductor Structure, Adv. Mater. Interfaces, 2022, 9, 2200248 CrossRef CAS
.
- S.-Y. Song, K.-K. Liu, J.-Y. Wei, Q. Lou, Y. Shang and C.-X. Shan, Deep-Ultraviolet Emissive Carbon Nanodots, Nano Lett., 2019, 19, 5553–5561 CrossRef CAS PubMed
.
- A. Trevisan, S. Piovesan, A. Leonardi, M. Bertocco, P. Nicolosi, M. G. Pelizzo and A. Angelini, Unusual High Exposure to Ultraviolet-C Radiation, Photochem. Photobiol., 2006, 82, 1077–1079 CrossRef CAS PubMed
.
- S. Zaffina, V. Camisa, M. Lembo, M. R. Vinci, M. G. Tucci, M. Borra, A. Napolitano and V. Cannatà, Accidental Exposure to UV Radiation Produced by Germicidal Lamp: Case Report and Risk Assessment, Photochem. Photobiol., 2012, 88, 1001–1004 CrossRef CAS PubMed
.
- M. Kneissl, T.-Y. Seong, J. Han and H. Amano, The emergence and prospects of deep-ultraviolet light-emitting diode technologies, Nat. Photonics, 2019, 13, 233–244 CrossRef CAS
.
- S. Liang and W. Sun, Recent Advances in Packaging Technologies of AlGaN-Based Deep Ultraviolet Light-Emitting Diodes, Adv. Mater. Technol., 2022, 7, 2101502 CrossRef CAS
.
- V. A. Pustovarov, K. V. Ivanovskikh, Y. E. Khatchenko, Q. Shi and M. Bettinelli, Energy conversion in LiSrPO4 doped with Pr3+ ions, Radiat. Meas., 2019, 123, 39–43 CrossRef CAS
.
- E. L. Cates and F. Li, Balancing intermediate state decay rates for efficient Pr3+ visible-to-UVC upconversion: the case of β-Y2Si2O7:Pr3+, RSC Adv., 2016, 6, 22791–22796 RSC
.
- T. A. Johnson, E. A. Rehak, S. P. Sahu, D. A. Ladner and E. L. Cates, Bacteria Inactivation via X-ray-Induced UVC Radioluminescence:
Toward in Situ Biofouling Prevention in Membrane Modules, Environ. Sci. Technol., 2016, 50, 11912–11921 CrossRef CAS PubMed
.
- L. Li, P. Li, X. Lv, C. Cai, T. Li, X. Shi, D. Peng and Y. Yang, Ultraviolet-C mechanoluminescence from NaYF4:Pr3+, Appl. Phys. Lett., 2024, 124, 101108 CrossRef CAS
.
- X. Wang and Y. Mao, Emerging Ultraviolet Persistent Luminescent Materials, Adv. Opt. Mater., 2022, 10, 2201466 CrossRef CAS
.
- M. Laroche, A. Braud, S. Girard, J. L. Doualan, R. Moncorge, M. Thuau and L. D. Merkle, Spectroscopic investigations of the 4f5d energy levels of Pr3+ in fluoride crystals by excited-state absorption and two-step excitation measurements, J. Opt. Soc. Am. B, 1999, 16, 2269–2277 CrossRef CAS
.
- P. A. Tanner, C. S. K. Mak, M. D. Faucher, W. M. Kwok, D. L. Phillips and V. Mikhailik, 4f-5d transitions of Pr3+ in elpasolite lattices, Phys. Rev. B, 2003, 67, 115102 CrossRef
.
- Y. Wei, G. Xing, K. Liu, G. Li, P. Dang, S. Liang, M. Liu, Z. Cheng, D. Jin and J. Lin, New strategy for designing orangish-red-emitting phosphor via oxygen-vacancy-induced electronic localization, Light:Sci. Appl., 2019, 8, 15 CrossRef PubMed
.
- X. Lv, Y. Liang, Y. Zhang, D. Chen, X. Shan and X.-J. Wang, Deep-trap ultraviolet persistent phosphor for advanced optical storage application in bright environments, Light:Sci. Appl., 2024, 13, 253 CrossRef CAS PubMed
.
- A. A. Bol and A. Meijerink, Luminescence of nanocrystalline ZnS:Pb2+, Phys. Chem. Chem. Phys., 2001, 3, 2105–2112 RSC
.
- M. Mehnaoui, R. Ternane, G. Panczer, M. Trabelsi-Ayadi and G. Boulon, Structural and luminescent properties of new Pb2+-doped calcium chlorapatites Ca10−xPbx(PO4)6Cl2 (0 ≤ x ≤ 10), J. Phys.:Condens. Matter, 2008, 20, 275227 CrossRef CAS PubMed
.
- H. Li, Q. Liu, J.-P. Ma, Z.-Y. Feng, J.-D. Liu, Q. Zhao, Y. Kuroiwa, C. Moriyoshi, B.-J. Ye, J.-Y. Zhang, C.-K. Duan and H.-T. Sun, Theory-Guided Defect Tuning through Topochemical Reactions for Accelerated Discovery of UVC Persistent Phosphors, Adv. Opt. Mater., 2020, 8, 1901727 CrossRef CAS
.
- X. Zhao, F. Liu, Z. Yu, X. Li, C. Wang, F. Chen and X.-j. Wang, Sunlight stimulated solar-blind ultraviolet phosphor, Phys. Rev. Res., 2022, 4, L012028 CrossRef CAS
.
- Y. Zhang, S. Yan, F. Xiao, X. Shan, X. Lv, W. Wang and Y. Liang, Long-persistent far-UVC light emission in Pr3+-doped Sr2P2O7 phosphor for microbial sterilization, Inorg. Chem. Front., 2023, 10, 5958–5968 RSC
.
- Q. Liu, Z.-Y. Feng, H. Li, Q. Zhao, N. Shirahata, Y. Kuroiwa, C. Moriyoshi, C.-K. Duan and H.-T. Sun, Non-Rare-Earth UVC Persistent Phosphors Enabled by Bismuth Doping, Adv. Opt. Mater., 2021, 9, 2002065 CrossRef CAS
.
- L. Liu, S. Peng, Y. Guo, Y. Lin, X. Sun, L. Song, J. Shi and Y. Zhang, Designing X-ray-Excited UVC Persistent Luminescent Material via Band Gap Engineering and Its Application to Anti-Counterfeiting and Information Encryption, ACS Appl. Mater. Interfaces, 2022, 14, 41215–41224 CrossRef CAS PubMed
.
- Y.-M. Yang, Z.-Y. Li, J.-Y. Zhang, Y. Lu, S.-Q. Guo, Q. Zhao, X. Wang, Z.-J. Yong, H. Li, J.-P. Ma, Y. Kuroiwa, C. Moriyoshi, L.-L. Hu, L.-Y. Zhang, L.-R. Zheng and H.-T. Sun, X-ray-activated long persistent phosphors featuring strong UVC afterglow emissions, Light:Sci. Appl., 2018, 7, 88 CrossRef PubMed
.
- X. Lv, X. Shan, Y. Zhang and Y. Liang, Ambient light stimulation enabling intense and long-lasting ultraviolet-C persistent luminescence from Pr3+-doped YBO3 in bright environments, J. Mater. Chem. C, 2023, 11, 4492–4499 RSC
.
- S. Yan, Y. Liang, Y. Chen, J. Liu, D. Chen and Z. Pan, Ultraviolet-C persistent luminescence from the Lu2SiO5:Pr3+ persistent phosphor for solar-blind optical tagging, Dalton Trans., 2021, 50, 8457–8466 RSC
.
- S. Yan, Y. Liang, Y. Zhang, B. Lou, J. Liu, D. Chen, S. Miao and C. Ma, A considerable improvement of long-persistent luminescence in LiLuSiO4:Pr3+ phosphors by Sm3+ co-doping for optical tagging applications, J. Mater. Chem. C, 2022, 10, 17343–17352 RSC
.
- Y. Zhang, X. Shan, X. Lv, D. Chen, S. Miao, W. Wang and Y. Liang, Multimodal luminescence in Pr3+ single-doped Li2CaSiO4 phosphor for optical information storage and anti-counterfeiting applications, Chem. Eng. J., 2023, 474, 145886 CrossRef CAS
.
- M. H. Wang, H. H. Zhang, C. K. Chan, P. K. H. Lee and A. C. K. Lai, Experimental study of the disinfection performance of a 222-nm Far-UVC upper-room system on airborne microorganisms in a full-scale chamber, Build. Environ., 2023, 236, 110260 CrossRef
.
- Y. Lv, X. Chen, W. Wu, F. Wu, X. Wu, W. Yuan and C. Qu, Experimental and numerical study on upper-room Far-UVC system under different ventilation schemes to disinfect airborne microorganisms in indoor environments, Build. Environ., 2024, 266, 112108 CrossRef
.
- K. Guo, Y. Pan, H. F. R. Chan, K.-F. Ho and C. Chen, Far-UVC disinfection of airborne and surface virus in indoor environments: Laboratory experiments and numerical simulations, Build. Environ., 2023, 245, 110900 CrossRef
.
- J. Zhao, E. M. Payne, B. Liu, C. Shang, E. R. Blatchley, W. A. Mitch and R. Yin, Making waves: Opportunities and challenges of applying far-UVC radiation in controlling micropollutants in water, Water Res., 2023, 241, 120169 CrossRef CAS PubMed
.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B, 1996, 54, 11169–11186 CrossRef CAS PubMed
.
- G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B, 1993, 47, 558–561 CrossRef CAS PubMed
.
- M. Cococcioni and S. de Gironcoli, Linear response approach to the calculation of the effective interaction parameters in the LDA+U method, Phys. Rev. B, 2005, 71, 035105 CrossRef
.
- A. M. Srivastava, Aspects of Pr3+ luminescence in solids, J. Lumin., 2016, 169, 445–449 CrossRef CAS
.
- S. Zhang, F. Zhao, S. Liu, Z. Song and Q. Liu, An improved method to evaluate trap depth from thermoluminescence, J. Rare Earths, 2025, 43, 262–269 CrossRef
.
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.