Skip to main content

Next generation lanthanide doped nanoscintillators and photon converters

Abstract

Scintillators are of significance for the realization of indirect X-ray detection and X-ray excited optical luminescence (XEOL) imaging. However, commercial bulk scintillators not only require complex fabrication procedures, but also exhibit non-tunable XEOL wavelength and poor device processability. Moreover, thick crystals usually generate light scattering followed by evident signal crosstalk in a photodiode array. Lanthanide doped fluoride nanoscintillators (NSs) prepared with low-temperature wet-chemical method possess several advantages, such as low toxicity, cheap fabrication cost, convenient device processability and adjustable emission wavelengths from ultraviolet to visible and extending to second near infrared window. In addition, they exhibit X-ray excited long persistent luminescence (XEPL) making them suitable for broadening the scope of their applications. This review discusses and summarizes the XEOL and XEPL characteristics of lanthanide doped fluoride NSs. We discuss design strategies and nanostructures that allow manipulation of excitation dynamics in a core–shell geometry to simultaneously produce XEOL, XEPL, as well as photon upconversion and downshifting, enabling emission at multiple wavelengths with a varying time scale profile. The review ends with a discussion of the existing challenges for advancing this field, and presents our subjective insight into areas of further multidisciplinary opportunities.

1 Introduction

X-rays are electromagnetic waves with short wavelength and strong penetrability in physical matter including live organisms [1,2,3,4]. Scintillators that are capable of converting X-rays into ultraviolet (UV), visible or near infrared (NIR) photons [5, 6], are widely employed to realize indirect X-ray detection and XEOL imaging in medical diagnosis [7, 8], computed tomography (CT) [9, 10], space exploration [11, 12] as well as in non-destructive industrial material [13, 14] and security inspections [15, 16]. Commercial bulk scintillators, such as CaWO4 [17], NaI: Tl [18], (Lu,Y)2SiO5:Ce (LYSO:Ce) [19] and Bi4Ge3O12 (BGO) [20], possess high light yield (LY) and superior energy resolution; however, they suffer from serval drawbacks, i.e., complex fabrication procedures, expensive experimental equipment, nontunable XEOL wavelength and poor device processability [21, 22]. Moreover, they all produce emission in the visible spectral range, but having XEOL in the NIR range may find more interesting applications in biomedicine [23, 24]. Also, thick crystals usually generate light scattering followed by evident signal crosstalk in a photodiode array [22, 25, 26]. Recently, metal halide perovskites, which can be prepared by facile low temperature or even by room-temperature approaches have been investigated for X-ray detection [27, 28]. Unfortunately, these materials exhibited some intrinsic limitations as well, such as poor photo-/environmental-stability, heavy metal toxicity and low LY [29,30,31]. Thus, the search for developing new generation of scintillators is still one of the topic of the day.

Lanthanide doped fluoride NSs not only avoid the above mentioned limitations of bulk scintillators and metal halide perovskites, but also exhibit many useful properties. By employing a cheap and convenient wet-chemical method, the size, shape and core@shell structures of the lanthanide doped fluoride NSs can be tuned and designed on demand [32, 33]. Benefiting from the abundant energy levels of lanthanide activators, the emission wavelengths can be tuned from UV to visible and be extended to the second NIR window [34, 35]. Furthermore, these NSs show superior photostability [36,37,38,39], low toxicity [40, 41] and convenient device processability, which make them promising candidates for next generation NSs and in the emerging fields of XEOL imaging. Moreover, they exhibit XEPL property, which showing promising applications in biomedicine and optical information encoding [42, 43]. The combination of XEOL and XEPL making them suitable for broadening the scope of their applications [44]. In recent years, significant advances have been made in NS development, thus it is very timely to summarize present results for further enhancing their performances and expanding the applications.

In this review, we discuss the XEOL and XEPL characteristics of lanthanide doped fluoride NSs, which includes their mechanisms, tunable emission wavelengths from UV to visible then extending to second near infrared (NIR-II), performance indexes and optimization strategies as well as their corresponding applications. The emerging applications of NSs are summarized. We discuss design strategies and nanostructure that allow manipulation of excitation dynamics in a core–shell geometry that simultaneously produce XEOL, XEPL, as well as photon upconversion (UC) and downshifting (DS), enabling emission at multiple wavelengths and at varying time scales. A major feature of this review is summarizing the existing challenges in advancing this field, together with our subjective insight into areas of further multidisciplinary opportunities.

2 Fundamentals of scintillation

X-ray shows strong penetration ability even in dense matters [45, 46], which was first found by Wilhelm Roentgen in 1895 [47] and then broadly used in many fields [48, 49]. The hard X-ray (0.01–0.1 nm) is generally used for space exploration and non-destructive inspection of industrial equipment [11, 12], while the soft X-ray (0.1–10 nm) is widely employed in medical diagnosis and CT [9, 35]. Two types of X-ray detector are generally used to realize these practical applications; one is direct conversion of X-ray photons into electrical signals utilizing semiconductor materials (i.e., amorphous selenium) [50, 51]. The other one is indirect conversion into low-energy photons (i.e., ultraviolet and/or visible light) via scintillators, which are then detected by low cost arrayed photodetectors (i.e., amorphous Si photodiodes [52], photomultiplier tubes [53], silicon avalanche photodiode [54], charge-coupled devices [55], or complementary metal-oxide semiconductor [56]). Owing to rapid development of plentiful high-performance scintillators, cheap commercialized sensing arrays and flexible combinations, the indirect conversion route is becoming more popular for the detection of X-rays.

2.1 History and comparison of inorganic scintillators

Scintillators are the core component of an intact indirect X-ray detector device, which experienced more than hundred years of development [57,58,59,60]. Inorganic scintillators can be classified into intrinsic and extrinsic based on the XEOL mechanism [61, 62], or halides and oxides based on the anion type at present [63, 64]. The intrinsic XEOL mainly includes free-exciton luminescence (i.e., ZnO [65] and GaN [66]), self-trapped exciton (STE) luminescence (i.e., CaF2 [67], SrF2 [68] and BaF2 [69]), Auger free luminescence (i.e., BaF2 [69], BaMgF4 [70], CsF [71] and Cs2ZnCl4 [72]) and self-activation luminescence (i.e., BGO [73], CeF3 [74] and CeBr3 [75]). The extrinsic XEOL originates from doped activators, i.e., Tl+, Bi3+, In+, Sn2+, Mn2+, Cr3+ and lanthanide ions [76,77,78,79], or defects such as F-centers [80, 81], which offers the prospect to tune the emission wavelengths and improve the performance.

In the year 1895, Wilhelm Roentgen first reported that both barium platino-cyanide (Ba [Pt(CN)4]) and calcium sulphide (CaS) emitted visible photons upon X-ray irradiation which could pass through 15 mm thick aluminum sheet [47]. Then CaWO4 [82] and ZnS [83] were used as the first generation of scintillators. Because of the poor experimental conditions in this early stage, the occurrence of scintillation process was judged by the naked eye. During 1940s to 1960s, the second generation of scintillators, such as NaI (Tl) [84], CsI (Tl) [85], CaF2 (Eu) [86] and CdWO4 [87], was developed. CaF2 and CdWO4 are known as fluorite and scheelite in mineralogy, respectively [88]. After the development of BGO single crystal in 1973 [89], the Ce3+ activators incorporated third generation of scintillators were developed, i.e., Lu2SiO5: Ce (LSO: Ce) [90], LYSO: Ce [91] and Gd3(Al,Ga)5O12: Ce (GAGG:Ce) [92]. Up to date, BaF2, CdWO4, BGO, Gd2SiO5: Ce (GSO: Ce), LSO: Ce, LaBr3: Ce, GAGG: Ce, Lu3Al5O12: Ce (LuAG: Ce), Gd2O2S: Tb bulk scintillators and thick CsI (Tl) films are widely commercialized [61, 93].

Each kind of scintillator has its own advantages and disadvantages, for example, the NaI (Tl) crystals exhibits high LY (38,000–68,000 photons/MeV) [94], but poor water stability and are fragile [95]; BaF2 possesses ultra-fast decay rate [96], but low LY (10,400–11,800 photons/MeV) [97], are fragile and poor energy resolution [98]; GSO crystal shows higher LY than that of BGO and superior water stability [99], but requires expensive fabrication procedures as well as poor energy resolution than that of CsI (Tl) [100]. Furthermore, these traditional scintillators suffer from several common drawbacks. First of all, these scintillators are generally prepared via complex and expensive fabrication procedures. For examples, LYSO:Ce is mainly synthesized by the Czochralski method at temperatures higher than 1500 °C [101] and LuAG:Ce is prepared by vacuum sintering (< 10–3 pa) at 1750 °C [102]. Secondly, these crystals are inherently brittle and fragile, which are difficult to be coupled with flexible detectors for XEOL imaging [103]. Thirdly, the XEOL wavelengths are mainly located in the visible range and hard to be tuned owing to the fixed electron transition processes [104]. Fourthly, the large thickness of these scintillators brings light scattering and then leads to evident signal crosstalk in a photodiode array [25, 26]. At last, radioluminescence afterglow can cause image artifacts and influence the high-contrast X-ray imaging quality, especially in CT and modern digital radiography [105, 106].

Recently, a large number of halide perovskites have been studied as a new class of scintillators. The high-Z ions, such as Cs, Pb, Bi and I, endow them large X-ray absorption coefficient and high LY up to 90,000 photons/MeV at 77 K [107]. In addition, their facile preparation processes make them easier to be formed as polycrystalline films, single crystals, and nanocrystals [22, 108]. Interested readers are referred to a few recent reviews covering the contents of synthesis, mechanism and performances [109]. Although great achievements have been made on these scintillators, their poor photo-/environmental- stability, high toxicity of heavy metal ions such as Pb, and low room temperature LY greatly restrict their practical applications. In addition, XEOL in these systems mainly originates from band gap transition, thus, the emission wavelengths usually remain limited to the visible range.

Lanthanide ions possess plentiful ladder-like energy levels and can produce emission wavelengths from UV, to visible, to NIR [110, 111]. Alkali or alkaline-earth fluoride nanoparticles are widely employed for the incorporation of lanthanide ions with high solubility up to 100%, exhibiting superior photostability (i.e., no photobleaching), stable emission capability (i.e., no blinking) and low toxicity [112,113,114]. Attributing to their mature and facile wet-chemical synthesis routes, such as co-precipitation, solvo-/hydro- thermal, and thermal decomposition, it is convenient to tune the particle size, shape, crystal phase, and chemical compositions, to modify the hydrophobic or hydrophilic surface, and to construct multifunctional core@shell nanoarchitectures [32, 33, 115]. Moreover, the high-Z lanthanide ions produce enhancement of X-ray absorption efficiency and therefore LY. For example, the LY of hexagonal NaLuF4:Gd/Tb NSs was reported to be 39,460 ph/MeV [116]. These characteristics endow them next generation candidates for NSs and much suitable for the emerging fields of XEOL imaging and biomedicine, which will be discussed in the application section. Furthermore, the occurrence and tunability of XEPL make them suitable for broadening the scope of their applications. The history and comparison of common inorganic scintillators is shown in Fig. 1 [11, 86,87,88, 117,118,119,120,121,122,123].

Fig. 1
figure 1

History and comparison of common inorganic scintillators

2.2 Performance index of scintillator

X-ray absorption coefficient: The absorption coefficient (α) is the efficiency of a scintillator to absorb the incident X-ray photon energy in the conversion stage, which can be expressed by the following equation [1]:

$$\mathrm{\alpha }=\uprho {\mathrm{Z}}^{4}/{\mathrm{AE}}^{3}$$
(1)

where ρ, Z, A and E are density, atomic number, atomic mass and radiation energy, respectively. The X-ray stopping power (also named as attenuation coefficient) of a scintillator with a given thickness mainly depends on α. Thus, the scintillators with high density and high-Z usually show better scintillation property and more suitable for X-ray detection [109]. For example, the absorption coefficient of NaLuF4 (atomic number Zmax = 71, Kα = 63.31 keV) is larger than that of NaYF4 (Zmax = 39, Kα = 17.05 keV), hence the XEOL and XEPL intensities of NaLuF4:Tb NSs are stronger than that of NaYF4:Tb NSs [44].

LY: LY is one of the most important performance indices for scintillators [124]. Absolute light yield refers to the ratio of total energy of scintillation photons to energy deposited by ionizing radiation in a scintillator, and can be characterized by the number of emitted photons per each 1 MeV radiation energy absorbed by a scintillator [125]. The number of emitted photons (Nph) produced in scintillation can be calculated by the following equation [126]:

$$N_{ph}= \frac{E}{\beta {E}_{g}}\times SQ$$
(2)

where E (eV) is the energy of the incoming X-ray photon and Eg (eV) is the bandgap of the scintillator. S and Q are the quantum efficiencies in the transport and luminescence stages, respectively. β, indicating the average energy required to generate one thermalized electron–hole pair, is a phenomenological parameter which is typically found to be 2 ~ 3. Because a number of emitted photons might be lost before being captured by the photodetector, the actual light yield of a scintillator is always lower than the Nph value.

Response—decay time: The response time represents how fast the scintillator converts X-ray photons to emitted photons such as UV or visible [109]. The decay rate of the luminescence center itself is defined by its transition dipole moment from the excited state to ground state. Introducing nonradiative quenching or energy transfer processes away from the excited state can accelerate the decay rate, but also leads to the decrease of parameter Q in Eq. (2) and then reduces the light yield [127]. In the simplest case of an exponential decay, the XEOL intensity I(t) is

$$\mathrm{I }(\mathrm{t}) \sim \mathrm{ exp }[-\mathrm{t}/\uptau ]$$
(3)

where τ is called the decay time [103]. For parity and/or spin forbidden transitions of most trivalent lanthanide activators, the decay times are typically several tens of μs up to ms. In the case of allowed 5d-4f transitions, such as in Ce3+ and Eu2+, the decay times are down to tens of ns [101]. Both short response time and fast decay rate are essential for dynamic real-time X-ray imaging in some specific medical diagnostics such as CT.

Spatial resolution: The spatial resolution represents the contrast and the amount of blurs over a certain range of spatial frequencies of an image, which is generally evaluated by the modulation transfer function (MTF) [128]. A thinner scintillation layer benefits the realization of high spatial resolution, while a thicker one can improve the X-ray blocking ability [129]. Thus, the thickness of a scintillation layer should be optimized in practical applications.

Chemical stability and radiation resistance. Chemical stability is mainly referring to the hygroscopicity of scintillators, which determines their longtime operation in the open air environment. Radiation resistance of scintillators regards mainly the variation and instability of the performances upon long-time X-ray irradiation [103]. The microstructure of fluoride NSs, especially the surface coordination environment, can greatly influence their scintillation performances, which should be regarded as an important stability parameter.

Furthermore, for some specific applications, the linearity of XEOL response with the incident X-ray dose and intensity, the emission wavelength, the energy resolution and the cost are important criteria for the selection of scintillators as well [108, 122]. For dynamic real-time X-ray imaging, the afterglow of a scintillator should be restricted to avoid sickle artifacts [4].

3 Photon conversion and mechanisms in lanthanide doped fluoride NSs

Lanthanide doped fluoride NSs have been widely studied in the NIR-triggered UC and UV-excited DS fields [130,131,132], and have recently attracted great interests in the scintillation field [133, 134]. Although the final scintillation profiles of lanthanide activators are similar to those reported in UC and DS emissions, the electron population pathways to the excited levels are significantly different [134]. Owing to the high energy of X-ray photons and various possible scintillation processes, all the trivalent lanthanide activators from Ce3+ to Yb3+ except Pm3+ ions can be activated by X-rays. Thus, the XEOL emission wavelengths covering from UV, to visible and to NIR are achieved through doping appropriate lanthanide activators [126, 135]. In addition, employing a tetragonal matrix with a low crystal symmetry which benefits the energy level splitting, the XEOL spectra at room temperature might be used for the analysis of electronic fine structures of lanthanide activators [134]. The main emitting levels of various lanthanide activators are schematically illustrated in Fig. 2 [136,137,138,139]. Typically, with employing the low frequency phonon crystal lattices of fluorides as hosts, Ce3+, Pr3+, Nd3+ and Gd3+ ions can emit UV photons, Pr3+, Nd3+, Sm3+, Eu3+, Tb3+, Dy3+ Ho3+, Er3+, and Tm3+ ions can emit visible photons, Nd3+, Ho3+, Er3+, Tm3+ and Yb3+ ions can emit NIR photons.

Fig. 2
figure 2

a Partial 4f-energy level diagram for trivalent lanthanide activators. b Main luminescent transitions of the lanthanide activators in the electromagnetic spectrum

3.1 Mechanisms of XEOL processes in lanthanide doped fluoride NSs

The role of X-ray can be classified into chemical, biological, and physical effects such as penetration, ionization, fluorescence, thermal as well as optical interference, diffraction, reflection and refraction [140,141,142,143]. The interaction between X-rays and matter through an ionization process consists of photoelectric effect, Compton scattering, Thomson scattering, Rayleigh scattering and electron pair effect. Upon X-ray irradiation, an electron in the inner shell of an atom is excited and then escapes out of the atom, which is known as the photoelectric effect and is dominated when the incident X-ray photon energy is below ~ 500 keV [144]. When the incident X-rays with photon energy from a few hundred keV to several MeV interact with matter, both the energy and the direction are changed, a phenomenon which was first discovered by Arthur Holly Compton in 1923 and then named as Compton scattering [109, 145]. In contrast, Thomson scattering is an elastic and coherent interaction, which occurs between a low energy photon and a single free electron [146]. Rayleigh scattering is the superposition of multiple Thomson scattering, in which the incident photon energy is not changed but its directions is altered [147]. Both the Thomson scattering and Rayleigh scattering decrease with an increase of the irradiation energy. The electron pair effect represents the generation of a positive and negative electron pair during the interaction between the incident X-ray photons and the nucleus of an atom, which happens when the irradiation energy is over 1.02 MeV. Generated by the X-ray of 1.02 MeV, one positive electron interacts with a surrounding negative electron via annihilation generating two photons with the energy of 0.51 MeV [148,149,150].

The XEOL processes in lanthanide doped fluoride NSs are similar to those of activators doped bulk scintillators, which mainly include three stages of conversion, transport and luminescence, as illustrated in Fig. 3 [60, 103]. Generally, the maximum X-ray photon energy is equal to the tube voltage multiplied by the electron charge. At the first stage of conversion, the incident X-ray photons interact with the lattice heavy atoms (i.e., Cs+, Ba2+, Ln3+, Bi3+) to generate hot electrons and deep holes, mainly through the photoelectric effect and Compton scattering effect. Massive secondary electrons are then produced via electron–electron scattering and the Auger process, resulting in the generation of lower kinetic energy charge carriers. These energetic charge carriers are thermally dissipated via interacting with phonons, and then a large amount of low energy electrons and holes are populated in the conduction band and valence band, respectively. This conversion stage takes place in subpicoseconds [109].

Fig. 3
figure 3

Schematic illustration of the scintillation process in lanthanide doped fluoride NSs. CB, conduction band. VB, valence band. ET, energy transfer between lanthanide ions [109]

At the second stage of transport, the produced large number of electrons and holes are transported to luminescence centers, captured by defects such as ionic vacancies, surface defects or Frenkel defects, or be self-trapped in the crystal lattice. This process usually takes place in 10–12 ~ 10–8 s. The defects either increase the nonradiative possibilities of the lanthanide activators, which leads to the decrease of XEOL intensity, or contribute to the production of afterglow [44]. Generally, limited afterglow benefits the increase of the signal-to-noise ratio and then produces high contrast images without lag, which is of significance for real-time dynamic XEOL imaging [105, 106]. However, bright and long XEPL show promising application in biomedicine, which will be discussed in the application section.

At the last stage of luminescence, after the excited levels of lanthanide activators are populated via the absorption of energy from the recombination of the low kinetic energy carriers, XEOL is generated through 5d-4f or 4f-4f transitions. It is emphasized that the electrons deposited in the defect states can migrate to the excited levels through energy tunneling during X-ray irradiation, which will influence the continuous XEOL stability [44, 151]. Different from traditional bulk scintillators and metal halide NSs, it is facile to introduce energy transfer processes in lanthanide doped fluoride NSs via a codoping strategy and inhibit energy migration from activators to surface quenchers through coating shells, so as to improve the XEOL intensity [152, 153]. Moreover, through doping multiple activators in different layers of core/multi-shell fluoride NSs, multiple emission wavelengths produced by upconversion as well as down shifting can be realized, but it has not been extensively studied in XEOL until now.

3.2 X-ray to UV

UV emissions in lanthanide ions are mainly originated from the 5d-4f transitions, such as in Ce3+ and Pr3+ [122, 154]. The Ce3+ ion with the 4f1 configuration shows efficient broadband emission corresponding to the 5d-4f parity allowed electric dipole transition. The energy gap between the 4f ground state and the 5d excited state of the free Ce3+ ion is about 6.2 eV (50 000 cm−1) [155, 156]. Since the 5d orbitals are susceptible to their surrounding crystal field, the symmetry, the anion polarizability, covalence and the surrounding surface ligands, the excitation and the emission wavelengths can be modulated by changing the chemical composition and the structure of the hosts [122, 157]. In common fluoride crystals, such as alkali/alkaline-earth rare-earth fluorides, or alkaline-earth fluorides, the doped Ce3+ ions are generally used to absorb the UV photons via 4f-5d transition and then sensitize Sm3+, Eu3+, Tb3+ and Dy3+ ions [158, 159]. Upon X-ray irradiation, the broad UV emission in Ce3+ from 280 to 350 nm, which contains two emission peaks of ~ 308 nm originated from 5d → 4f (2F5/2) and ~ 325 nm originated from 5d → 4f (2F7/2) transitions, was recorded in the tetragonal LiYF4:Ce microcrystals. This strong UV emission was further used to activate wavelength shifters such as CdSe/ZnS quantum dots to modify the XEOL emission across the visible range. Their reported polymer composites were physical mixtures of LiYF4:Ce microcrystals and CdSe/ZnS quantum dots [160]; a controlled synthesis of heterostrcutures between them might improve the XEOL intensity. For extending the XEOL wavelength to UVC range, the cubic elpasolite Cs2NaYF6 host with high defect-bearing characteristic was employed to incorporate Pr3+ ions. As a result, the UVC emission peaking at 250 nm corresponding to the Pr3+: 4f5d → 3H4 transition was achieved [161]. This reported Cs2NaYF6 host was synthesized via a solid state reaction where the structures such as size, shape and surface molecules were hard to be modified. In such case, the structure—XEOL property relationships in this host are not studied yet. In addition, in the LuF3 [162] or BaLu2F8 [163] single crystal, sharp emission of Nd3+ (~ 170 nm) corresponding to the 5d-4f transition was observed.

Except 5d-4f transitions, the radiative transitions from high energy 4f levels can result in UV emission as well, such as in Gd3+ and Nd3+. For example, in the hexagonal NaLuF4:Gd/Tb NSs, the Gd3+ ions were used to promote the population of Tb3+ ions via energy transfer and then enhance the XEOL intensity, while its emission peaking at 311 nm corresponding to the 6P7/2 → 8S7/2 transition was recorded simultaneously. The XEOL intensity of Gd3+ ions were much weaker than that of Tb3+ ions, which might be attributed to energy transfer from Gd3+ to Tb3+. With codoping Gd3+/Nd3+ ions in the NaLuF4 NSs, the emission at 385 nm corresponding to Nd3+: 4D3/2 → 4I11/2 transition was observed [44].

3.3 X-rays to visible

The 4f-4f transition in a 4fN electronic configuration is in principle electric dipole forbidden by the Laporte’s rule. However, since the sites occupied by lanthanide ions in the fluorides do not present a center of inversion, Laporte’s rule is relaxed to some extent due to odd parity terms in the ligand field Hamiltonian [164]. Thus, various visible emissions in lanthanide ions corresponding to the 4f-4f transitions are observed after the absorption of X-rays, whose profiles are much similar to those of common UC and DS in luminescent materials. For examples, characteristic transitions originated from Tb3+ (green, 5D4 → 7F3-6), Eu3+ (red, 5D0 → 7F0-6) and Sm3+ (red, 4G5/2 → 6H5/2–11/2) ions were detected when employing the NaLuF4 NSs as the host. The Tb3+ ion usually exhibits stronger XEOL intensity than those of Sm3+, Eu3+ and Dy3+ ions in a same host [44, 165], and the green wavelength matches well with the visible photodetectors. Thus, Tb3+ ions incorporated scintillators are broadly studied [44].

Similar to the Ce3+-sensitized DS luminescent materials, Gd3+ ions might be used as efficient bridge centers to improve the XEOL intensities of Tb3+, Sm3+, Eu3+ and Dy3+ ions as well [108]. Because X-ray photons possess much higher energy than that of UV photons, the strong XEOL emissions of Er3+, Ho3+ and Tm3+, which are widely used in UC materials, are easier to be recorded in the doped NaLuF4 NSs. Compared with bulk materials, NSs have a larger surface-to-volume ratio and more surface defects [166, 167]. Building a core@shell architecture has been verified to be an effective route to restrict the energy migration from activators to surface quenchers, followed by the decreased nonradiative transition possibilities and improved luminescence intensity [140, 168, 169]. The influence of shell thickness on UC and XEOL is quite different. For the Yb/Tm: NaYF4 NSs with a mean particle size of 30 nm, the optimal NaYF4 shell thickness for bright UC emission intensity but small size was found to be about 6.3 nm [170]; for the Yb/Er: NaYF4 NSs (ellipsoids with a length of 25 nm and a width of 21 nm), the optimal NaYF4 shell thickness was about 5.5 nm [171]. It should be noted that the UC emission intensities did not decrease with further increasing the shell thickness over the optimal value. However, it was reported that the XEOL intensity of Eu3+ ions doped LuF3 NSs (~ 4.4 nm) was evidently decreased when the LuF3 shell thickness increased over ~ 1 nm [172]. Although it was previously speculated that a thicker shell could increase the fraction of non-emissive volume in the NSs and then lower the probability of radiative recombination of charge carriers at the luminescent centers followed by the decreased of XEOL intensity, but it still needs more theoretical and experimental results to clarify this phenomenon.

3.4 X-rays to NIR

Under X-ray irradiation, NIR emissions originating from the Nd3+ (4F3/2 → 4I9/2, ~ 866 nm/ ~ 893 nm; 4F3/2 → 4I11/2, ~ 1064 nm), Ho3+ (5I6 → 5I8, ~ 1180 nm), Er3+ (4I11/2 → 4I15/2, ~ 980 nm; 4I13/2 → 4I15/2, ~ 1525 nm), Tm3+ (3H4 → 3H6, ~ 800 nm; 1G4 → 3H4, ~ 1120 nm; 3H4 → 3F4, ~ 1475 nm) and Yb3+ (2F5/2 → 2F7/2, ~ 980 nm) ions were recorded by employing the NaYF4 NSs as host [173]. NIR luminescence exhibits deeper penetration depth than those of UV/visible light, which is suitable for biomedical application. Achieving strong XEOL emission in the NIR range under a low irradiation dose is of significant importance from the view point of biosafety [174, 175]. Modulating the chemical compositions and constructing a core@shell structure in the lanthanide doped fluoride NSs are general feasible routes to enhance the XEOL emission intensity [42, 176]. Because the radiation energy of the recombination of charge carriers is much larger than that of NIR photons, the high energy levels of lanthanide activators are prior populated, and then multi-step nonradiation relaxations or cross-relaxations are required to produce NIR emissions [173]. Thus, introducing appropriate assisted energy levels to increase the electrons population in the low excited energy levels is a useful tool for the enhancement of XEOL emission in the NIR range.

4 X-ray excited persistent luminescence

4.1 XEPL mechanisms

Long persistent luminescent materials, also known as afterglow luminescent materials, have attracted great attention over the past few decades [177]. Long persistent luminescence refers to an optical phenomenon in which the stored excitation energy in traps is released to produces UV, visible or NIR photons which lasts for minutes, hours or even days after ceasing the excitation. The formation of suitable charge carrier traps such as type, concentration and depth in the crystal structure plays a vital role in the XEPL process [178]. Anion Frenkel defects were recognized as traps in fluoride hosts, such as NaLuF4 and NaYF4. The dislocation of fluorine anions (F) to interstitial sites through elastic collisions with large-momentum X-ray photons leads to the formation of fluorine vacancies (E-traps) and interstitials (H-traps), accompanied by the production and trapping of many energetic electrons in Frenkel defect-related trap states [44, 151]. The XEPL mechanism in lanthanide doped fluoride NSs is illustrated in Fig. 4. The X-rays induced low kinetic energy electrons and holes in the conversion stage of the XEOL process are partially captured by the traps. After ceasing the X-rays, the captured electrons and holes are released from the traps to the conduction band and the valence band, respectively, due to stimulation by heat. The excited levels of lanthanide activators are then populated via the recombination of these electrons and holes. Finally, the XEPL is generated through 5d-4f or 4f- 4f transitions.

Fig. 4
figure 4

a Schematic illustrating creation of Frenkel-related trap states in fluoride crystal lattices upon X-ray irradiation [44]. Copyright 2021 Springer Nature Limited. b The XEPL mechanism in lanthanide doped fluoride NSs [151]

4.2 XEPL in fluoride hosts

UVC light in the 200–280 nm range shows the capability of germicidal activity, which can be used to kill bacteria, viruses, and other pathogens by destroying nucleic acids and disabling their ability to multiply [161, 179]. To generate UVC persistent luminescence, the host should have a large bandgap, thus, X-ray is appropriate as a stimuli source. The band gap of pristine Cs2NaYF6 is ~ 9.67 eV, which was verified to be suitable for the generation of XEPL in the UVC range from Pr3+ activators. As shown in the Fig. 5a, the XEPL peaking at 250 nm (4f5d → 3H4), 270 nm (4f5d → 3H5), 486 nm (3P0 → 3H4) and 610 nm (3P0 → 3H6) were recorded in the Cs2NaYF6: Pr3+. The existence of fluorine vacancy-related defect levels was verified by density functional theory calculations and thermoluminescence analysis. The trap depths (E) relative to the conduction band minimum (CBM) was calculated by the formula of E = 0.002Tm, in which Tm is the temperature corresponding to the maximum thermoluminescence peak. Three shallow traps, with respect to those > 1.3 eV below the CBM, were calculated to have the activation energies of 0.69, 0.83, and 1.02 eV at 70, 141, and 238 °C, respectively. The XEPL at 250 nm lasts over 2 h, and the intensity is still over one order of magnitude stronger than the background signal after 2 h (Fig. 5d, g). This XEPL intensity could be improved by tuning the Pr3+ ions concentration and the X-ray irradiation duration [122].

Fig. 5
figure 5

a XEPL spectra of the Cs2NaY0.99F6: 0.01Pr3+ recorded at different times [161]. Copyright 2018 Springer Nature Limited. b XEOL spectra of the NaLuF4: Tb3+@NaYF4 core@shell nanocrystals, recorded after cessation of X-rays (50 kV) for 0.5–168 h or 30 days [44]. Copyright 2021 Springer Nature Limited. c Tunable XEPL spectra of lanthanide-doped NaYF4 based core@shell NSs [173]. Copyright 2021 Springer Nature Limited. d XEPL decay curve of the Cs2NaY0.99F6: 0.01Pr3+ detected at 250 nm [161]. Copyright 2018 Springer Nature Limited. e XEPL decay curve of the NaLuF4:Tb3+ and NaLuF4: Tb3+@NaYF4 nanocrystals monitored at 546 nm [44]. Copyright 2021 Springer Nature Limited. f NIR-II XEPL decay curves of typical emission bands of lanthanide-doped NaYF4 based core@shell NSs [173]. Copyright 2021 Springer Nature Limited. g UVC images of the Cs2NaY0.99F6: 0.01Pr3+ phosphors taken at different afterglow times [161]. Copyright 2018 Springer Nature Limited. h NIR-II XEPL images of lanthanide-doped NaYF4 based core@shell NSs [173]. Copyright 2021 Springer Nature Limited. i XEOL and XEPL images of NaLuF4 NSs doped with various activators [44]. Copyright 2021 Springer Nature Limited

The XEPL in the visible range has been achieved in a series of trivalent lanthanides, such as Pr3+, Nd3+, Sm3+, Tb3+, Dy3+, Ho3+, Er3+ and Tm3+. For example, a bright and ultra-long (over 30 days) green XEPL was reported in the NaLuF4:Tb@NaYF4 core@shell NSs with a mean particle size of 27 nm (Fig. 5b, i). After coating the shell, the XEOL and XEPL intensities were enhanced by 1.5-fold and 6.5-fold, respectively (Fig. 5e), both of which were much stronger than in the commercial plastic scintillators and conventional persistent phosphors, including SrAl2O4:Eu2+/Dy3+ powder, ZnS:Cu2+/Co2+ powder, SrAl2O4:Eu2+/Dy3+ NSs and ZnGa2O4:Cr3+ NSs. First-principles calculations based on the density functional theory suggested that interstitial fluorine ions gradually diffuse back to original vacancies when the proximity between them is less than 3 Å. When the separation of these defect pairs is over 3 Å, the interstitial fluorine ions can be stabilized due to increased energy barriers, except under stimulation with heating or light exposure [44]. Similar to alkali rare-earth fluorides, the alkaline-earth fluorides are also used as hots for the realization of XEPL. The core CaF2: Dy NSs (~ 13.1 nm) exhibited extremely weak XEPL, whose intensity was significantly enhanced after coating an inert NaYF4 shell. Its intensity increased gradually with an increase of the shell thickness from 0.25 nm to 2 nm and then reached a plateau, but it was not evidently changed with tuning the shell compositions including NaYF4, NaYF4:Yb/Er, NaYF4:Yb/Tm, NaYF4:Ca, NaGdF4, NaYbF4 and LiYF4 [180].

Through doping Nd3+, Ho3+, Er3+ or Tm3+ ions in the NaYF4 or NaGdF4 NSs, NIR-II XEPL were achieved, which lasted more than 72 h (Fig. 5c, f, h). The initial XEPL intensity enhanced with an increase of the X-ray dosage and reached a plateau at higher dosages of X-rays (> 200 Gy), which was probably due to a limitation in the harvesting of the excitation energy by the NSs. The morphology and the crystal structure showed high photostability during cycles of X-ray recharging. For the optimal NaYF4:3Er@NaYF4 NSs with 35 nm cores and 7 nm shells, its XEPL intensity was approximately 25-fold than that of the core NSs of 22 nm [173]. Comparison of afterglow duration in different lanthanide doped fluoride systems is shown in Table 1.

Table 1 Comparison of XEPL among various of lanthanide doped fluoride NSs

4.3 XEPL tuning strategies

The XEPL wavelength, intensity and duration should be modified to satisfy a specific application; for example, the XEPL intensity should be strictly limited for real-time dynamic imaging, but need to be strong enough for bio-sensing [176]. The 4f-4f transitions of lanthanide ions are almost independent of the local chemical environment owing to the shielding of 4f electrons by the outer 5s and 5p shells [122, 184], the XEPL wavelength is mainly tuned by doping different activators. It is possible to tune the XEPL wavelength of the Ce3+ ion by structural engineering because its 5d-4f transition is susceptible to the local environment [163]. There are several structural and external stimuli parameters which can be used to modify the XEPL intensity and duration: (1) A core@shell structure which is an effective route to passivate the NS surface defects and then enhance the XEPL intensity [180]; (2) Crystal phase of the host, i.e., the XEPL intensity of the NaYF4:3Er in the hexagonal phase was more than ten times stronger than that in the cubic phase [165]; (3) Crystal structure of the host, i.e., with increasing the Gd3+ doping content in the NaYF4:Gd/3Er@NaYF4 from 0 to 97 mol%, the XEPL intensity first decreased and then increased, which might be ascribed to the increased lattice defects that origin from the atomic size mismatch between Gd3+ and Y3+ [173]. (4) The activators doping concentration, core size, shell thickness, irradiation dose and time, trap depth as well as storage temperature could also influence the XEPL [44]. (5) Codoping Ce3+ ions was used to reduce the electron population in the excited energy levels of lanthanide ions and then significant weaken the XEPL [108]. Accompany with these initial results, more efforts are needed to clarify the related mechanisms, and it is remaining a need to develop novel methods to significantly amplify or restrict the XEPL for specific applications.

4.4 Simultaneous XEOL, XEPL, UC and DS emissions using multiple shells

Lanthanide doped fluoride NSs have been broadly used to generate NIR triggered photon UC and UV excited DS emissions in addition to XEOL and XEPL (Fig. 6a) [130,131,132]. Core/multi-shells nanostructure can be employed to simultaneously produce XEOL, XEPL, as well as UC and DS, which allow researchers to manipulate the excitation dynamics of lanthanide activators. Because the electron population pathways in these processes are significantly different, several interesting physical processes such as excitation-dependent multiple emissions and thermal-response multicolors are facile to be manipulated. For example, when different lanthanide activators are used to generate diverse emission wavelengths of XEOL, UC and DS in core@shell@shell NSs (Fig. 6b), plentiful multicolors can be modulated on demand through controlling the excitation wavelength and/or power, which show promising applications in dynamic display and multi-level anti-counterfeiting. It should be noted that the relative intensities of XEOL, UC and DS can be tuned by changing the spatial distributions of the activators in core@shell@shell structures. Especially, through modulating the incident photons at different timescale based on a specific requirement, time-response characteristic transitions can be designed, which show promising in multi-mode bio-imaging and bio-sensing.

Fig. 6
figure 6

a XEOL, XEPL, DS and UC processes in lanthanide doped fluoride NSs. b Schematic illustration of the multimode color evolution based on fluoride core@shell@shell NSs. P represents excitation power

5 Applications

5.1 XEOL imaging

The fundamental working principle of XEOL imaging is to record the attenuation of X-rays after penetrating the specific subjects by using a scintillator and then imaging with a camera (Fig. 7). The scintillator screen is placed under the target to absorb the transmitted X-ray photons. For examples, a low dose of X-rays penetrating live organisms enables the application of computed tomography, while penetrating nonliving matter enables product quality and security inspection [7,8,9,10,11,12,13,14,15,16]. The X-ray irradiation dose should be low enough to assure the safety, while the high resolution and distinct contrast are important for image analysis. The maximum permissible dose for gonads and red bone marrow is ~ 5 rem/year; for skin, bone and thyroid is ~ 30 rem/year; for hands, forearms, feet and ankles is ~ 75 rem/ year; for any other single organ is ~ 15 rem/year [185].

Fig. 7
figure 7

Schematic illustration of the X-ray imaging system

The ethylenediaminetetraacetate (EDTA) capped NaGdF4:Ce/Tb NSs exhibited reduced afterglow and highly efficient XEOL, which was demonstrated for high resolution XEOL imaging. The Ce3+ dopants can promote the energy migration from a trap center to surface quenchers via the Gd3+ sublattice and then greatly reduce the population in traps to produce weak afterglow. When employing an ultrathin transparent NaGdF4: Ce/Tb involved film (0.045 mm) as a NS screen, a high spatial resolution of 18.6 lp mm−1 (lp: line pairs) was achieved. The MTF of a image is calculated by the slanted-edge method to reveal the spatial resolution [186, 187]. The spatial resolution is defined to be the spatial frequency at MTF = 0.2. The formula is as follows:

$$MTF\left(v\right)=F\left(LSF\left(x\right)\right)=F\left(\frac{dESF(x)}{dx}\right)$$
(4)

where \(\nu \) is the spatial frequency, and \(x\) is the position of the pixels. The configuration inside of circuit boards with different electronic components, the thread of screw inside an opaque capsule, obvious biological tissue phase contrast and the bone shape of a chicken claw were clearly imaged by a simple XEOL imaging system (Fig. 8) [108]. It was reported that the LiLuF4 possessed a higher absorption coefficient than those of CdTe, BaF2, CdZnTe, and CsPbBr3 materials at a higher photon energy, and the detection limit of LiLuF4:15 Tb NSs was measured to be 36.31 nGy s−1, which is much lower than typically used for X-ray diagnostics (5.50 µGy s−1). The LiLuF4:15 Tb NSs involved film was verified to be used for XEOL imaging to reveal the details of circuit board with a spatial resolution better than 20 lp mm−1 [188]. For dynamic real-time X-ray imaging, it remains a great challenge to develop divalent lanthanide or metal transition activators doped fluoride NSs with fast decay time.

Fig. 8
figure 8

Copyright 2021 American Chemical Society

a Photograph of a circuit board A (top) and its XEOL image (below). b Photograph of circuit board B (top), its XEOL images with using NaGdF4: 10Ce/18 Tb film (below, left) and CsI (Tl) single crystal (below, right). c Photographs of a capsule containing a screw inside and its XEOL image. d Photographs of one peanut and its XEOL image. e XEOL image of a chicken claw. (dose rate: 42.08 mGyair s−1, voltage: 50 kV) [108].

Lanthanide-doped NSs featuring strong XEOL intensity and long XEPL were studied for X-ray luminescence extension imaging (Xr-LEI) as well (Fig. 9). By employing a flexible detector prepared by embedding NaLuF4:15 Tb@NaYF4 NSs into a polydimethylsiloxane (PDMS) substrate, the internal structures of a highly curved electronic circuit board was visualized and high-resolution 3D Xr-LEI was achieved though combining XEPL and graphical simulations, while only overlapped imaging of the electronic circuit board was recorded when using a typical flat-panel X-ray detector (Fig. 9c). The achieved spatial resolution was more than 20 lp mm−1, which is much higher than that of conventional flat-panel X-ray detectors (typically less than 5 lp mm−1) (Fig. 9e, f) [44]. Xr-LEI is an interesting and effective route to reveal the three dimensional structures of objects, which is inaccessible by conventional flat-panel X-ray detectors or synchrotron-based X-ray microscopy.

Fig. 9
figure 9

Copyright 2021 Springer Nature Limited

a Schematic description of 3D electronic scintillation imaging through a NSs based stretchable X-ray detector. b Xr-LEI of a 3D circuit board by the NaLuF4: Tb3+(15 mol%)@NaYF4 based flexible X-ray detector (50 kV, 80 °C). c X-ray imaging on the same electronic board by a traditional packaged flat panel detector. d Xr-LEI of inner circuits of an iPhone. Inset: digital photo of the same integrated circuits. e Photograph of the flexible scintillation layer. f High-quality Xr-LEI by applying the flexible X-ray detector [44].

5.2 XEOL based biomedicine

X-ray, as an ionizing radiation with deep penetration depth in human body, has been broadly studied for radiotherapy and bioimaging applications [1,2,3,4, 167]. Benefiting from the feasible surface modification, various photosensitizers such as rose bengal (FDA-approved), verteporfin, ZnO, and choline e6 can be linked on the fluoride NSs via chemical conjugation or electrostatic adherence. The strong XEOL can activate the photosensitizers to generate reactive oxygen species, which then directly slow or even stop tumor growth by photodynamic therapy causing inflammation and compromising microvasculature [112].

The Ce3+ ions involved systems, such as CeF3@verteporfin [189], CeF3@ZnO [190], CeLaF3/LaF3@Chlorine e6 [191], CeF3:Tb@CTAB-Chlorine e6 [192], CeF3:Gd/Tb@rose Bengal [110], LaF3:Tb@rose Bengal [193], NaCeF4:Gd,Tb [152] NSs were widely studied in X-ray response biomedicine. Loading rose Bengal on the mesoporous silica-coated CeF3:Gd/Tb NSs was used to generate efficient 1O2 upon X-ray irradiation [112]. This product can be employed as a multifunctional tool capable of being used synergistically with dual-modal imaging (CT and magnetic resonance imaging), guided combined non-radioactive radiotherapy (RT) and X-Ray effected photodynamic therapy (XPDT) under a low X-ray irradiation dose. Combining synergistic RT + XPDT treatment modalities with untargeted global metabolomics can be used for the analysis of relevant metabolic tumor and serum biomarkers and their patterns. The synergistic RT + XPDT of NaCeF4:Gd/Tb NSs showed better tumor inhibition efficiency than RT alone in both A549 lung tumor model and CT26 colon cancer model (Fig. 10) [152]. In addition, the NaYF4:Gd/Tb NSs was reported as a light transducer for depth-independent NO release and on-demand gas-sensitized cancer therapy. Owing to the energy transfer from NaYF4:Gd/Tb NSs to the NO donor (RBS), ultralow soft X-ray dosage (~ 0.85 mGy) triggered NO release was realized in deep tissues even up to 3 cm depth, which broke the depth limitation suffered by the traditional UV/vis and NIR light. Upon X-ray irradiation, the NaYF4:Gd/Tb-RBS agent could be used for the inhibition of tumor growth [194].

Fig. 10
figure 10

Copyright 2019 American Chemical Society

Digital photographs of mice on the 30th day after different treatments, 2′,7′- dichlorofluorescein diacetate staining of A549 tumor for ROS (reactive oxygen species) visualization, images of Hematoxylin and eosin H&E and TdT-mediated dUTP nick end labeling stained tumor slices from different groups of mice and images of anti-Bax and anti-Bcl-2 stained tumor slices from different groups of mice from top to bottom. The energy and dose rate of X-ray were 160 kV and 1.103 Gy/min respectively [152].

XEOL bio-medical imaging takes advantage over traditional fluorescence imaging due to the unlimited tissue penetration depth of X-rays. UV–vis and NIR can only penetrate a depth of 0.5–2.5 mm and 8–10 mm, respectively, while X-rays can easily reach several-centimeter penetration [176]. Although a great achievement has been achieved in the photoluminescence imaging field, some intrinsic disadvantages still exist, such as poor anatomical and physiological detail in vivo. Fluoride NSs can not only be used as CT contrast agents for CT imaging [195,196,197,198], but also employed for CT and photoluminescence imaging simultaneously to realize precision diagnosis. For example, NaYF4:Nd3+@NaLuF4@polydopamine nanocomposites can be used in NIR-II optical and X-ray CT dual-modal imaging, which generated clearer in vivo tumor information than the grayscale magnetic resonance imaging (Fig. 11a) [199]. Moreover, tri-model biological imaging was realized in the NaGdF4:Yb,Tm@NaGdF4 nanosystem. 980 nm laser activated energy transfer UC from Yb3+ to Tm3+ was used for high-resolution fluorescence imaging at the cellular level, X-ray CT was used for the revelation of tumor localization and magnetic resonance imaging originated from the paramagnetic Gd3+ ions was used for providing superior details and functional information about tissues (Fig. 11b) [200]. For the traditional bulk scintillators, in addition to the limitations discussed in the introduction section, they cannot be used for in vivo biomedicine. However, the bio-combinable lanthanide doped fluoride NSs show great promises in bio-medicine field. One of the most challenges is developing facile strategy to improve the LY, so as to reduce the employed dose rate of dangerous X-rays.

Fig. 11
figure 11

a X-ray CT images of the HeLa tumors-bearing nude mice before (top) and after (bottom) intratumoral injection of NaYF4:Nd3+@NaLuF4@polydopamine [199]. Copyright 2017 American Chemical Society b. UC images of HeLa cells incubated with NaGdF4:Yb,Tm@NaGdF4 at 37 °C for 0.5, 1, and 3 h. Scale bar: 50 μm (top left). In vivo T1-weighted magnetic resonance images of tumor-bearing mice before and after injection (top right). In vivo CT images of a tumor-bearing mouse before and after injection (bottom) [200]. Copyright 2019 Royal Society of Chemistry

5.3 XEPL based biomedicine

The XEPL in UVC range can be used for sterilization and in vivo killing of pathogens and cancer cells [179]. Fluorides with large band gap and facile creation of anionic defects are appropriate for the generation of UVC persistent luminescence from 5d → 4f transitions of lanthanide activators. Through doping Pr3+ ions into the defect-bearing cubic elpasolite Cs2NaYF6, bright XEPL peaking at ~ 250 nm (4f5d → 3H4) can last more than 2 h. Experimental characterizations combined with first-principles calculations suggested that oxygen introduction-induced fluorine vacancies acted as electron traps. As shown in Fig. 12a, 100% viability was maintained when keeping Pseudomonas aeruginosa PAO1 under ambient conditions (i.e., room light, normal atmosphere) for 30 min, while it was greatly decreased when increasing the XEPL intensity upon prolonging the irradiation time [161].

Fig. 12
figure 12

a Inactivation of P. aeruginosa PAO1 using UVC-afterglow phosphor sheets. They were the control sample, irradiated by X-ray for 2 min, 5 min, 10 min, 16 min to inactivate P. aeruginosa PAO1 and the dependence of P. aeruginosa PAO1 survival ratios on the X-ray irradiation time of the given phosphor sheets, respectively. The live and dead cells show green and red colors, respectively [161]. Copyright 2018 Springer Nature Limited. b Dual-channel in vivo PL imaging of organs in a living mouse. Nd-doped NSs were used to image the gastrointestinal tract through oral administration, and Er-doped NSs were used to image visceral organs via a tail vein injection. Scale bar, 1 cm [173]. Copyright 2021 Springer Nature Limited

Tunable XEPL in the NIR-II window was realized in the NaYF4: (Nd3+, Ho3+, Er3+ or Tm3+)@NaYF4 core@shell NSs, which are promising for in vivo imaging applications, including high-contrast abdominal vessels, tumor imaging and ureter tracking, as well as multispectral in vivo deep-tissue viscera imaging and the multimodal XEPL-magnetic resonance-positron emission tomography imaging of tumors. As shown in the Fig. 12b, Er-doped NSs were first injected into a living mouse via the tail vein, followed by gavaging of Nd-doped NSs after 10 min; as a result, the overlayed NIR-II XEPL image clearly showed the main organs with a high contrast [173]. High resolution is significantly important in medical imaging. The full-width at half-maxima of 331.1 μm and 457.4 μm were achieved in the NIR-II XEPL imaging, which were 0.83-fold sharper than those obtained with NIR-II fluorescence imaging [173].

5.4 Information encoding

Fluoride NSs are one of most suitable hosts for UC, thus, dual optical properties of XEPL and upconvesion are facile to be realized by a core@shell structure, such as CaF2:Dy@NaYF4:Yb/Er and CaF2:Dy@NaYF4:Yb/Tm NSs, and time-dependent color evolution can be realized as well [180]. For example, in the Na3HfF7:Yb/Er NSs, the red UC intensity remained unchanged at a fixed pumping power and the bright green XEPL decreased gradually over time, which lead to the output color evolution from green to red naturally (Fig. 13a) [134]. Fluoride NSs prepared via wet-chemical method are well dispersed in a polar or nonpolar solvent, which are suitable for the design of patterns via printing and then employed for optical information storage. For example, NaYF4:(Tb, Dy or Ho)@NaYF4 with strong XEPL was printed on a glass sheet to form three quick-response codes, which was interpreted into three readable codes after passing through 545BP, 570BP, and 605LP optical filters (Fig. 13b) [153]. This route enables 3D information storage on a single layer of recording medium. Furthermore, the fluoride NSs could be used for information encryption and decryption. As shown in Fig. 13c a designed pattern with employing NaMgF3:Tb3+@NaMgF3 NSs was clearly observed under X-ray irradiation and then faded once the cessation of excitation source. After heating several seconds, the stored pattern appeared again (Fig. 13c) [181]. Considering the UC and DS fluoride nanosystems show promising application in information encryption and encoding [201, 202], the combination of XEOL, XEPL, UC and DS may greatly improve the security level.

Fig. 13
figure 13

a Schematic illustration for the realization of a time-dependent color variation display [134]. Copyright 2021 Royal Society of Chemistry. b1 Schematic illustration of the application in 3D optical information storage. b2-b3 The original images using NaYF4:Ln3+@NaYF4 and monochromatic images after passing through 545BP, 570BP and 605LP filters. The scale bars are 100 μm [153]. Copyright 2021 Springer Nature Limited. c Photographs of NaMgF3:Tb3+@NaMgF3 emission images during X-ray irradiation, at the moment of turning off, with a delay time of 10 min, and heated to 400 K [181]. Copyright 2021 Wiley–VCH

5.5 Broadband photodetection

Photodetectors have a wide range of applications in biomedical sensing [203, 204], camera imaging [205, 206], optical communications [207, 208], and night vision [209, 210]. In commercial photodetectors, crystalline inorganic semiconductors such as silicon or III-V compounds are employed as photodiodes and phototransistors, which do not effectively respond to a broad scope of photon energy covering X-ray, ultraviolet–visible (UV–vis), and NIR light [211]. To solve this issue, NaYF4:Yb/Tm (30/1 mol%)@NaYF4@mSiO2@MAPbX3 core@shell nanoparticles were developed. As shown in Fig. 14, under NIR excitation, lanthanide-doped fluoride layer emitted UV–vis light through energy-transfer UC processes, and then the subsequential radiation re-absorption process from lanthanide activators to perovskite layer occurs. Upon X-ray or UV excitation, visible emission from perovskite layer is produced through recombination of electrons in the CB and holes in the VB [212]. This nanotransducer exhibited a wide linear response to X-rays with various dose rates, as well as UV and NIR photons at different power densities. As discussed in Sect. 4.4, without integrating perovskite layer, lanthanide-doped fluoride NSs can be used for the generation of XEOL, UC and DS as well, which might be possible for the realization of broadband detection in theory and need more study in the future.

Fig. 14
figure 14

Copyright 2021 Wiley–VCH

Schematic illustration of the structure design of lanthanide-perovskite core@shell nanoparticles for detection of X-ray, UV, and NIR photons. Lanthanide doped upconversion nanocrystals are coated with a layer of mSiO2 shell and organic–inorganic lead-halide perovskite nanocrystals are in situ synthesized inside the pores of mesoporous mSiO2 [212].

6 Existing challenges and future opportunities

Lanthanide doped fluoride nanoparticles are suitable candidates for next generation NSs owing to their low bio-toxicity, high photo-/environmental- stability, facile device processability, tunable XEOL and XEPL properties, and other useful features [44, 112,113,114]. To promote the development of high performance fluoride NSs and the practical applications, we further discuss the existing challenges and future multidisciplinary opportunities in this field below.

Understanding the XEOL mechanism benefits the design and exploration of new fluoride NSs. At present, how the generated low kinetic energy charge carriers are transported to the luminescent centers or are captured by defects as well as the corresponding influence factors, are unclear. For example, it is better to calculate or characterize the energy differences among these charge carries, the first populated nonradiative excited levels and the radiative levels of lanthanide activators, which will guide the design of energy transfer processes to match the energy differences followed by the enhanced light yield. High LY is a perquisite for the realization of ultra-low dose rate applications.

The life-times of 4f-4f transitions in most trivalent lanthanide ions vary from a few µs to tens of ms [213, 214], which are not suitable for real-time dynamic XEOL imaging. Although the decay rate of Ce3+ ions is in the nanosecond range, XEOL emission in fluoride NSs doped by cerium ions is in the UV region [188, 189], which do not match well with commonly used visible detectors. So to achieve bright XEOL intensity with fast decay rate and appropriate emission wavelength range, it is better to design the local crystal environment of 5d orbitals of Ce3+ and Eu2+ ions in fluoride NSs for the realization of better XEOL performances.

Scintillators with the quantum-cutting process, which take place in some lattices, have promising properties for a number application due to high quantum efficiency (up to 190%) of resulted visible emission. An example of well-investigated material is LiGdF4 doped with Eu3+, where Gd3+ acts as an intermediator of UV photon to Eu3+, where quantum cutting occurs [215]. In addition, quantum cutting of UV photon is efficient in Tb3+ via cross-relaxation energy transfers between Tb3+ ions, leading to intense green emission [216]. Apparently, next step in this direction would be XEOL materials, where quantum cutting results to produce a pair of visible/NIR, or NIR/NIR photons, which can be realized in the lanthanide doped fluoride NSs [173]. However, it is remaining a great challenge to develop an effective method to improve its LY, owing to the big energy differences between the generated low secondary electrons and NIR excited levels. Quantum cutting processes via Pr3+/Yb3+, Tb3+/Yb3+ or Ce3+/Nd3+ co-dopants might be helpful for the increase of LY in NIR range, which is a promising research direction.

Through incorporating multiple lanthanide activators into fluoride based core@shell NSs, broadband photon detectors covering X-ray, UV–vis, and NIR light might be realized, which has not been studied yet. To inhibit deleterious cross-relaxations between different activators and achieve superior performances, the influences of activators type and their spatial distributions, the core@shell structures (i.e., composition, phase, shell thickness) and surface ligands are needed to be systematically studied.

The ability to trap X-ray photon energy in NSs for persistent radioluminescence, can be used in biomedical applications for wireless optogenetic control of neuron activities, cellular signaling pathways, and cell fate regulations. In this case, no real-time excitation would be required, thus causing less photodamage [173]. It is recognized that the XEPL is originated from the deposited electrons in defects that formduring X-rays irradiation [44, 151], however, it remains a challenge to clarify the influencing factors on the formation of defects and develop strategies to enhance the XEPL intensity and prolong its duration.

Both the fluoride host and scintillating screen structures are of significance for XEOL performance; it is thus valuable to study the structure dependent X-ray detection ability in the future. Most of the studies are mainly focused on the bulk scintillators or perovskite NSs, but scarcely in the fluoride based NSs field. Special structured scintillators with large radiation sensing area, composed of small units through meso- and macro- structure engineering [58], can exhibit extra-low level radiation sensitivity. Portable devices based on this technology could be useful for efficient remote environmental radiation monitoring.

The recent development of X-ray imaging technologies, including digital radiography, which in contrast to traditional film-screen radiography show a much wider and linear dynamic range and, therefore, reduces the risk of overexposure or underexposure [217]. This new technology inspires us to explore NSs based low-dose digital radiography in the future, and for proper processing and subsequent further analysis requires application of machine learning and artificial intelligence.

Abbreviations

XEOL:

X-ray excited optical luminescence

GAGG:Ce:

Gd3(Al,Ga)5O12:Ce

NSs:

Nanoscintillators

GSO:Ce:

Gd2SiO5:Ce

XEPL:

X-ray excited long persistent luminescence

LuAG:Ce:

Lu3Al5O12: Ce

UV:

Ultraviolet

Nph :

Number of emitted photons

NIR:

Near infrared

MTF:

Modulation transfer function

CT:

Computed tomography

E-traps:

Fluorine vacancies

LYSO:Ce:

(Lu,Y)2SiO5:Ce

H-traps:

Fluorine interstitials

BGO:

Bi4Ge3O12

EDTA:

Ethylenediaminetetraacetate

LY:

Light yield

lp:

Line pairs

NIR-II:

Second near infrared

Xr-LEI:

X-ray luminescence extension imaging

UC:

Upconversion

PDMS:

Polydimethylsiloxane

DS:

Downshifting

LSO:Ce:

Lu2SiO5:Ce

STE:

Self-trapped exciton

XPDT:

X-Ray effected photodynamic therapy

Ba[Pt(CN)4]:

Barium platino-cyanide

ROS:

Reactive oxygen species

CaS:

Calcium sulphide

CBM:

Conduction band minimum

References

  1. H. Lusic, M.W. Grinstaff, X-ray-computed tomography contrast agents. Chem. Rev. 113, 1641–1666 (2013)

    Article  Google Scholar 

  2. G. Robb, Ultra-tunable graphene light source. Nat. Photon. 10, 3–4 (2016)

    Article  ADS  Google Scholar 

  3. X. Chen, J. Song, X. Chen, H. Yang, X-ray-activated nanosystems for theranostic applications. Chem. Soc. Rev. 48, 3073–3101 (2019)

    Article  Google Scholar 

  4. L. Lu, M. Sun, Q. Lu, T. Wu, B. Huang, High energy X-ray radiation sensitive scintillating materials for medical imaging, cancer diagnosis and therapy. Nano Energy 79, 105437 (2021)

    Article  Google Scholar 

  5. G. Blasse, Scintillator materials. Chem. Mater. 1994(6), 1465–1475 (1994)

    Article  Google Scholar 

  6. A. Kamkaew, F. Chen, Y. Zhan, R.L. Majewski, W. Cai, Scintillating nanoparticles as energy mediators for enhanced photodynamic therapy. ACS Nano 10, 3918–3935 (2016)

    Article  Google Scholar 

  7. Z. Wang et al., Non-invasive classification of microcalcifications with phase-contrast X-ray mammography. Nat. Commun. 5, 3797 (2014)

    Article  ADS  Google Scholar 

  8. G. Wang et al., A deep-learning pipeline for the diagnosis and discrimination of viral, non-viral and COVID-19 pneumonia from chest X-ray images. Nat. Biomed. Eng. 5, 509–521 (2021)

    Article  Google Scholar 

  9. A. Momose, T. Takeda, Y. Itai, K. Hirano, Phase-contrast X-ray computed tomography for observing biological soft tissues. Nat. Med. 2, 473–475 (1996)

    Article  Google Scholar 

  10. O. Rabin, J.M. Perez, J. Grimm, G. Wojtkiewicz, R. Weissleder, An X-ray computed tomography imaging agent based on long-circulating bismuth sulphide nanoparticles. Nat. Mater. 5, 118–122 (2006)

    Article  ADS  Google Scholar 

  11. L.-J. Xu, X. Lin, Q. He, M. Worku, B. Ma, Highly efficient eco-friendly X-ray scintillators based on an organic manganese halide. Nat. Commun. 11, 4329 (2020)

    Article  ADS  Google Scholar 

  12. P. Li et al., 4th generation synchrotron source boosts crystalline imaging at the nanoscale. Light Sci. Appl. 11, 73 (2022)

    Article  ADS  Google Scholar 

  13. R. Chen, P. Liu, T. Xiao, L.X. Xu, X-ray imaging for non-destructive microstructure analysis at SSRF. Adv. Mater. 26, 7688–7691 (2014)

    Article  Google Scholar 

  14. M. Holler et al., High-resolution non-destructive three-dimensional imaging of integrated circuits. Nature 543, 402–406 (2017)

    Article  ADS  Google Scholar 

  15. H. Wei et al., Sensitive X-ray detectors made of methylammonium lead tribromide perovskite single crystals. Nat. Photon. 10, 333–339 (2016)

    Article  ADS  Google Scholar 

  16. S. Akcay, T. Breckon, Towards automatic threat detection: a survey of advances of deep learning within X-ray security imaging. Pattern Recogn. 122, 108245 (2022)

    Article  Google Scholar 

  17. Y.G. Zdesenko et al., Scintillation properties and radioactive contamination of CaWO4 crystal scintillators. Nucl. Instrum. Methods. Phys. Res. Sect. A. 538, 657–667 (2005)

    Article  ADS  Google Scholar 

  18. J.A. Shepherd, S.E. Sobottka, M.B. Williams, Performance and fabrication of thin film NaI(Tl) scintillators for use on imaging photomultiplier tubes. IEEE Trans. Nucl. Sci. 40, 413–416 (1993)

    Article  ADS  Google Scholar 

  19. I.G. Valais et al., Luminescence properties of (Lu, Y)2SiO5: Ce and Gd2SiO5: Ce single crystal scintillators. IEEE Trans. Nucl. Sci. 54, 11–18 (2007)

    Article  ADS  Google Scholar 

  20. B.D. Milbrath, A.J. Peurrung, M. Bliss, W.J. Weber, Radiation detector materials: an overview. J. Mater. Res. 23, 2561–2581 (2008)

    Article  ADS  Google Scholar 

  21. J.K. Chen, N. Shirahata, H.T. Sun, Metal-free scintillators excite X-ray community. Nat. Photon. 15, 171–172 (2021)

    Article  ADS  Google Scholar 

  22. Q. Chen et al., All-inorganic perovskite nanocrystal scintillators. Nature 561, 88–93 (2018)

    Article  ADS  Google Scholar 

  23. H. Chen et al., LiGa5O8:Cr-based theranostic nanoparticles for imaging-guided X-ray induced photodynamic therapy of deep-seated tumors. Mater. Horiz. 4, 1092–1101 (2017)

    Article  Google Scholar 

  24. D. Ding et al., X-ray-activated simultaneous near-infrared and short-wave infrared persistent luminescence imaging for long-term tracking of drug delivery. ACS Appl. Mater. Interfaces 13, 16166–16172 (2021)

    Article  Google Scholar 

  25. R. Yasuda, M. Kataigiri, M. Matsubayashi, Influence of powder particle size and scintillator layer thickness on the performance of Gd2O2S: Tb scintillators for neutron imaging. Nucl. Instrum. Methods. Phys. Res. Sect. A. 680, 139–144 (2012)

    Article  ADS  Google Scholar 

  26. W. Ma et al., Highly resolved and robust dynamic X-ray imaging using perovskite glass-ceramic scintillator with reduced light scattering. Adv. Sci. 8, 2003728 (2021)

    Article  Google Scholar 

  27. B. Yang et al., Lead-free halide Rb2CuBr3 as sensitive X-ray scintillator. Adv. Mater. 31, 1904711 (2019)

    Article  Google Scholar 

  28. L. Lian et al., Efficient and reabsorption-free radioluminescence in Cs3Cu2I5 nanocrystals with self-trapped excitons. Adv. Sci. 7, 2000195 (2020)

    Article  Google Scholar 

  29. T. Jiang et al., Power conversion efficiency enhancement of low-bandgap mixed Pb-Sn perovskite solar cells by improved interfacial charge transfer. ACS Energy Lett. 4, 1784–1790 (2019)

    Article  Google Scholar 

  30. Y. Gao et al., Highly stable lead-free perovskite field-effect transistors incorporating linear π-conjugated organic ligands. J. Am. Chem. Soc. 141, 15577–15585 (2019)

    Article  Google Scholar 

  31. T. Jiang et al., Realizing high efficiency over 20% of low-bandgap Pb-Sn-alloyed perovskite solar cells by in situ reduction of Sn4+. Sol. RRL 4, 1900467 (2019)

    Article  Google Scholar 

  32. D. Liu et al., Three-dimensional controlled growth of monodisperse sub-50 nm heterogeneous nanocrystals. Nat. Commun. 7, 10254 (2016)

    Article  ADS  Google Scholar 

  33. Q. Su et al., The effect of surface coating on energy migration-mediated upconversion. J. Am. Chem. Soc. 134, 20849–20857 (2012)

    Article  Google Scholar 

  34. Y. Fan et al., Lifetime-engineered NIR-II nanoparticles unlock multiplexed in vivo imaging. Nat. Nanotech. 13, 941–946 (2018)

    Article  ADS  Google Scholar 

  35. Y. Zhong et al., In vivo molecular imaging for immunotherapy using ultra-bright near-infrared-IIb rare-earth nanoparticles. Nat. Biotechnol. 37, 1322–1331 (2019)

    Article  Google Scholar 

  36. D.K. Chatterjee, A.J. Rufaihah, Y. Zhang, Upconversion fluorescence imaging of cells and small animals using lanthanide doped nanocrystals. Biomaterials 29, 937–943 (2008)

    Article  Google Scholar 

  37. J. Zhou, Z. Liu, F. Li, Upconversion nanophosphors for small-animal imaging. Chem. Soc. Rev. 41, 1323–1349 (2012)

    Article  Google Scholar 

  38. M. Tan et al., Rare-earth-doped fluoride nanoparticles with engineered long luminescence lifetime for time-gated in vivo optical imaging in the second biological window. Nanoscale 10, 17771–17780 (2018)

    Article  Google Scholar 

  39. F. Wang et al., Tuning upconversion through energy migration in core-shell nanoparticles. Nat. Mater. 10, 968–973 (2011)

    Article  ADS  Google Scholar 

  40. D. Wang et al., ICG-sensitized NaYF4: Er nanostructure for theranostics. Adv. Opt. Mater. 6, 1701142 (2018)

    Article  Google Scholar 

  41. Y. Li, S. Zeng, J. Hao, Non-invasive optical guided tumor metastasis/vessel imaging by using lanthanide nanoprobe with enhanced down-shifting emission beyond 1500 nm. ACS Nano 13, 248–259 (2019)

    Article  Google Scholar 

  42. Z. Chen et al., Low dose of X-ray-excited long-lasting luminescent concave nanocubes in highly passive targeting deep-seated hepatic tumors. Adv. Mater. 31, 1905087 (2019)

    Article  Google Scholar 

  43. Y. Zhuang, L. Wang, Y. Lv, T. Zhou, R. Xie, Optical data storage and multicolor emission readout on flexible films using deep-trap persistent luminescence materials. Adv. Funct. Mater. 28, 1705769 (2018)

    Article  Google Scholar 

  44. X. Ou et al., High-resolution X-ray luminescence extension imaging. Nature 590, 410–415 (2021)

    Article  ADS  Google Scholar 

  45. P. Thibault, M. Dierolf, A. Menzel, O. Bunk, C. David, F. Pfeiffer, High-resolution scanning X-ray diffraction microscopy. Science 321, 379–382 (2008)

    Article  ADS  Google Scholar 

  46. M. Dierolf et al., Ptychographic X-ray computed tomography at the nanoscale. Nature 467, 436–439 (2010)

    Article  ADS  Google Scholar 

  47. W.C. Rontgen, On a new kind of rays. Nature 53, 274–276 (1895)

    Google Scholar 

  48. M. Suga et al., Native structure of photosystem II at 1.95 Å resolution viewed by femtosecond X-ray pulses. Nature 517, 99–103 (2015)

    Article  ADS  Google Scholar 

  49. C. Leveille et al., Ultrafast time-evolution of chiral Neel magnetic domain walls probed by circular dichroism in x-ray resonant magnetic scattering. Nat. Commun. 13, 1412 (2022)

    Article  ADS  Google Scholar 

  50. S. Yakunin et al., Detection of X-ray photons by solution-processed lead halide perovskites. Nat. Photon. 9, 444–449 (2015)

    Article  ADS  Google Scholar 

  51. J. Zhao et al., Perovskite-filled membranes for flexible and large-area direct-conversion X-ray detector arrays. Nat. Photon. 14, 612–617 (2020)

    Article  ADS  Google Scholar 

  52. M. Spahn, X-ray detectors in medical imaging. Nucl. Instrum. Methods Phys. Res. Sect. A. 731, 57–63 (2013)

    Article  ADS  Google Scholar 

  53. C. D’Ambrosio, F. De Notaristefani, H. Leutz, D. Puertols, E. Rosso, X-ray detection with a scintillating YAP-window hybrid photomultiplier tube. IEEE Trans. Nucl. Sci. 47, 6–12 (2000)

    Article  ADS  Google Scholar 

  54. C. Carrier, R. Lecomte, Recent results in scintillation detection with silicon avalanche photodiodes. IEEE Trans. Nucl. Sci. 37, 209–214 (1990)

    Article  ADS  Google Scholar 

  55. H. Wei, J. Huang, Halide lead perovskites for ionizing radiation detection. Nat. Commun. 10, 1066 (2019)

    Article  ADS  Google Scholar 

  56. B.K. Cha et al., Investigation of the performance of scintillator-based CMOS flat panel detectors for X-ray and thermal neutron imaging. IEEE Trans. Nucl. Sci. 57, 1409–1413 (2010)

    Article  ADS  Google Scholar 

  57. F. Cao et al., Shining emitter in stable host: design halide perovskite scintillators for X-ray imaging from commercial concept. ACS Nano 14, 5183–5193 (2020)

    Article  Google Scholar 

  58. Z. Lin, S. Lv, Z. Yang, J. Qiu, S. Zhou, Structured scintillators for efficient radiation detection. Adv. Sci. 9, 2102439 (2022)

    Article  Google Scholar 

  59. J.B. Birks, Scintillations from organic crystals: specific fluorescence and relative response to different radiations. Proc. Phys. Soc. A 64, 874 (1951)

    Article  ADS  Google Scholar 

  60. F. Maddalena et al., Inorganic, organic, and perovskite halides with nanotechnology for high-light yield X- and γ-ray scintillators. Curr. Comput.-Aided Drug Des. 9, 88 (2019)

    Google Scholar 

  61. M.J. Weber, Inorganic scintillators: today and tomorrow. J. Lumin. 100, 35–45 (2002)

    Article  Google Scholar 

  62. Y. Wang, X. Yin, J. Chen, Y. Wang, Z. Chai, S. Wang, Gleaming uranium: an emerging emitter for building X-ray scintillators. Chem.-Eur. J. 26, 1900–1905 (2020)

    Article  Google Scholar 

  63. A. Chaudhry, R. Boutchko, S. Chourou, G. Zhang, N. Gronbech-Jensen, A. Canning, First-principles study of luminescence in Eu2+-doped inorganic scintillators. Phys. Rev. B 89, 155105 (2014)

    Article  ADS  Google Scholar 

  64. C. Dujardin et al., Needs, trends, and advances in inorganic scintillators. IEEE Trans. Nucl. Sci. 65, 1977–1997 (2018)

    Article  ADS  Google Scholar 

  65. T. Yanagida et al., Optical and scintillation properties of bulk ZnO crystal. Phys. Status Solidi C 9, 2284–2287 (2012)

    Article  ADS  Google Scholar 

  66. T. Yanagida, Y. Fujimoto, M. Koshimizu, Evaluation of scintillation properties of GaN. Surf. Sci. Nanotechnol. 12, 396–399 (2014)

    Article  Google Scholar 

  67. V.B. Mikhailik, H. Kraus, J. Imber, D. Wahl, Scintillation properties of pure CaF2. Nucl. Instrum. Methods Phys. Res. A 566, 522–525 (2006)

    Article  ADS  Google Scholar 

  68. R.Y. Shendrik, A. Radzhabov, A.I. Nepomnyashchikh, Scintillation properties of SrF2 and SrF2-Ce3+ crystals. Tech. Phys. Lett. 39, 587–590 (2013)

    Article  ADS  Google Scholar 

  69. M. Laval et al., Barium fluoride-inorganic scintillator for subnanosecond timing. Nucl. Instrum. Meth. Phys. Res. 206, 169–176 (1983)

    Article  ADS  Google Scholar 

  70. T. Yanagida et al., Growth and scintillation properties of BaMgF4. Nucl. Instrum. Methods Phys. Res. A 621, 473–477 (2010)

    Article  ADS  Google Scholar 

  71. M. Moszynski, R. Allemand, M.R. Odru, J. Vacher, Laval Recent progress in fast timing with CsF scintillators in application to time-of-flight positron tomography in medicine. Nucl. Instrum. Methods Phys. Res. Sect. A 205, 239–249 (1983)

    Article  Google Scholar 

  72. Y. Tokuda, K. Sakaguchi, T. Nishihara, K. Takano, T. Fukushima, M. Hangyo, X-ray detection capability of a Cs2ZnCl4 single- crystal scintillator. Appl. Phys. Exp. 7, 062602–1–062602–4 (2014)

  73. M.J. Weber, R.R. Monchamp, Luminescence of Bi4Ge3O12: spectral and decay properties. J. Appl. Phys. 44, 5495–5499 (1973)

    Article  ADS  Google Scholar 

  74. D.F. Anderson, Properties of the high-density scintillator cerium fluoride. IEEE Trans. Nucl. Sci. 36, 137–140 (1989)

    Article  ADS  Google Scholar 

  75. E. Garcia-Torano, B. Caro, V. Peyres, M. Mejuto, Characterization of a CeBr3 detector and application to the measurement of some materials from steelworks. Nucl. Instrum. Methods Phys. Res. A 837, 63–68 (2016)

    Article  ADS  Google Scholar 

  76. Y. Fujimoto et al., Thallium magnesium chloride: a high light yield, large effective atomic number, intrinsically activated crystalline scintillator for X-ray and gamma-ray detection. Jpn. J. Appl. Phys. 55, 090301–1–090301–3 (2016).

  77. H. Masai, T. Yanagida, Y. Fujimoto, M. Koshimizu, T. Yoko, Scintillation property of rare earth-free SnO-doped oxide glass. Appl. Phys. Lett. 101, 191906 (2012)

    Article  ADS  Google Scholar 

  78. T. Kato, G. Okada, T. Yanagida, Optical, scintillation and dosimeter properties of MgO transparent ceramic doped with Mn2+. J. Ceram. Soc. Jpn. 124, 559–563 (2016)

    Article  Google Scholar 

  79. T. Kato, G. Okada, T. Yanagida, Optical, scintillation and dosimeter properties of MgO translucent ceramic doped with Cr3+. Opt. Mater. 54, 134–138 (2016)

    Article  ADS  Google Scholar 

  80. Y. Futami, T. Yanagida, Y. Fujimoto, Optical, dosimetric, and scintillation properties of pure sapphire crystals. Jpn. J. Appl. Phys. 53, 02BC12 (2014)

    Article  Google Scholar 

  81. T. Kato, G. Okada, T. Yanagida, Optical, scintillation and dosimeter properties of MgO transparent ceramic and single crystal. Ceram. Int. 42, 5617–5622 (2016)

    Article  Google Scholar 

  82. T.A. Edison, Communication to Lord Kelvin. Nature 53, 470 (1896)

    Google Scholar 

  83. W. Crookes, The emanations of radium. Proc. R. Soc. Lond. 71, 405–408 (1903)

    Article  Google Scholar 

  84. E. Sakai, Recent measurements on scintillator-photodetector systems. IEEE Trans. Nucl. Sci. 34, 418–422 (1987)

    Article  ADS  Google Scholar 

  85. I. Holl, E. Lorenz, G. Mageras, A measurement of the light yield of common inorganic scintillators. IEEE Trans. Nucl. Sci. 35, 105–109 (1988)

    Article  ADS  Google Scholar 

  86. Y. Shimizu, M. Minowa, W. Suganuma, Y. Inoue, Dark matter search experiment with CaF2(Eu) scintillator at Kamioka Observatory. Phys. Lett. B 633, 195–200 (2006)

    Article  ADS  Google Scholar 

  87. B.C. Grabmaier, Crystal scintillators. IEEE Trans. Nucl. Sci. 31, 372–376 (1984)

    Article  ADS  Google Scholar 

  88. T. Yanagida, Inorganic scintillating materials and scintillation detectors. Proc. Jpn. Acad. Ser. B 94, 75–97 (2018)

    Article  ADS  Google Scholar 

  89. M.J. Weber, R.R. Monchamp, Luminescence of Bi4Ge3O12-spectral and decay properties. J. Appl. Phys. 44, 5495–5499 (1973)

    Article  ADS  Google Scholar 

  90. C.L. Melcher, J.S. Schweitzer, Cerium-doped lutetium oxyorthosilicate-a fast, efficient new scintillator. IEEE Trans. Nucl. Sci. 39, 502–505 (1992)

    Article  ADS  Google Scholar 

  91. L. Pidol et al., High efficiency of lutetium silicate scintillators, Ce-doped LPS, and LYSO crystals. IEEE Trans. Nucl. Sci. 51, 1084–1087 (2004)

    Article  ADS  Google Scholar 

  92. K. Kamada et al., Composition engineering in cerium-doped (Lu, Gd)3(Ga, Al)5O12 single-crystal scintillators. Cryst. Growth Des. 11, 4484–4490 (2011)

    Article  Google Scholar 

  93. R.T. Williams, W.W. Wolszczak, X. Yan, D.L. Carroll, Perovskite quantum-dot-in-host for detection of ionizing radiation. ACS Nano 14, 5161–5169 (2020)

    Article  Google Scholar 

  94. R.-Y. Zhu, C.L. Woody, Inorganic scintillators. Phys. Lett. B 667, 286 (2008)

    Google Scholar 

  95. H. Nakamura, H. Kitamura, R. Hazama, Development of a new rectangular NaI(Tl) scintillator and spectroscopy of low-energy charged particles. Rev. Sci. Instrum. 81, 013104 (2010)

    Article  ADS  Google Scholar 

  96. P. Schotanus, P. Dorenbos, C.W.E. van Eijk, H.F. Lamfers, Suppression of the slow scintillation light output of BaF2 crystals by La3+ doping. Nucl. Instrum. Methods Phys. Res. A 281, 162–166 (1989)

    Article  ADS  Google Scholar 

  97. R. Shendrik, E. Radzhabov, Absolute light yield measurements on SrF2 and BaF2 doped with rare earth ions. IEEE Trans. Nucl. Sci. 61, 406–410 (2014)

    Article  ADS  Google Scholar 

  98. T. Seo, New method for the centroid-shift analysis of picosecond lifetime measurements using BaF2 scintillators. Nucl. Instrum. Methods Phys. Res. A 325, 176–186 (1993)

    Article  ADS  Google Scholar 

  99. K. Takagi, T. Fukazawa, Cerium activated Gd2SiO5 single crystal scintillator. Appl. Phys. Lett. 42, 43 (1983)

    Article  ADS  Google Scholar 

  100. V.V. Avdeichikov et al., Light output and energy resolution of CsI, YAG, GSO, BGO and LSO scintillators for light ions. Nucl. Instrum. Methods Phys. Res. A 349, 216–224 (1994)

    Article  ADS  Google Scholar 

  101. Z. Ma, X. Bi, X. Liu, D. Li, J. Li, X. Sun, Preliminary exploration on the preparation of LYSO: Ce single crystal using verneuil method. Mater. Sci. Forum 1003, 247–253 (2020)

    Article  Google Scholar 

  102. J. Xu, L. Fan, Y. Shi, J. Li, J. Xie, F. Lei, Effects of Ce3+ doping concentrations on microstructure and luminescent properties of Ce3+: Lu3Al5O12 (Ce: LuAG) transparent ceramics. Opt. Mater. 36, 1954–1958 (2014)

    Article  ADS  Google Scholar 

  103. M. Nikl, Scintillation detectors for X-rays. Meas. Sci. Technol. 17, R37–R54 (2006)

    Article  ADS  Google Scholar 

  104. M.V.S. Rezende, P.J.R. Montes, A.B. Andrade, Z.S. Macedo, M.E.G. Valerio, Mechanism of X-ray excited optical luminescence (XEOL) in europium doped BaAl2O4 phosphor. Phys. Chem. Chem. Phys. 18, 17646–17654 (2016)

    Article  Google Scholar 

  105. S.J. Duclos et al., Development of the HiLightTM scintillator for computed tomography medical imaging. Nucl. Instrum. Methods Phys. Res. A 505, 68–71 (2003)

    Article  ADS  Google Scholar 

  106. G.B. Schober, J.N. Anker, Radioluminescence imaging of drug elution from biomedical implants. Adv. Funct. Mater. 32, 2106508 (2022)

    Article  Google Scholar 

  107. V. Kumar, Z. Luo, A review on x-ray excited emission decay dynamics in inorganic scintillator materials. Photonics 8, 71 (2021)

    Article  Google Scholar 

  108. J. Ma et al., Highly efficient NaGdF4: Ce/Tb nanoscintillator with reduced afterglow and light scattering for high-resolution X-ray imaging. ACS Appl. Mater. Interfaces 13, 44596–44603 (2021)

    Article  Google Scholar 

  109. Y. Zhou, J. Chen, O.M. Bakr, O.F. Mohammed, Metal halide perovskites for X-ray imaging scintillators and detectors. ACS Energy Lett. 6, 739–768 (2021)

    Article  Google Scholar 

  110. J.-C.G. Bunzli, C. Piguet, Taking advantage of luminescent lanthanide ions. Chem. Soc. Rev. 34, 1048–1077 (2005)

    Article  Google Scholar 

  111. X. Zhu et al., Temperature-feedback upconversion nanocomposite for accurate photothermal therapy at facile temperature. Nat. Commun. 7, 10437 (2016)

    Article  ADS  Google Scholar 

  112. F. Ahmad et al., Codoping enhanced radioluminescence of nanoscintillators for X-ray-activated synergistic cancer therapy and prognosis using metabolomics. ACS Nano 13, 10419–10433 (2019)

    Article  Google Scholar 

  113. N.M. Idris, M.K. Gnanasammandhan, J. Zhang, P.C. Ho, R. Mahendran, Y. Zhang, In vivo photodynamic therapy using upconversion nanoparticles as remote-controlled nanotransducers. Nat. Med. 18, 1580–1585 (2012)

    Article  Google Scholar 

  114. J. Wang et al., Enhancing multiphoton upconversion through energy clustering at sublattice level. Nat. Mater. 13, 157–162 (2013)

    Article  ADS  Google Scholar 

  115. K. Lingeshwar Reddy, R. Balaji, A. Kumar, V. Krishnan, Lanthanide doped near infrared active upconversion nanophosphors: fundamental concepts, synthesis strategies, and technological applications. Small 14, 1801304 (2018)

    Article  Google Scholar 

  116. D.R. Cooper, J.A. Capobianco, J. Seuntjens, Radioluminescence studies of colloidal oleate-capped β-Na (Gd, Lu)F4: Ln3+ nanoparticles (Ln= Ce, Eu, Tb). Nanoscale 10, 7821–7832 (2018)

    Article  Google Scholar 

  117. V. Yakovyna, Y. Zhydachevskii, V.B. Mikhailik, I. Solskiic, D. Sugakc, M. Vakivc, Effect of thermo-chemical treatments on the luminescence and scintillation properties of CaWO4. Opt. Mater. 30, 1630–1634 (2008)

    Article  ADS  Google Scholar 

  118. A.A. Demidenko et al., Scintillation parameters of BaF2 and BaF2: Ce3+ ceramics. Opt. Mater. 32, 1291–1293 (2010)

    Article  ADS  Google Scholar 

  119. X. Wang et al., Photo/cathodoluminescence and stability of Gd2O2S: Tb, Pr green phosphor hexagons calcined from layered hydroxide sulfate. J. Am. Ceram. Soc. 101, 5477–5486 (2018)

    Article  Google Scholar 

  120. M. Yoneyama et al., Evaluation of GAGG: Ce scintillators for future space applications. J. Instrum. 13, P02023 (2018)

    Article  Google Scholar 

  121. R. Ma, L. Zhang, R. Zhu, Optical and scintillation properties of inorganic scintillators in high energy physics. IEEE Trans. Nucl. Sci. 55, 2425–2431 (2008)

    Article  ADS  Google Scholar 

  122. X. Qin, X. Liu, W. Huang, M. Bettinelli, X. Liu, Lanthanide-activated phosphors based on 4f–5d optical transitions: theoretical and experimental aspects. Chem. Rev. 117, 4488–4527 (2017)

    Article  Google Scholar 

  123. W.M. Higgins, A. Churilov, E. van Loef, J. Glodo, M. Squillante, K. Shah, Crystal growth of large diameter LaBr3: Ce and CeBr3. J. Cryst. Growth 310, 2085–2089 (2008)

    Article  ADS  Google Scholar 

  124. C.L. Melcher, Scintillators for well logging applications. Nucl. Instrum. Methods Phys. Res. A 40, 1214–1218 (1989)

    Article  Google Scholar 

  125. E. Sysoeva, V. Tarasov, O. Zelenskaya, Comparison of the methods for determination of scintillation light yield. NNucl. Instrum. Methods Phys. Res. A 486, 67–73 (2002)

    Article  ADS  Google Scholar 

  126. M. Nikl, A. Yoshikawa, Recent R&D trends in inorganic single-crystal scintillator materials for radiation detection. Adv. Opt. Mater. 3, 463–481 (2015)

    Article  Google Scholar 

  127. F. Zhou, Z. Li, W. Lan, Q. Wang, L. Ding, Z. Jin, Halide perovskite, a potential scintillator for X-Ray detection. Small Methods 4, 2000506 (2020)

    Article  Google Scholar 

  128. W. Huda, R.B. Abrahams, X-ray-based medical imaging and resolution. Am. J. Roentgenol. 204, W393–W397 (2015)

    Article  Google Scholar 

  129. J. Zhou et al., Impact of lanthanide nanomaterials on photonic devices and smart applications. Small 14, 1801882 (2018)

    Article  Google Scholar 

  130. L. Lei et al., Amplifying upconversion by engineering interfacial density of state in sub-10 nm colloidal core/shell fluoride nanoparticles. Nano Lett. 24, 10222–10229 (2021)

    Article  Google Scholar 

  131. F. Wang, R. Deng, X. Liu, Preparation of core-shell NaGdF4 nanoparticles doped with luminescent lanthanide ions to be used as upconversion-based probes. Nat. Protoc. 9, 1634–1644 (2014)

    Article  Google Scholar 

  132. G. Chen et al., Core/shell NaGdF4:Nd3+/NaGdF4 nanocrystals with efficient near-infrared to near-infrared downconversion photoluminescence for bioimaging applications. ACS Nano 6, 2969–2977 (2012)

    Article  Google Scholar 

  133. L. Song et al., Low-dose X-ray activation of W(VI)-doped persistent luminescence nanoparticles for deep-tissue photodynamic therapy. Adv. Funct. Mater. 28, 1707496 (2018)

    Article  ADS  Google Scholar 

  134. W. Xu, L. Lei, Y. Wang, E. Liu, L. Chen, S. Xu, Modulating electron population pathways for time-dependent dynamic multicolour displays. Mater. Horiz. 8, 3443–3448 (2021)

    Article  Google Scholar 

  135. W. Fan et al., Breaking the depth dependence by nanotechnology-enhanced X-ray-excited deep cancer theranostics. Adv. Mater. 31, 1806381 (2019)

    Article  Google Scholar 

  136. F. Wang, X. Liu, Multicolor tuning of lanthanide-doped nanoparticles by single wavelength excitation. Accounts Chem. Res. 47, 1378–1385 (2014)

    Article  Google Scholar 

  137. Y. Wu, W. Wu, Combinations of superior inorganic phosphors for level-tunable information hiding and encoding. Adv. Opt. Mater. 9, 2100281 (2021)

    Article  Google Scholar 

  138. B. Liu, C. Li, P. Yang, Z. Hou, J. Lin, 808-nm-Light-excited lanthanide-doped nanoparticles: rational design, luminescence control and theranostic applications. Adv. Mater. 29, 1605434 (2017)

    Article  Google Scholar 

  139. G. Gao et al., Up-conversion fluorescent labels for plastic recycling: a review. Adv. Sustain. Syst. 1, 1600033 (2017)

    Article  Google Scholar 

  140. P.M. Kraus, M. Zurch, S.K. Cushing, D.M. Neumark, S.R. Leone, The ultrafast X-ray spectroscopic revolution in chemical dynamics. Nat. Rev. Chem. 2, 82–94 (2018)

    Article  Google Scholar 

  141. E.F. Garman, Developments in x-ray crystallographic structure determination of biological macromolecules. Science 343, 1102–1108 (2014)

    Article  ADS  Google Scholar 

  142. C.M. Jeffries et al., Small-angle X-ray and neutron scattering. Nat. Rev. Methods Primers 1, 70 (2021)

    Article  Google Scholar 

  143. J. Scharf et al., Bridging nano- and microscale X-ray tomography for battery research by leveraging artificial intelligence. Nat. Nanotechnol. (2022). https://doi.org/10.1038/s41565-022-01081-9

    Article  Google Scholar 

  144. A. Einstein, Uber einen Erzeugung und Verwandlung des lichtes betreffenden heuristischen Gesichtspunkt. Ann. Phys. 17, 132–148 (1905)

    Article  MATH  Google Scholar 

  145. A.H. Compton, A quantum theory of the scattering of X-rays by light elements. Phys. Rev. 21, 483 (1923)

    Article  ADS  Google Scholar 

  146. J.J. Thomson, On electrical oscillations and the effects produced by the motion of an electrified sphere. Proc. Lond. Math. Soc. 1, 197–219 (1883)

    Article  MathSciNet  MATH  Google Scholar 

  147. G. Matt, M. Feroci, M. Rapisarda, E. Costa, Treatment of Compton scattering of linearly polarized photons in Monte Carlo codes. Radiat. Phys. Chem. 48, 403–411 (1996)

    Article  ADS  Google Scholar 

  148. M. Rizzi, M. D’Aloia, B. Castagnolo, Semiconductor detectors and principles of radiation-matter interaction. J. Appl. Sci. 10, 3141–3155 (2010)

    Article  ADS  Google Scholar 

  149. H.J. Ache, Chemistry of the positron and of positronium. Angew. Chem. Int. Edit. 11, 179–199 (1972)

    Article  Google Scholar 

  150. T. Enoto et al., Photonuclear reactions triggered by lightning discharge. Nature 551, 481–484 (2017)

    Article  ADS  Google Scholar 

  151. L. Li et al., Mechanism of the trivalent lanthanides’ persistent luminescence in wide bandgap materials. Light Sci. Appl. 11, 1–8 (2022)

    Google Scholar 

  152. X. Zhong et al., NaCeF4: Gd, Tb scintillator as an X-ray responsive photosensitizer for multimodal imaging-guided synchronous radio/radiodynamic therapy. Nano Lett. 19, 8234–8244 (2019)

    Article  ADS  Google Scholar 

  153. Y. Zhuang et al., X-ray-charged bright persistent luminescence in NaYF4: Ln3+@ NaYF4 nanoparticles for multidimensional optical information storage. Light Sci. Appl. 10, 1–10 (2021)

    Article  Google Scholar 

  154. T. Passuello et al., Structural and optical properties of Vernier phase lutetium oxyfluorides doped with lanthanide ions: interesting candidates as scintillators and X-ray phosphors. J. Mater. Chem. 22, 10639–10649 (2012)

    Article  Google Scholar 

  155. P. Dorenbos, Relation between Eu2+ and Ce3+ f↔d-transition energies in inorganic compounds. J. Phys. Condens. Matter 15, 4797–4807 (2003)

    Article  ADS  Google Scholar 

  156. G. Li, Y. Tian, Y. Zhao, J. Lin, Recent progress in luminescence tuning of Ce3+ and Eu2+-activated phosphors for pc-WLEDs. Chem. Soc. Rev. 44, 8688–8713 (2015)

    Article  Google Scholar 

  157. L. Lei et al., Modulation of surface energy transfer cascade for reversible photoluminescence pH sensing. Chem. Mater. 31, 8121–8128 (2019)

    Article  Google Scholar 

  158. J. Liu et al., Simultaneously excited downshifting/upconversion luminescence from lanthanide-doped core/shell fluoride nanoparticles for multimode anticounterfeiting. Adv. Funct. Mater. 28, 1707365 (2018)

    Article  Google Scholar 

  159. H. Huang et al., Lanthanide-doped core@ multishell nanoarchitectures: multimodal excitable upconverting/downshifting luminescence and high-level anti-counterfeiting. Small 16, 2000708 (2020)

    Article  Google Scholar 

  160. Z. Qiu, S. Wang, W. Wang, S. Wu, Polymer composites entrapped Ce-doped LiYF4 microcrystals for high-sensitivity X-ray scintillation and omaging. ACS Appl. Mater. Interfaces 12, 29835–29843 (2020)

    Google Scholar 

  161. Y.-M. Yang et al., X-ray-activated long persistent phosphors featuring strong UVC afterglow emissions. Light Sci. Appl. 7, 1–11 (2018)

    Article  Google Scholar 

  162. K. Fukuda, T. Yanagida, Y. Fujimoto, N. Kawaguchi, H. Sekiya, Study of vacuum ultraviolet scintillation in rare earth doped Lutetium fluoride. 2013 IEEE NSS/MIC 1–3 (2013).

  163. J. Pejchal et al., Crystal growth and scintillation properties of selected fluoride crystals for VUV scintillators. J. Cryst. Growth 401, 833–838 (2014)

    Article  ADS  Google Scholar 

  164. O.L. Malta, L.D. Carlos, Intensities of 4f–4f transitions in glass materials. Quim. Nova 26, 889–895 (2003)

    Article  Google Scholar 

  165. K. Li, X. Liu, Y. Zhang, X. Li, H. Lian, J. Lin, Host-sensitized luminescence properties in CaNb2O6: Ln3+ (Ln3+= Eu3+/Tb3+/Dy3+/Sm3+) phosphors with abundant colors. Inorg. Chem. 54, 323–333 (2015)

    Article  Google Scholar 

  166. C.L. Hsu, S.J. Chang, Doped ZnO 1D nanostructures: synthesis, properties, and photodetector application. Small 10, 4562–4585 (2014)

    Article  Google Scholar 

  167. S. Zhang et al., Highly sensitive InSb nanosheets infrared photodetector passivated by ferroelectric polymer. Adv. Funct. Mater. 30, 2006156 (2020)

    Article  Google Scholar 

  168. L. Tu, X. Liu, F. Wu, H. Zhang, Excitation energy migration dynamics in upconversion nanomaterials. Chem. Soc. Rev. 44, 1331–1345 (2015)

    Article  Google Scholar 

  169. R. Deng, F. Qin, R. Chen, M. Hong, X. Liu, Temporal full-colour tuning through non-steady-state upconversion. Nat. Nanotech. 10, 237–242 (2015)

    Article  ADS  Google Scholar 

  170. X. Xu et al., Optimising passivation shell thickness of single upconversion nanoparticles using a time-resolved spectrometer. APL Photon. 4, 026104 (2019)

    Article  ADS  Google Scholar 

  171. L. Liu, F. Qin, H. Zhao, T. Lv, Z. Zhang, W. Cao, Shell thickness dependence of upconversion luminescence of β-NaYF4: Yb, Er/β-NaYF4 core-shell nanocrystals. Opt. Lett. 38, 2101–2103 (2013)

    Article  ADS  Google Scholar 

  172. L.G. Jacobsohn et al., Synthesis, luminescence and scintillation of rare earth doped lanthanum fluoride nanoparticles. Opt. Mater. 33, 136–140 (2010)

    Article  ADS  Google Scholar 

  173. P. Pei et al., X-ray-activated persistent luminescence nanomaterials for NIR-II imaging. Nat. Nanotechnol. 16, 1011–1018 (2021)

    Article  ADS  Google Scholar 

  174. C. Li et al., In vivo real-time visualization of tissue blood flow and angiogenesis using Ag2S quantum dots in the NIR-II window. Biomaterials 35, 393–400 (2014)

    Article  Google Scholar 

  175. G. Hong et al., Ultrafast fluorescence imaging in vivo with conjugated polymer fluorophores in the second near-infrared window. Nat. Commun. 5, 4206 (2014)

    Article  ADS  Google Scholar 

  176. J. Yan, B. Li, P. Yang, J. Lin, Y. Dai, Progress in light-responsive lanthanide nanoparticles toward deep tumor theranostics. Adv. Funct. Mater. 31, 2104325 (2021)

    Article  Google Scholar 

  177. Y. Liu et al., Significantly enhanced afterglow brightness via intramolecular energy transfer. ACS Materials Lett. 3, 713–720 (2021)

    Article  Google Scholar 

  178. Y. Li, M. Gecevicius, J. Qiu, Long persistent phosphors-from fundamentals to applications. Chem. Soc. Rev. 45, 2090–2136 (2016)

    Article  Google Scholar 

  179. H. Singh, S.K. Bhardwaj, M. Khatri, K.-H. Kim, N. Bhardwaja, UVC radiation for food safety: an emerging technology for the microbial disinfection of food products. Chem. Eng. J. 417, 128084 (2021)

    Article  Google Scholar 

  180. K. Huang et al., Three-dimensional colloidal controlled growth of core-shell heterostructured persistent luminescence nanocrystals. Nano Lett. 21, 4903–4910 (2021)

    Article  ADS  Google Scholar 

  181. Y. Wang et al., NaMgF3:Tb3+@NaMgF3 nanoparticles containing deep traps for optical information storage. Adv. Opt. Mater. 9, 2100624 (2021)

    Article  Google Scholar 

  182. Z. Li, H. Li, H. Sun, X-ray-activated UVA long persistent luminescence from defective fluoride elpasolites. J. Rare Earths 38, 124–129 (2020)

    Article  Google Scholar 

  183. Y. Hu et al., X-ray-excited super-long green persistent luminescence from Tb3+ monodoped β-NaYF4. J. Phys. Chem. C 124, 24940–24948 (2020)

    Article  Google Scholar 

  184. G. Chen, C. Yang, P.N. Prasad, Nanophotonics and nanochemistry: Controlling the excitation dynamics for frequency up- and down-conversion in lanthanide-doped nanoparticles. Accounts Chem. Res. 46, 1474–1486 (2013)

    Article  Google Scholar 

  185. K. Sankaranarayanan, J.S. Wassom, Reflections on the impact of advances in the assessment of genetic risks of exposure to ionizing radiation on international radiation protection recommendations between the mid-1950s and the present. Mutation Res. 658, 1–27 (2008)

    Article  Google Scholar 

  186. K. Masaoka, T. Yamashita, Y. Nishida, M. Sugawara, Modified slanted-edge method and multidirectional modulation transfer function estimation. Opt. Express 22, 6040–6046 (2014)

    Article  ADS  Google Scholar 

  187. F. Viallefont-Robinet et al., Comparison of MTF measurements using edge method: towards reference data set. Opt. Express 26, 33625–33648 (2018)

    Article  ADS  Google Scholar 

  188. H. Lu, X. Xu, G. Feng, B. Sun, S. Wang, S. Wu, Terbium doped LiLuF4 nanocrystal scintillator-based flexible composite film for high resolution X-ray imaging. RSC Adv. 12, 4615–4623 (2022)

    Article  ADS  Google Scholar 

  189. S. Clement, W. Deng, E. Camilleri, B.C. Wilson, E.M. Goldys, X-ray induced singlet oxygen generation by nanoparticle-photosensitizer conjugates for photodynamic therapy: Determination of singlet oxygen quantum yield. Sci. Rep. 6, 19954 (2016)

    Article  ADS  Google Scholar 

  190. T. Rimoldi et al., CeF3-ZnO scintillating nanocomposite for self-lighted photodynamic therapy of cancer. J. Mater. Sci. Mater. Med 27, 159 (2016)

    Article  Google Scholar 

  191. D.R. Cooper, K. Kudinov, P. Tyagi, C.K. Hill, S.E. Bradforth, J.L. Nadeau, Photoluminescence of cerium fluoride and cerium-doped lanthanum fluoride nanoparticles and investigation of energy transfer to photosensitizer molecules. Phys. Chem. Chem. Phys. 16, 12441–12453 (2014)

    Article  Google Scholar 

  192. M.Y. Losytskyy, L.V. Kuzmenko, O.B. Shcherbakov, N.F. Gamaleia, A.I. Marynin, V.M. Yashchuk, Energy transfer in Ce0.85Tb0.15F3 nanoparticles-CTAB shell-Chlorin e6 system. Nanoscale Res. Lett. 12, 294 (2017)

    Article  ADS  Google Scholar 

  193. Y. Tang, J. Hu, A. Elmenoufy, X. Yang, Highly efficient FRET system capable of deep photodynamic therapy established on X-ray excited mesoporous LaF3: Tb scintillating nanoparticles. ACS Appl. Mater. Interfaces 7, 12261–12269 (2015)

    Article  Google Scholar 

  194. M. Jiang, Z. Xue, Y. Li, H. Liu, S. Zeng, J. Hao, A soft X-ray activated lanthanide scintillator for controllable NO release and gas-sensitized cancer therapy. Nanoscale Horiz. 5, 268–273 (2020)

    Article  ADS  Google Scholar 

  195. J. Zhou, X. Zhu, M. Chen, Y. Sun, F. Li, Water-stable NaLuF4-based upconversion nanophosphors with long-term validity for multimodal lymphatic imaging. Biomaterials 33, 6201–6210 (2012)

    Article  Google Scholar 

  196. Y. Liu, K. Ai, J. Liu, Q. Yuan, Y. He, L. Lu, A high-performance ytterbium-based nanoparticulate contrast agent for in vivo X-ray computed tomography imaging. Angew. Chem. Int. Ed. 124, 1466–1471 (2012)

    Article  ADS  Google Scholar 

  197. L. Xu et al., Construction of high quality ultrathin lanthanide oxyiodide nanosheets for enhanced CT imaging and anticancer drug delivery to efficient cancer theranostics. Biomaterials 230, 119670 (2020)

    Article  Google Scholar 

  198. C. Ronda, H. Wieczorek, V. Khanin, P. Rodnyi, Review-Scintillators for medical imaging: a tutorial overview. ECS J. Sol. State Sci. Technol. 5, 3121–3125 (2016)

    Article  Google Scholar 

  199. Y. Dai et al., Mussel-inspired polydopamine-coated lanthanide nanoparticles for NIR-II/CT dual imaging and photothermal therapy. ACS Appl. Mater. Interfaces 9, 26674–26683 (2017)

    Article  Google Scholar 

  200. T. Jia et al., Mesoporous cerium oxide-coated upconversion nanoparticles for tumor-responsive chemo-photodynamic therapy and bioimaging. Chem. Sci. 10, 8618–8633 (2019)

    Article  Google Scholar 

  201. Y. Song et al., Energy migration control of multimodal emissions in an Er3+-doped nanostructure for information encryption and deep-learning decoding. Angew. Chem. Int. Edit. 133, 23983–23989 (2021)

    Article  ADS  Google Scholar 

  202. Y. Xie, Y. Song, G. Sun, P. Hu, A. Bednarkiewicz, L. Sun, Lanthanide-doped heterostructured nanocomposites toward advanced optical anti-counterfeiting and information storage. Light Sci. Appl. 11, 1–10 (2022)

    Article  Google Scholar 

  203. Y. Khan et al., A flexible organic reflectance oximeter array. Proc. Natl. Acad. Sci. USA 115, E11015–E11024 (2018)

    Article  ADS  Google Scholar 

  204. H. Xu, J. Liu, J. Zhang, G. Zhou, N. Luo, N. Zhao, Flexible organic/inorganic hybrid near-infrared photoplethysmogram sensor for cardiovascular monitoring. Adv. Mater. 29, 1700975 (2017)

    Article  Google Scholar 

  205. A. El Gamal, H. Eltoukhy, CMOS image sensors. IEEE Circuits Devices Mag. 21, 6–20 (2005)

    Article  Google Scholar 

  206. Y. Song et al., Digital cameras with designs inspired by the arthropod eye. Nature 497, 95–99 (2013)

    Article  ADS  Google Scholar 

  207. T. Mueller, F. Xia, P. Avouris, Graphene photodetectors for high-speed optical communications. Nat. Photon. 4, 297–301 (2010)

    Article  Google Scholar 

  208. C. Bao et al., High performance and stable all-inorganic metal halide perovskite-based photodetectors for optical communication applications. Adv. Mater. 30, 1803422 (2018)

    Article  Google Scholar 

  209. M. Boberl, M.V. Kovalenko, S. Gamerith, E.J.W. List, W. Heiss, Inkjet-printed nanocrystal photodetectors operating up to 3 μm wavelengths. Adv. Mater. 19, 3574–3578 (2007)

    Article  Google Scholar 

  210. J. Chen et al., A submicrosecond-response ultraviolet-visible-near-infrared broadband photodetector based on 2D tellurosilicate inSiTe3. ACS Nano (2022). https://doi.org/10.1021/acsnano.1c11628

    Article  Google Scholar 

  211. J. Michel, J. Liu, L. Kimerling, High-performance Ge-on-Si photodetectors. Nat. Photon. 4, 527–534 (2010)

    Article  ADS  Google Scholar 

  212. L. Xie et al., Broadband detection of X-ray, ultraviolet, and near-infrared photons using solution-processed perovskite-lanthanide nanotransducers. Adv. Mater. 33, 2101852 (2021)

    Article  Google Scholar 

  213. G. Anjaiah, S.K.N. Rasool, P. Kistaiah, Spectroscopic and visible luminescence properties of rare earth ions in lead fluoroborate glasses. J. Lumin. 159, 110–118 (2015)

    Article  Google Scholar 

  214. Y. Wang et al., Integrating positive and negative thermal quenching effect for ultrasensitive ratiometric temperature sensing and anti-counterfeiting. ACS Appl. Mater. Interfaces 13, 23951–23959 (2021)

    Article  Google Scholar 

  215. J. Legendziewicz, M. Guzik, J. Cybinska, VUV spectroscopy of double phosphates doped with rare earth ions. Opt. Mater. 31, 567–574 (2009)

    Article  ADS  Google Scholar 

  216. Y. Liu, J. Zhang, C. Zhang, J. Jiang, H. Jiang, High efficiency green phosphor Ba9Lu2Si6O24:Tb3+: visible quantum cutting via cross-relaxation energy transfers. J. Phys. Chem. C 120, 2362–2370 (2016)

    Article  Google Scholar 

  217. X. Ou et al., Recent development in X-ray imaging technology: future and challenges. Research 2021, 9892152 (2021)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This work was supported by the Zhejiang Provincial Natural Science Foundation of China (No. LZ21A040002), the National Natural Science Foundation of China (No. 52172164). The Institute for Lasers, Photonics and Biophotonics acknowledges support from the Office of Vice President for Research and Economic Development at the university at Buffalo.

Author information

Authors and Affiliations

Authors

Contributions

Lei Lei and Yubin Wang contributed equally to this work. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Lei Lei, Shiqing Xu or Paras N. Prasad.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lei, L., Wang, Y., Kuzmin, A. et al. Next generation lanthanide doped nanoscintillators and photon converters. eLight 2, 17 (2022). https://doi.org/10.1186/s43593-022-00024-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s43593-022-00024-0