Innovations in scintillator materials for X-ray detection

Kuilin Li , Wenqing Li, Qi Nie* and Xiao Luo*
School of Optoelectronic Science and Engineering, University of Electronic Science and Technology of China (UESTC), Chengdu 610054, PR China. E-mail: uestcnq@outlook.com; luox@uestc.edu.cn

Received 7th March 2025 , Accepted 7th April 2025

First published on 8th April 2025


Abstract

Since their introduction in the early 20th century, scintillators have become essential components of a wide range of applications, including high-energy physics, medical imaging, cryptography, and nuclear detection. As the demand for high-performance scintillating materials continues to rise in particle physics experiments and medical imaging technologies, the development of novel scintillator materials has become a critical area of research. In recent years, advancements in scintillators have flourished, presenting new opportunities for practical applications. This review presents a comprehensive overview of standard performance parameters for scintillators, aimed at enhancing our understanding and evaluating their advancements. Unlike previous reviews focusing on isolated material categories, this work provides a cross-comparative analysis of emerging scintillators, with particular emphasis on challenges for high-precision detection and low-dose imaging. We highlight the latest developments in scintillator materials, emphasizing research from the past three years and focusing on their intrinsic properties. Our analysis covers the perovskite scintillators, nanocluster scintillators, rare-earth ion-doped scintillators, organic scintillators, and scintillators with specialized structures. This classification offers a scientific perspective on the overall progress in the field of scintillators, and several forward-looking insights into the future development of scintillators are proposed, employing a problem-oriented approach. Future scintillator development requires synergistic material design integrating computational modeling and scalable fabrication techniques to enhance stability, radiation tolerance, and light yields. Prioritizing lead-free systems and defect-tolerant lattice engineering will address environmental and operational challenges, and advancements in hybrid architectures, and novel optical structures promise breakthroughs in low-dose imaging, industrial nondestructive testing and sustainable radiation detection technologies. Eventually, we discuss the challenges encountered in scintillator development, explore future prospects, and provide valuable insights for improving their performances and expanding their applications.


1. Introduction

X-rays are a form of electromagnetic radiation with exceedingly short wavelengths, typically ranging from 0.01 to 10 nm. Their remarkable penetrating power and high energy enable them to traverse various materials, including soft tissues in the human body and denser substances. Due to their exceptional capacity for penetration and imaging, X-rays find extensive application across multiple domains, such as medical imaging, industrial non-destructive testing, encryption and security screenings.1–5 Since the groundbreaking discovery of X-rays by Wilhelm Röntgen in 1895,6 this technology has rapidly evolved and made significant strides, gradually becoming an indispensable part of modern science and technology. Röntgen's discovery not only spurred in-depth investigations into the properties and behavior of X-rays but also ushered in a new era in imaging science, providing a novel tool for medical diagnosis.7 Through modalities such as X-ray films, computed tomography (CT), and digital imaging, physicians can observe the internal structures and lesions of patients with remarkable clarity, vastly improving the quality and efficacy of medical services. Consequently, the development of materials capable of detecting X-rays at low radiation doses is crucial.8 Furthermore, this technology adeptly identifies welding defects, internal cracks, and other structural issues, ensuring product quality and operational safety.

X-ray detection can be categorized into direct and indirect methods based on different detection principles (Fig. 1). Direct detection involves the direct absorption of X-rays, which are then converted into electronic or biochemical signals.9 In contrast, indirect detection involves the conversion of X-rays into visible light or ultraviolet light using scintillation materials, followed by detection with optical devices such as photomultiplier tubes, charge-coupled devices, or complementary metal–oxide–semiconductor detectors.10 Indirect detection is generally more cost-effective than direct detection, and the availability of various scintillators allows for diverse applications, making them increasingly popular. Practically, take biological X-ray imaging as an example, X-rays pass through the biological specimen and create an image on the scintillator film. Then the scintillator film converts the high-energy X-ray image into a signal that can be detected by conventional imaging devices. Subsequently, this signal is processed by the corresponding software and transformed into an image that is recognizable by the human eye. Therefore, high-quality scintillator films are crucial in the application process.


image file: d5qi00671f-f1.tif
Fig. 1 Schematic diagram of the principles of direct and indirect X-ray detection.

Scintillators are indispensable components of indirect detection. When high-energy particles or electromagnetic radiation traverse the scintillators, they interact with the atoms within the material, exciting them to excited states. As these atoms return to their ground states, they release energy in the form of photons, emitting visible light or radiation of other wavelengths. After a long period of development, numerous traditional scintillators, characterized by their high light yield (LY) and exceptional energy resolution, have achieved commercial viability for their respective applications. For instance, due to their exceptionally high LYs, NaI:Tl and CsI:Tl have been utilized in medical imaging, X-ray detection, and medical radiography.11–13 Bi4Ge3O12 (BGO) is widely accepted as an commercial-applicable inorganic scintillator due to its mechanically stable oxide with proportional performance.14,15 Moreover, single-crystalline Lu1.8Y0.2SiO5:Ce is employed in positron emission tomography (PET), while GdWO4 and Gd2O2S:Tb are used in CT.16–18 Additionally, Y3Al5O12:Ce and BaF2 are utilized in high energy physics experiments.19,20 Despite their widespread application, these traditional scintillators still suffer from some drawbacks. The preparation of traditional scintillator materials typically entails a complex process, high temperatures, and the use of capital-intensive equipment, resulting in considerable costs.16,21,22 Furthermore, due to their intrinsic characteristics, these materials inevitably experience interference from a long afterglow, light scattering, and relatively low detection limit.23 Additionally, their inflexible nature and biotoxicity limit their potential for advancement. Moreover, the challenges in converting the radiation emission wavelength into the visible light spectrum further hinder their applicability.24

Recently, the rapid advancement of scintillator materials, driven by extensive and in-depth research, has captivated more researchers each year. Among these, metal halides have garnered significant attention due to their intriguing luminescence properties and high LYs. Since Chen et al. utilized fully inorganic perovskite nanocrystals as scintillators in 2018,24 achieving strong X-ray absorption, intense radioluminescence (RL) at visible wavelengths, flexibility, and ultra-low detection limits, the development of metal halides has progressed by leaps and bounds. Accordingly, many studies have comprehensively summarized recent research advancements in scintillator materials. For instance, in 2023, Wang et al. summarized strategies for enhancing the LY of halide perovskite scintillators and presented a roadmap for improving X-ray imaging performance.25 Zhou et al. summarized the latest advancements in low-dimensional metal halide scintillators and proposed future development directions.26 Some studies have predominantly concentrated on specific categories of scintillator materials. Fan et al. focused on reviewing the progress of organic–inorganic hybrid Mn(II) halide scintillators27 and Zhao et al. summarized advanced techniques for preparing rare-earth nanocrystalline scintillators and provided essential guidance for their application in biomedical research.28 Besides, rare-earth-doped nano-scintillators have emerged as prominent scintillators in the field of biomedical applications. For example, Hong et al. presented a summary of the latest developments in advanced X-ray luminescence, focusing on its applications in imaging, biosensing, theragnostics, and neuromodulation through optogenetics.8 In 2024, Md. Helal Miah et al. summarized advancements in X-ray detection technologies across various materials, with a focus on methodologies, structural designs, and device architectures. Notably, the review provides crucial insights into advances in perovskite materials for X-ray detection and imaging applications, emphasizing strategies to optimize efficiency, stability, and scalability for next-generation radiation sensing systems. To conclude, the development of scintillators is both extensive and profound. However, there is a noticeable absence of a comprehensive return to the development of advanced scintillator materials.

In this review, we initially summarize a set of standard performance parameters for scintillators to facilitate a deeper understanding and evaluation of their development. We update the latest developments in scintillator materials based on their inherent properties, with a particular focus on research from the past three years. Then we focus on scintillators containing lead and lead-free scintillators, Cu-based clusters, rare-earth-ion-doped scintillators, organic scintillators as well as certain specialized structures. Our classification provides a scientific overview of the overall development of scintillators, aiding researchers in gaining a comprehensive understanding of the latest advancements in the field. Ultimately, this review also addresses the challenges faced in the development of scintillators, analyzes the prospects for their future advancement, and offers insightful visions for enhancing the performance of scintillators and their applications. Unlike prior reviews that adopt siloed discussions of conventional material systems, this study pioneers a systematic cross-disciplinary assessment of next-generation scintillators, encompassing metal halide perovskites, copper-based nanoclusters, rare-earth nanocrystallites and triplet-harvesting organic matrices. By critically evaluating their interdependent photophysical limitations, such as ion migration in perovskites, environmental instability in Cu-based architectures, and triplet exciton quenching in organic systems, we have a clear and comprehensive understanding of the advantages and disadvantages of each material in the new generation of scintillators.

2. Principle

In this section, we focus on the intricate details of both material and device characteristics, with particular emphasis on the interdependence between them. The performance and features of the device are inevitably shaped by inherent properties of the materials employed. This comprehensive exploration aims to elucidate the complex relationship between material characteristics and the corresponding performance metrics of the device, thereby providing the reader with a deeper understanding of the subsequent content.

2.1 Light yield and overall conversion efficiency for X-rays

The LY represents the ability of a scintillator to convert high-energy rays into low-energy visible light, which can be expressed as follows:
 
image file: d5qi00671f-t1.tif(1)
where S represents the transmission efficiency of electron–hole pairs to the optical emission centers, Q denotes the photoluminescence (PL) efficiency, β is a constant typically valued at 2.5, and Eg is the energy band gap. In scintillators, the standard time gate is usually set between 100 ns and 10 μs, and the exact value is determined through practical experimentation. A high-energy LY is essential for scintillators. Moreover, the overall conversion efficiency of X-rays in the scintillator can be defined as the number of UV/visible photons (Nph) generated during the scintillation conversion process per unit energy (E) of an incoming X-ray photon:
 
image file: d5qi00671f-t2.tif(2)

2.2 Attenuation coefficient

The attenuation coefficient (AC) indicates how effectively the detection material can absorb high-energy rays. It has a positive correlation with both its density ρ and effective atomic number Zeff. The attenuation coefficient is directly proportional to these factors and can be expressed as:
 
image file: d5qi00671f-t3.tif(3)
where A is the atomic mass and E is the radiation energy. This equation indicates that heavy elements and a high crystal density prevail in scintillators.

2.3 Afterglow

Afterglow refers to the time required for the intensity of RL in scintillators to diminish to a certain level after X-ray exposure. Due to the superimposition of previously exposed signals, prolonged afterglow can result in imaging artifacts. Therefore, a shorter afterglow duration is more desirable for common occasions. Yet along with the extension of scintillator applications, a long afterglow can be welcomed in some particular situations, like imaging scenes with memory properties, which are discussed in the following sections.

2.4 Spatial resolution

Spatial resolution pertains to the capacity to discern adjacent details within an object and their relative clarity. Specifically, higher spatial resolution is indicative of superior image quality. In scintillators, spatial resolution is primarily influenced by the geometric shape and intrinsic morphology of the scintillator itself, including its shape, size, and thickness. It has been widely accepted that the modulation transfer function (MTF), equivalent to 0.2, is derived from X-ray imaging to characterize spatial resolution. This also reflects the limitations of human visual discernment, typically measured in line pairs per millimeter (lp mm−1). It is important to note that the thickness of the scintillator significantly impacts imaging resolution. Thus, for discussions concerning scintillators in studies, it is essential to address both spatial resolution and thickness concurrently.

2.5 Linear response of scintillators and matching with RL and the photodetector's spectrum

RL refers to visible or near-visible light emitted by a scintillator under the excitation of ionizing radiation. For the RL of scintillators, the radiation intensity must have a linear relationship with the dose of X-rays. Moreover, the wavelength of the RL needs to match the efficient operational wavelength region of the detector. It is also very important to select an appropriate detector for different scintillators.

2.6 Detection limit

The detection limit is defined as the equivalent dose rate that generates a signal three times that of the noise level. The captured signal should be recognized as valid only when the incident X-ray dose exceeds this limit. To lower this threshold and enhance sensitivity, it is crucial to minimize the dark current and improve the signal-to-noise ratio. In medical imaging, a smaller detection limit reduces the exposure to X-rays and thereby minimizes the potential harm to biological tissues.

2.7 Chemical stability and radiation resistance

Chemical stability is a crucial indicator of a scintillator's ability to maintain its performance under various environmental conditions, including moisture resistance, temperature fluctuations, and oxidation resistance. For instance, CsI and NaI:Tl may undergo degradation in high-humidity environments, resulting in decreased luminescence efficiency. As for the radiation resistance, it is a key factor when assessing the ability of scintillator materials to preserve their performance upon exposure to high-energy radiation. Under exposure to high doses of X-rays, the lattice structure of scintillator materials may sustain damage, leading to a reduction in light output and a decline in energy resolution. Enhancing both chemical stability and radiation resistance can significantly improve the performance of X-ray scintillators. This enables them to maintain excellent detection capabilities and application efficacy in various harsh environments.

3. Technology innovations

This section provides a categorized overview of commonly used scintillator materials, highlighting their types, characteristics, recent representative works, and potential issues (Fig. 2). First, we introduce lead-based perovskite scintillators, which have been the subject of extensive research due to the strong X-ray absorption properties of Pb. This category remains active, with new studies continuously emerging. Next, we focus on lead-free scintillators, which have gained significant attention in recent years due to environmental and sustainable development concerns. Among these, copper-based scintillators are considered an excellent alternative to Pb and represent the main category within lead-free materials, as discussed in section 3.2.2. Additionally, there has been increasing interest in emerging copper-based nanoclusters, which are also being explored as scintillator materials, due to their exceptional optical properties. Rare earth elements, commonly used in optical materials, have attracted significant attention for their ability to regulate energy levels, with numerous applications in X-ray scintillators, which we address in detail in section 3.4. Furthermore, organic materials, due to their controllability and structural diversity, have also been extensively researched within the field of scintillators. Lastly, we present an overview of special structures in the scintillator domain, incorporating recent advancements and relevant studies.
image file: d5qi00671f-f2.tif
Fig. 2 Brief overview of this review.

3.1 Lead-based perovskite scintillators

Benefiting from adjustable bandgaps, high spatial resolution, fast response times, high LYs and considerable X-ray attenuation,24,29–32 lead-based perovskite scintillators have consistently been a major focus of research efforts and many valuable efforts have explored these charming materials (Table 1).24,33–52 Nevertheless, they still suffer from several drawbacks. The instability of a perovskite is always a non-negligible issue because of its hygroscopicity. Water and oxygen expand perovskite molecules, causing aggregation and affecting their optical properties, which will even be prominent under irradiation with X-rays,53 and the typical method for improving a perovskite's stability is to passivate its surface or reduce inner defects. Favorable and accepted methods are to embed them in solid substances or encapsulate them with polymers like polyethylene glycol terephthalate (PET), polytetrafluoroethylene (PTFE) and poly(methyl methacrylate) (PMMA).54
Table 1 The parameters of lead-based scintillators
Material Type Light yield (photons per MeV) Detection limit (nGyair s−1) Spatial resolution (lp mm−1) Ref.
CsPbBr3 Nanocrystals 9001 9 57
CsPbBr3 Nanocrystals 58
CsPbBr3/Cs4PbBr6 Nanocrystals 33[thin space (1/6-em)]800 79 8.9 59
CsPbBr3 Nanocrystals 20 60
MAPbBr3 MOFs 170 14.7 61
CsPbBr3 Nanocrystals 326 13.9 62
MAPbMgZnCdBr3 Nanocrystals 21[thin space (1/6-em)]200 462 11.7 63
CsPb2Br5:K+ Nanocrystals 25[thin space (1/6-em)]160 162.3 21 64


Based on similar methodologies, some groups have already completed related work,55,56 and in the past two years more noteworthy studies have been reported. Lv et al. combined CsPbBr3 nanocrystals with PMMA to fabricate a perovskite film of over 100 cm2 with a root-mean-square roughness of 1.06 nm, which retained 96% initial luminous intensity in water and an ambient environment for 480 h and 2880 h (Fig. 3a).57 Zheng et al. selected poly(maleic anhydride-alt-1-octadecene) (PMAO) with a low molecular weight (Mn = 1911) as the multi-dentate ligands and synthesized PMAO–CsPbBr3 lead perovskite nanocrystals (PNCs) via hot injection. It shows preeminent water stability, maintaining its fluorescence for 2600 min while traditional PNCs became completely dysfunctional after 100 min, as observed in Fig. 3b.58 For the embedded-in-matrix strategy, Shen et al. applied cold-sintered technology to bulk scintillators to fabricate CsPbBr3 nanocrystals embedded in the Cs4PbBr6 structure based on an “emitter-in-matrix” principle (Fig. 3c).59 This group employed a rather low temperature of 90 °C during the sintering process and DMSO was introduced as the antisolvent to promote its densification (Fig. 3d). Compared to solid-state sintering at high temperature, cold sintering demonstrates highly compacted grains with enhanced densification and transparency, thus scattering can be well controlled. These bulk scintillators exhibit a high LY of 33[thin space (1/6-em)]800 photons per MeV, a low detection limit of 79 nGyair s−1 and outstanding spatial resolution of 8.9 lp mm−1.


image file: d5qi00671f-f3.tif
Fig. 3 (a) Photographs of the CsPbBr3@PMMA@VmB1 film in air at various times to show its stability. (b) PL evolution of PNCs after being immersed in water for different times (inset: optical images of the PMAO-PNCs in water under vigorous stirring). (c) Design of the fabrication process to obtain bulk CsPbBr3/Cs4PbBr6 ceramic composite pellets via a CSP. (d) Light scattering of residual pores and grain boundaries in conventional solid-state sintered ceramics, reducing energy resolution and transparency (left). Reduction of light scattering from grain boundaries in cold-sintered ceramics to enhance resolution and transparency (right). (e) SEM images of the PNC/silicone films with APTES/OA ligands (upper) and APTES/APTES-PA ligands (bottom). (f) Schematic illustration of the synthesis of bMOF–MAPbBr3 and bMOF–MAPbBr3–PMMA films. (g) Comparison of LY between the MAPbMgZnCdBr3/PVDF-HFP scintillator and LuAG:Ce.

Apart from those tactics mentioned for enhancing perovskite stability, some other interesting methods have appeared. Chen et al. found it detrimental to cross-linking of the organosilicon matrix that the silicon-based CsPbBr3 encapsulated nanocrystals were linked with long-carbon-chain ligands, thus suffering from poor transparency and surface roughness.60 Yet this can be improved by substituting long chain ligands for short ones (Fig. 3e). Then a dual-organosilicon ligand system was proposed, consisting of (3-aminopropyl)triethoxysilane (APTES) and (3-aminopropyl)-triethoxysilane with pentanedioic anhydride (APTES-PA), and the scintillator showed a satisfying spatial resolution above 20 lp mm−1 and high stability in various harsh environments. Wu et al. applied the host–guest strategy, involving confining MAPbBr3 nanocrystals in metal–organic frameworks (MOFs) (Fig. 3f).61 The in situ spatial confinement provides quantum confinement and surface passivation, resulting in significant stability against air, heat and irradiation. It exhibits a detection limit of 170 nGyair s−1 and an impressive spatial resolution of 14.7 lp mm−1. Likewise, Li et al. presented a glass-ceramic CsPbBr3 perovskite film with high transparency, in which CsPbBr3 crystals grew in situ.62 The thickness was controlled by the crystal coordination–topology growth and in situ film formation strategy, resulting in little light scattering and a uniform distribution of the perovskite crystals. The as synthesized 250 μm film enables a detection limit of 326 nGyair s−1 and a high spatial resolution of 13.9 lp mm−1 with incredible stability under exposure to long-term X-ray irradiation, water soaking and high temperature.

Besides the stability improvements, progress was also made in the energy utilization efficiency by doping with new elements. Xiang et al. performed calculations on the surface formation energy and the vacancy defect energy and co-doped several bivalent metal ions to substitute Pb in MAPbBr3 to improve the photoluminescence quantum yield (PLQY) and decay time of the nanocrystal.63 Furthermore, the scintillator was fabricated using MAPbMgZnCdBr3 as the energy conversion layer with PVDF-HFP as the carrier and demonstrated a LY of 21[thin space (1/6-em)]200 photons per MeV (Fig. 3g) with a spatial resolution of 11.7 lp mm−1. Moreover, Qiu et al. doped non-emissive CsPb2Br5 with alkali metals, which caused lattice distortion and enhanced electron–phonon coupling.64 This led to the formation of potential transient energy wells, resulting in the creation of self-trapped excitons (STEs) generated by X-rays. The STE experiences radiative recombination, with a PLQY of 55.92%, thus X-rays can be transformed into red light. Furthermore, the K+-doped CsPb2Br5 based X-ray scintillator displays high stability and weak self-absorption with a detection limit of 162.3 nGyair s−1 and 21 lp mm−1.

Despite being a long-investigated scintillator material, it still suffers from non-negligible drawbacks. The predominant emission mechanism via band-edge excitons introduces substantial self-absorption effects owing to spectral overlap between its emission and absorption profiles. This fundamental limitation persists as a pervasive challenge inherent to virtually all perovskite-based systems. While lead-based perovskite scintillators exhibit unparalleled performance metrics, the absence of innovative stabilization strategies remains a critical bottleneck. No groundbreaking processing techniques addressing environmental robustness have emerged in the latest literature, which confines their practical deployment to laboratory-scale demonstrations or strictly controlled environments, as evidenced by rapid performance degradation. In addition, the potential environmental leakage risk of heavy metals further hinders large-scale applications.

3.2 Lead-free perovskite scintillators

Although Pb based scintillators exhibit promising absorption of X-rays, the heavy atom's intrinsic toxicity, doing harm to human health and environment, is still a non-negligible hindrance to its extensive applications. Therefore, the pursuit of ‘lead-free’ scintillators has been put forward, expecting materials with no toxic element yet retaining exceptional and controllable properties. In recent years, research on lead-free scintillators has significantly outpaced that on traditional lead-based scintillators, especially in the past two years (Table 2).65–67 These materials can be categorized into the following types: double perovskite scintillators, manganese-based halide scintillators and others; these are all introduced and discussed in the following sections.
Table 2 The parameters of lead-free scintillators
Material Type Light yield (photons per MeV) Detection limit (nGyair s−1) Spatial resolution (lp mm−1) Ref.
Cs3Cu2I5:Mn Single-crystal 95[thin space (1/6-em)]772 100
Rb2AgI3:Cu+ Single-crystal 53[thin space (1/6-em)]983 231 14.1 101
Cs3Cu2I5:HCOO Single crystals 61[thin space (1/6-em)]500 10 102
Cs3Cu2I5 0D 12[thin space (1/6-em)]760[thin space (1/6-em)]000 10 103
Cs3Cu2I5 Nanocrystals 105 14.3 104
Cs3Cu2I5 Nanocrystals 105
Cs3Cu2Cl5 Single crystal 95[thin space (1/6-em)]000 2700 105 106
CsCu2I3 Crystal 88.9 15.6 107
Cs5Cu3Cl6I2 1D 64[thin space (1/6-em)]800–67[thin space (1/6-em)]200 27 108
Cs5Cu3Cl6I2 1D 109
Cs5Cu3Cl6I2:Rb Crystal 64[thin space (1/6-em)]868 14 110
Rb2AgBr3:Cu+ Single crystal 79[thin space (1/6-em)]250 714.83 5.8 111
Rb2AgI3:Cu+ 1D single crystals 36[thin space (1/6-em)]293 1022 10.2 112
Rb2AgBr3:Cu+ Single-crystal 10.2 113
Cs2NaBiCl6 Nanocrystals 28[thin space (1/6-em)]350 45.2 14.76 74
Cs2TeCl6 Microcrystals 38[thin space (1/6-em)]523 258 15.9 66
Cs2HfCl6 Microcrystals 21[thin space (1/6-em)]700 55 11.2 67
Cs4Cd1−xMnxBi2Cl12 2D 34[thin space (1/6-em)]450 183.6 16.7 77
(C25H22P)2MnCl4 Single crystals 78[thin space (1/6-em)]937 25.61 82
(C38H34P2)MnBr4 Crystal 35[thin space (1/6-em)]945 5500 16.7 83
BPP2MnBr4 0D 35[thin space (1/6-em)]000 250 20 84
(MTP)2MnBr4 Crystal 67[thin space (1/6-em)]000 82.4 6.2 85
PrPP2MnBr4 Single crystal 43[thin space (1/6-em)]511 13.9 53.5 86
(Br-PrTPP)2MnBr4 0D 68[thin space (1/6-em)]000 45 12.78 87
(BTPP)2MnX4 0D 53[thin space (1/6-em)]000 89.9 14.1 88
A2MnBr4 0D 80[thin space (1/6-em)]100 30 14.06 89
DMA4InCl7:Sb3+ 0D 23[thin space (1/6-em)]500 175 90
TTA2In0.92Sb0.08Cl5 0D 60[thin space (1/6-em)]976 90 91
TpyBiCl5 Single crystal 196.31 92
Ag6S6L6 (SC-Ag) Metal cluster 17[thin space (1/6-em)]420 16 93
Ag2Cl2(dppb)2 Single crystal 79[thin space (1/6-em)]970 59.8 25 94
Cs2Ag.6Na.4In.85Bi.15Cl6 Nanocrystals 95
Cs2CdCl4:10%Mn Single crystal 88[thin space (1/6-em)]138 31.04 16.1 96
Tb3+–Cs2NaGdCl6 Single crystals 39[thin space (1/6-em)]100 41.32 10.75 78
Te4+ dopingCs2SnCl6 Nanocrystals 132 20 79


3.2.1 Copper-free scintillators. One of the important forms of copper-free scintillator materials is the double perovskite material. Their development stemmed from the recognition that, while perovskite materials have long been a focus of scintillator research due to the advantages in their synthesis, the isolated network of metal-halide octahedral units leads to poor carrier mobility and predominantly indirect bandgap characteristics.68–70 In contrast, the double perovskite materials, typically represented as A2B3X6 (where A is a monovalent metal ion like Ag+ and Cs+, B is a trivalent metal ion like In3+ and Sb3+, X refers to a halide anion), though suffering from lower carrier mobility and a larger indirect bandgap compared to the perovskites, possess a cubic structure with interlinked [B+X6]5− and [B+X6]3− sites and do not contain toxic Pb elements.69,71–73 This objectively enhances the material's stability and environmental friendliness, making it an attractive alternative to perovskite materials for researchers. Efforts have been made to circumvent these drawbacks. For example, Varnakavi et al. managed to inject 4.04% of Mn(II) into a double perovskite Cs2NaBiCl6 nanocrystal (Fig. 4a), which turned out to have a PLQY of 10.6% at 586 nm, circumventing self-absorption.74 The Cs2NaBiCl6:4.04%Mn@poly(methyl methacrylate) scintillator film demonstrates a satisfying LY of 28[thin space (1/6-em)]350 photons per MeV and a detection limit of 45.2 nGyair s−1, along with a rather exceptional spatial resolution of 14.76 lp mm−1. Recent advances in Bi3+-doped perovskite scintillators75 highlight their critical role in performance enhancement. In 2025, Chen et al. engineered 3D bismuth chloride perovskites via monovalent Na+ incorporation, achieving a sensitivity of 354.5 mCGy−1 cm−2 and an ultralow detection limit of 59.4 nGy s−1 with stable cycling over 4500 on–off cycles.76 Complementing this, Subagyo et al. demonstrated the efficacy of Bi3+ in optimizing radiation hardness, LY, and defect dynamics, underscoring its potential for sustainable high-performance X-ray detectors. Together, these studies exemplify the versatility of Bi3+ in advancing perovskite scintillators for precision radiation sensing.65 In parallel, Lai et al. discovered that the Te-based double perovskite, Cs2TeCl6 (CTC) microcrystal, showed a broadband yellow emission which was highly matched with CCD.66 Furthermore, the hafnium alloyed CTC scintillator, with PDMS as the binder, demonstrates a LY of 38[thin space (1/6-em)]523 photons per MeV (Fig. 4b), a detection limit of 258 nGyair s−1 (Fig. 4c) and an outstanding spatial resolution of 15.9 lp mm−1. Likewise, Zhang et al. came up with Cs2HfCl6, and under ambient conditions for 60 days, there were no extra diffraction peaks observed in XRD spectra, indicating promising stability, which was possibly due to short Hf–Cl bonds, with a spatial resolution of 11.2 lp mm−1.67 Moreover, Li et al. reported a 2D layered double perovskite, namely Cs4Cd1−xMnxBi2Cl12, which demonstrated preeminent scintillation properties, including exceptional emission response linearity and a high LY of ∼34[thin space (1/6-em)]450 photons per MeV.77 More admirably, the X-ray excited emission intensity remained at 92% and 94% respectively after being stored under an ambient atmosphere and exposed to a total dose of 11.4 Gy irradiation. After being mixed with PDMS, the scintillator screen exhibits an excellent spatial resolution of 16.7 lp mm−1. Subsequently, Naresh et al. proposed the Cs2NaGdCl6:5%Tb3+ films with 0.4 mm thickness and achieved a low detection limit of 41.32 nGyair s−1, an LY of 39[thin space (1/6-em)]100 photons per MeV and excellent radiation stability.78 After the introduction of Tb3+, the PLQY increased from 8.4% to 72.6% due to efficient energy transfer from STE to Tb3+. Likewise, Wang et al. introduced Te4+ into the double perovskite Cs2SnCl6 and fabricated flexible imaging screens with a resolution of 20 lp mm−1 and a low detection limit of 132 nGyair s−1.79
image file: d5qi00671f-f4.tif
Fig. 4 (a) Schematic diagram of Mn2+-doped Cs2NaBiCl6 showing STE dynamics and the Mn2+ PL mechanism. (b) RL of CHTC and CTC. (c) RL of CHTC and CTC as a function of dose rate. (d) RL spectra of (MTP)2MnBr4 and BGO under X-ray irradiation (left). Integrated RL intensity of (MTP)2MnBr4 as a function of X-ray dose rate (middle). Time-dependent RL intensity of (MTP)2MnBr4 upon continuous X-ray irradiation at a dose rate of 42.29 mGy s−1 for 60 min (right). (e) The MTF of the PrPP2MnBr4 scintillation film beveled edge image. (f) Schematic diagram of the emission mechanism with STE (T) and STE (S). (g) The MTF of the Ag2Cl2(dppb)2 (Ag1) scintillator edge images.

Manganese-based halides have only been utilized in the past five years. Due to their excellent LY and high energy radiation absorption performance, they are designed to be potential candidates for efficient and environmentally friendly scintillator materials.80,81 Here we focus mainly on research published in the past year. For instance, Zhou et al. designed and synthesized the thickness-tunable inch-sized Mn-based halide single crystal, (C25H22P)2MnCl4, with a PLQY up to 99.10% and the scintillator film showed a LY of 78[thin space (1/6-em)]937 photons per MeV and a spatial resolution of 25.61 lp mm−1.82 Liu et al. synthesized the (C38H34P2)MnBr4 organogel scintillator and made it water-stable, stretchable and self-healing by carefully arranging and activating dynamic hydrogen bonds and coordination bonds.83 The organogel scintillator is able to be stretched up to 13-fold while maintaining its property, and the imaging resolution reached 16.7 lp mm−1. The scintillator film demonstrated seamless self-healing within 30 min, providing a robust solution for the self-repair of flexible films that may experience bending or cutting damage during usage. Wang et al. reported a method to control the crystallization process of manganese halides through organic cation modulation.84 This approach aims to reduce uncontrolled crystallization during the fabrication of large-area flat-panel detectors, thereby enhancing their light-emitting properties. The large steric hindrance brought by the cation provides a high exciton binding energy and restrains non-radiative recombination. The as synthesized BPP2MnBr4 (BPP = C25H22P+) exhibits a superior spatial resolution of more than 20 lp mm−1, an exceptional LY of over 35[thin space (1/6-em)]000 photons per MeV and an ultra-low detection limit of less than 250 nGyair s−1. Likewise, Zhang et al. reported the potential of applying (MTP)2MnBr4 (MTP refers to methyltriphenylphosphonium) as an X-ray scintillator for its near-unity PLQY, high resistance to thermal quenching and decent stability.85 They fabricated a thin film scintillator and determined its impressive LY of 67[thin space (1/6-em)]000 photons per MeV and superior detection limit of 82.4 nGy s−1 (Fig. 4d).

Moreover, Xiao et al. synthesized PrPP2MnBr4 (PrPP2 refers to diisopropylammonium) by a dual-solvent evaporation method, and the 10 cm × 10 cm area scintillation screen based on the solidified melt salt exhibited a LY of 43[thin space (1/6-em)]511 photons per MeV, an unprecedented high spatial resolution of 53.5 lp mm−1 (Fig. 4e) and an outstanding detection limit of 13.9 nGyair s−1.86 Also, Zhang et al., recognizing the satisfying optical and scintillator properties of 0D organic–inorganic hybrid metal halides (OIMHs), synthesized (Br-PrTPP)2MnBr4 (Br-PrTPP refers to (3-bromopropyl) triphenylphosphonium) via a facile saturated crystallization method.87 Attributed to the tetrahedrally coordinated [MnBr4]2− polyhedron, it demonstrates an excellent LY of up to 68[thin space (1/6-em)]000 photons per MeV and lowest detection limit of 45 nGyair s−1. Li et al. prepared 0D (BTPP)2MnBr4 (BTPP refers to benzyltriphenylphosphonium) as a promising scintillator with excellent air and radiation stability, which showed an appreciable LY of 53[thin space (1/6-em)]000 photons per MeV, a low detection limit of 89.9 nGyair s−1 and an ultra-low afterglow comparable to the commercial BGO single crystal.88 Furthermore, after fabricated as a flexible screen combined with PDMS, a rather high spatial resolution of 14.1 lp mm−1 was achieved. In addition, Gong et al. presented a novel group of 0D hybrid manganese halides of A2MnBr4 (A = BzTPP, Br-BzTPP, and F-BzTPP).89 Benefiting from near-unity PLQY, these hybrids show a state-of-art LY of 80[thin space (1/6-em)]100 photons per MeV for hybrid manganese halides. The 30 nGyair s−1 ultra-low detection limit, 14.06 lp mm−1 qualified spatial resolution and 0.3 ms short afterglow make it a promising scintillator for further applications.

Apart from all these copper-free scintillators mentioned, there are still several noteworthy studies we would like to present. For example, Wang et al. discovered the introduction of a small amount of Sb3+ into the indium halide DMA4InCl7, resulting in DMA4InCl7:10%Sb3+, largely enhancing the PLQY by 9-fold, which originated from efficient energy transfer between Sb3+ ions and STE states.90 The STE emission yields a high LY of 23[thin space (1/6-em)]500 photons per MeV and a low detection limit of 175 nGyair s−1. Likewise, Huang et al. discovered the novel influence of the A-site on the band-edge structures and optoelectronic properties, and some Sb3+-doped 0D halide perovskite derivatives were designed as experimental proof.91 Benefiting from the deformation of the lattice brought about by Sb3+, STE states are harnessed properly (Fig. 4f), resulting in a qualified LY of up to 60[thin space (1/6-em)]976 photons per MeV and a detection limit of 90 nGyair s−1. Moreover, Wang et al. designed and synthesized a molecular hybrid perovskite, TpyBiCl5, which managed to circumvent thermal quenching via multi-excited state switching, originating from the inter-reaction among its counterparts in the perovskite.92 It also shows a rigid framework structure, resulting in 12 times higher PLQY than its organic ligand (Tpy) at room temperature. TpyBiCl5 maintains an ideal detection limit of 196.31 nGy s−1 even at 353 K through thermally activated delayed fluorescence (TADF), and a clear image at 413 K, effectively addressing potential annealing issues observed in other films. This characteristic offers a promising solution for the application of scintillators under extreme conditions, such as in situ imaging of thermal loss during mechanical operation. Furthermore, Wang et al. synthesized Ag6S6L6 (SC-Ag) metal clusters through a simple solvothermal reaction.93 The heavy atom guarantees an effective TADF mechanism, and satisfyingly, SC-Ag demonstrates a high PLQY of 91.6% and a LY of 17[thin space (1/6-em)]420 photons per MeV, along with a preeminent spatial resolution of 16 lp mm−1. Additionally, the rigid core structure endows SC-Ag with strong resistance to humidity and radiation. Additionally, Yuan et al. pioneered high-performance TADF Ag-based scintillators, namely Ag2Cl2(dppb)2 (dppb refers to 1,2-bis(diphenylphosphino)benzene), with an exceptionally high RL intensity and a low detection limit of 59.8 nGy s−1.94 Its superior performance, compared to the Cu series, mostly results from effective utilization of high excitons due to a small singlet–triplet energy gap. The flexible film fabricated has a high spatial resolution of 25 lp mm−1 (Fig. 4g). Another work worth mentioning is that of O'Neill et al. who synthesized and characterized Cs2Ag0.6Na0.4In0.85Bi0.15Cl6 (CANIBIC), an inorganic mixed-cation double halide perovskite, and figured out its STE emission mechanism.95 Moreover, the CANIBIC-PMMA film demonstrates higher X-ray luminescence intensity than that of the pure pressed pellet of CANIBIC. Also, Wang et al. reported a Mn2+-activated all-inorganic 2D layered Ruddlesden–Popper perovskite, namely Cs2CdCl4:10%Mn.96 It possesses a bright orange–red emission with 90.47% PLQY, strong X-ray absorption, an ultra-high LY up to 88[thin space (1/6-em)]138 photons per MeV and 31.04 nGyair s−1. A 15 cm × 15 cm flexible screen was prepared by combining PDMS with Cs2CdCl4:10%, with a high spatial resolution of 16.1 lp mm−1 and qualified imaging capability under an extreme low dose of 16 μGyair s−1.

While copper-free scintillators, as a subclass of their lead-free counterparts, enable elements with analogous elements, such as Mn, Bi, and Te, to better manifest their intrinsic properties for optimizing spectral characteristics and enhancing performance, critical challenges persist. These include the generation of toxic species from undesired byproducts during the synthesis of organic components in hybrid organic–inorganic systems, inherent parity-forbidden transitions in double perovskites necessitating additional strategies to improve the LY,97 and the hygroscopic nature of lead-free variants imposing stringent encapsulation requirements that limit their practical deployment.98

3.2.2 Copper-based scintillators. Cu substitution is the best accepted solution, and Cs3Cu2I5 is a widely reported Cu-based scintillator with prominent feasibility, because the zero-dimensional (0D) structure of Cs3Cu2I5 effectively localizes and segregates multiple STEs, thereby mitigating Auger recombination.99 Several works have been reported on this appealing material;100 here we focus only on work from recent years, for which researchers have focused their efforts on synthesis procedure progress or property enhancement. For instance, to avoid lattice distortion caused by heteromorphic homologs in vacuum-evaporated metal halides (MHs), Peng et al. brought out a single-source vacuum evaporation method.101 A ∼10 μm thick film of Cu-MHs was synthesized by using this very strategy, to achieve an impressive LY of 53[thin space (1/6-em)]983 photons per MeV and a spatial resolution of 14.1 lp mm−1.

Doping is recognized as a potential method for enhancing efficiency and stability. For example, Wang et al. focused their efforts on multi-self-trapped excitons of HCOO-doped Cs3Cu2I5 single crystals with transient and steady-state spectroscopy, and figured out that it derived from the host lattice and external dopants separately, which could both be triggered independently.102 Then, after carefully tuning the excitation wavelength and selectively exciting different STEs, Cs3Cu2I5:HCOO demonstrates a remarkable PLQY of 99.01% and little self-absorption, with LY being improved by 5.4 times, compared to that of pristine Cs3Cu2I5 crystals, up to 61[thin space (1/6-em)]500 photons per MeV (Fig. 5a). Also, like the lead-based perovskite scintillators, combination with poly-carriers is considered a way of enhancing stability. To illustrate this, Moon et al. reported novel technology that integrated Cs3Cu2I5 perovskite nanoparticles into densified–delignified wood, namely Cs3Cu2I5@D-Wood, for an eco-friendly and biodegradable X-ray scintillator film with a spatial resolution of 10 lp mm−1.103 Similarly, Hao et al. developed an innovative integration of Cs3Cu2I5 with PVDF and PMMA, effectively preventing agglomeration and ensuring a more uniform size distribution, which, in turn, improved stability.104 The as controlled flexible scintillator film displays a satisfying spatial resolution of 14.3 lp mm−1 and a low detection limit of 105 nGy s−1 was achieved.


image file: d5qi00671f-f5.tif
Fig. 5 (a) RL spectra of pristine and HCOO-doped Cs3Cu2I5 single crystals. (b) Schematic of SIXS for multi-energy X-ray detection and imaging. (c) Dynamic imaging of a screw after 10 s of X-ray excitation using CsI:Tl and Cs5Cu3Cl6I2 films. The Cs5Cu3Cl6I2 film shows no residual image, while the image in the CsI:Tl film persists even after 3 s. (d) Crystal structure diagram of Rb2AgBr3 (left) and the schematic substitution model via Cu+ doping (right). (e) Schematic of the mechanisms in Cu4I4 cubic clusters. (f) The PL mechanism in the configuration coordinate diagram of [BzTPP]2Cu2I4. (g) Bright-field and X-ray images of a USB drive. (h) POM images of M3 at different temperatures.

To further exert the functionality of Cs3Cu2I5 and its derivatives, some praiseworthy investigations have been conducted. For instance, considering that the absence of X-ray energy distinction shall result in insufficient image contrast, Ran et al. presented an innovative multi-energy X-ray linear-array detector based on Cs3Cu2I5 (Fig. 5b).105 Benefiting from the negligible self-absorption of the material and side-illuminated scintillation scenarios, the incident X-ray spectra could be reconstructed by analyzing the distribution of the scintillator intensity while illuminated from the side. As a result, the outcome error can be controlled to be below 5.63%. Xiang et al. presented centimeter-sized Cs3Cu2Cl5 via a slow-cooling method, and the planar orientation, which was controlled in a space-confined chamber, required no further shaping before being fabricated as a scintillation screen.106 The synthesized Cs3Cu2Cl5 single crystal exhibits an impressive LY of up to 95[thin space (1/6-em)]000 photons per MeV and a low detection limit of 2.7 μGyair s−1. Moreover, it could retain its original phase for 6 months in a vial after the surface passivation procedure. Intriguingly, Ran et al., the same group mentioned above, also discovered a better match between CsCu2I3, once considered the oxidation product, and the spectral responsivity of regular flat-panel photodiode arrays than Cs3Cu2I5, showing potential extension for further application in X-ray scintillators.107

Another interesting and noteworthy Cu-based scintillator was presented in recent years, and has attracted attention in the frontier research field, namely Cs5Cu3Cl6I2. Its presenters, Wu et al., discovered it possessed benign grain boundaries without dangling bonds, which were commonly considered to contribute to RL lifetime extension and nonradiative recombination loss increase.108 After fabrication as a columnar morphology via close space sublimation, it exhibits negligible afterglow and a high LY (Fig. 5c). The imaging property was also determined; it demonstrated a spatial resolution of 27 lp mm−1 with a frame rate up to 33 fps. Also, Zhang integrated a Cs5Cu3Cl6I2 event-based scintillator film with negligible ghosting artifacts (0.1%) and an impressive data compression ratio of 23.7%.109 Shu et al. reported the Rb doped Cs5Cu3Cl6I2 crystal with considerable improvement in stability against heat and humidity and an outstanding LY.110 Combined with PDMS, the Cs5Cu3Cl6I2:Rb scintillator screen reaches a high spatial resolution of over 10 lp mm−1.

Besides, copper(I) itself can also be used as a dopant to enhance the STE state. To illustrate, Hu et al. discovered that the introduction of 0.05% Cu+ into Rb2AgBr3:Cu+ could largely enhance the PLQY to 98.8% (Fig. 5d).111 Consequently, a preeminent LY of 79[thin space (1/6-em)]250 photons per MeV and rather low detection limit of 714.83 nGy s−1 were achieved. Likewise, Yao et al. synthesized 1D Cu+-doped Rb2AgI3 with 76.48% PLQY, demonstrating a LY of 36[thin space (1/6-em)]293 photons per MeV, a low detection limit of 1.022 μGyair s−1 and a decay time of 465 ns.112 In very recent work, Wu et al. further enhanced the performance of Rb2AgBr3:Cu+, achieving single crystals with a PLQY of up to 99.2%.113 Through pulsed decay measurements and DFT calculations, they demonstrated that the incorporation of Cu increased the density of deep bound excitons, thereby significantly enhancing radiative recombination. By integrating the material with PDMS, they successfully fabricated a large-area flexible scintillator film, which exhibited a spatial resolution of 10.2 lp mm−1, indicating its potential applications in radiation monitoring and biomedical imaging.

Copper-based systems, serving as a viable substitute for their lead-containing counterparts, have emerged as one of the most promising non-toxic scintillator candidates. Nevertheless, even though elemental doping strategies have been employed to mitigate their degradation under ambient conditions, the degradation pathways, such as oxidation or segregation remain non-negligible.114 Furthermore, the spectral mismatch between their emission profile and the responsivity spectra of commercially available photodetectors imposes additional constraints on achieving optimal detection efficiency.115

3.3 Cu-based cluster scintillators

Another huge family of Cu-based scintillator materials are CuxIx clusters, usually combined with organic cations or polymers, which no doubt are emerging as new X-ray scintillators (Table 3). Blessed with double STE emission, these clusters usually exhibit a considerable LY, a large Stokes shift and high-efficiency white light emission.116–118 As a highly regarded material, major efforts have been reported. For instance, Lai et al. prepared organic cuprous iodide clusters, (IPP)CuI2 (IPP refers to iso-propyltriphenylphosphonium).119 These present the best scintillator property when IPP[thin space (1/6-em)]:[thin space (1/6-em)]CuI = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.75, with a super low detection limit of 62.29 nGy s−1 and an ultrahigh LY of 75[thin space (1/6-em)]793.83 photons per MeV. They also applied it to CT and obtained a clear reconstructed 3D model, due to its superior spatial resolution of 10.2 lp mm−1. Zhao et al. designed and fabricated a novel metal halide scintillator, (C12H28N)2Cu2I4, with white emission.120 Benefitting from the double self-trapped mechanism, it demonstrates a really high PLQY, which not only matches with the response of semiconductor-based sensors, but also decreases the doses to which the objects are exposed. The flexible transparent large-area scintillator film made from this shows a superior luminescence property, a high spatial resolution of 19.8 lp mm−1 and an incredibly low detection limit of 28.39 nGyair s−1; these values are four times higher and 194 times lower than the typical value in medical usage. Moreover, recognizing the trivial factor that X-ray luminescence efficiency in metal clusters is governed by the competition between radiative states from organic ligands and non-radiative ones (Fig. 5e), Zhang et al. reported a class of Cu4I4 cubes, including [DDPACDBFDP]2Cu4I4 (DDPACDBFDP refers to 10,10′-(4,6-bis(diphenylphosphanyl))). It presents an impressive emissive RL due to functionalizing the biphosphine ligands with acridine, which can effectively absorb ionization radiation to generate electron–hole pairs and subsequently transfer them to ligands during thermalization.121 It demonstrates a PLQY of 95%, benefiting from triplet-to-singlet conversion via a TADF matrix, the lowest detection limit of 77 nGy s−1 and a high spatial resolution of up to 12 lp mm−1. Similarly, Zou et al. employed a solvent-recyclable green process to synthesize stable three-dimensional nanoclusters, Cu4I4(C6H14N2)2, achieving a PLQY close to 100%.122 The sturdy 3D framework ensures the morphological stability of the molecules at high temperatures and humidity. It retains 90% of its room-temperature luminescence intensity at 110 °C and 88.34% of its initial luminescence intensity after 100 days under ambient conditions. This demonstrates a promising approach for the green, low-toxicity synthesis of high-performance, lead-free nanoclusters. In work recently reported by the same group, Zou et al. synthesized hybrid copper iodide single crystals, Cu2I2(3,4-DMP)4 (3,4-DMP = 3,4-dimethylpyridine), with a PLQY of 80%, utilizing ionic liquids as both the solvent and iodine source, and nitrogen-containing organic ligands for structural modulation.123 Notably, the material overcomes the common issue of moisture-induced degradation, demonstrating exceptional stability over 90 days, thereby providing a novel solution for the inkjet printing of scintillators.
Table 3 The parameters of Cu-based cluster scintillators
Material Type Light yield (photons per MeV) Detection limit (nGyair s−1) Spatial resolution (lp mm−1) Ref.
Cu4I6(bttmpe)2 Cluster 17 117
Cu6I8(bu-ted)2 19.6 20 118
(IPP)CuI2 Cluster 75[thin space (1/6-em)]793.83 62.29 10.2 119
(C12H28N)2Cu2I4 Single crystal 56[thin space (1/6-em)]000 28.39 19.8 120
[DDPACDBFDP]2Cu4I4 Cluster 77 12 121
Cu4I4(C6H14N2)2 3D 48[thin space (1/6-em)]538 30.68 122
Cu2I2(3,4-DMP)4 0D 38[thin space (1/6-em)]031 106.71 10 123
(18-crown-6)2Na2(H2O)3Cu4I6 (CNCI) 0D 109[thin space (1/6-em)]000 59.4 16.3 124
(BzTPP)2Cu2I4 0D 27[thin space (1/6-em)]706 352 4.928 125
(CISDM)4[Cu4I8] 2H2O 2D 41[thin space (1/6-em)]042 86.8 108 126
(4-bzpy)4Cu4I4 0D 60[thin space (1/6-em)]948 155.2 5 127
TPA2Cu2I4 0D 40[thin space (1/6-em)]124 126 5.5 128
Cu4I6(L1)2 0D 32[thin space (1/6-em)]600 96.4 30 129
(MTP)2CuI3 (M1) 0D 6400 72.6 20 130
Cu6I8(bu-ted)2 32 17 131
1-rod MOF 41[thin space (1/6-em)]000 34.6 20 132
Cu4I4(R3As)3L Cluster 15[thin space (1/6-em)]000 18.1 133
Cu2X2 (X = Cl, Br and I) Nanoclusters 175[thin space (1/6-em)]000 30 134
[CuI(PPh3)2R] (R = PH, Cu-1 or PH-Br, Cu-2) Cuprous complex 49.7 6.8 135


To extend the scintillator properties, Wang et al. synthesized (18-crown-6)2Na2(H2O)3Cu4I6 (CNCI) with little self-absorption and near-unity green light emission.124 It possesses the highest values of LY ever reported, up to ∼109[thin space (1/6-em)]000 photons per MeV, and other impressive scintillator properties, like an ultralow detection limit of 59.4 nGy s−1 and an appreciable spatial resolution of 16.3 lp mm−1 for CNCI–polymer film screens, which are further improved after modification by a silicon wave guide structure to 24.8 lp mm−1. Furthermore, Lin et al. designed (BzTPP)2Cu2I4 (BzTPP refers to benzyltriphenylphosphonium) as a scintillator, exhibiting a substantial Stokes shift of 167 nm and 27[thin space (1/6-em)]706 photons per MeV, the energy transfer mechanism of which is shown in Fig. 5f. This material stands out as the premier choice among copper-based scintillators.125 Also, Yang et al. presented a 2D-copper-based cluster, (CISDM)4[Cu4I8] 2H2O (CISDM refers to cis-2,6-dimethylmorpholine), as a qualified scintillator material, with 87.2% PLQY at 588 nm.126 It also demonstrates a strong linear response to X-rays, with a LY of 41[thin space (1/6-em)]042 photons per MeV, a low detection limit of 86.8 nGy s−1, and an exceptional resolution of 108 lp mm−1. These characteristics represent a state-of-the-art performance for lead-free metal halide hybrid materials in spatial detection. Also, Kong et al. combined a conjugated organic cation (4-benzylpyridine, 4-bzpy) with copper(I) iodide modules to grow organic copper halide single crystals, namely 0D cubane-like (4-bzpy)4Cu4I4.127 It demonstrates a high LY of 60[thin space (1/6-em)]948 photons per MeV and an exceptional linear relationship with X-ray dose, originating from the ligand-to-core charge transfer state. Similarly, Lv et al. developed a copper halide scintillator, TPA2Cu2I4, utilizing the principle of STE ultra-broadband emission.128 The material achieves a remarkably high PLQY of 94.27%, a LY of 40[thin space (1/6-em)]124 MeV−1, and a spatial resolution of 5.5 lp mm−1, thereby providing new insights into broadband-emission X-ray scintillators.

Also, a nano-rod strategy was employed to promote the performance. He et al. chose copper iodide ink, Cu4I6(L1)2 (L1 refers to 1-propyl-1,4-diazabicyclo[2.2.2]octan-1-ium) 0D nanorods, as the scintillator, which exhibited a PLQY of 95.3% for broadband green emission, and the imaging result of this scintillator film is shown in Fig. 5g.129 The subsequent scintillator screen shows a low detection limit of 96.4 nGy s−1, approximately 55 times lower than that required for standard medical diagnosis (5.5 μGy s−1). Furthermore, the spatial resolution exceeding 30 lp mm−1 was twice that of conventional scintillators. To enhance the stability, Bohan Li et al. fabricated three excellent glass phase 0D hybrid copper halides, (MTP)2CuI3 (M1) (MTP refers to methyltriphenylphosphonium), (MTP)2Cu4I6-α (M2)′(MTP)2Cu4I6-β (M3), which shared the same inorganic anions.130 The prepared glass samples were heated in air to promote the crystal–glass phase transition, and the self-assembly enabled recrystallization, resulting in a hybrid bulk glass-ceramic; the procedure is illustrated by polarized optical microscopy (POM), as shown in Fig. 4h. It demonstrates an outstanding LY of 64[thin space (1/6-em)]000 photons per MeV, a detection limit of 72.6 nGy s−1 and high stability for real-time X-ray imaging with spatial resolution above 20 lp mm−1.130 Likewise, Gu et al. reported the synthesis of Cu6I8(bu-ted)2 via a surfactant-assisted method, limiting the crystal size with the help of controlling surface tension.131 The material shows a near-unity PLQY, resulting in a light output 4.8 times higher than that of the commercial scintillator Lu3Al5O12:Ce3+, as well as an impressively low detection limit of 32 nGy s−1 and a spatial resolution of 17 lp mm−1 while integrating SDBS into a flexible screen. Also, for the purpose of preparing a high quality, flexible and stable scintillator screen, Peng et al. selected a macrocyclic bridging ligand with an aggregation-induced emission feature as the MOF, which could endow the scintillator with high efficiency and preeminent stability, combined with a copper iodide cluster.132 The as-fabricated scintillator film exhibits exceptional flexibility, enabling their integration with flexible OLEDs for the precise imaging of complex surfaces or internal structures of objects. This capability holds significant potential for applications in industrial inspection and biological imaging. Moreover, the introduction of in situ polyvinyl pyrrolidone during the synthesis process leads to the regular rod-shaped microcrystal, enhancing the processibility even further. The scintillator realizes dynamic X-ray flexible imaging for the first time with a real time ultra-high spatial resolution of 20 lp mm−1. Also, Demyanov et al. presented cubane Cu4I4–triarsenic clusters [Cu4I4(R3As)3L] (R refers to the organic functional group, L refers to no ligand or nitrile) with a record ultralow detection limit of 18.1 nGy s−1 and an exquisite linear response over a wide range of X-ray doses (0–640 μGyair s−1).133 Yuan et al. reported their discovery that Cu2X2 (X = Cl, Br and I) was a TADF nano-cluster scintillator suffering from no mechanochromism and blessed with a considerable X-ray absorption cross section, which contributed to a flexible scintillator screen with excellent radiation and humidity resistance, an ultra-high LY of 175[thin space (1/6-em)]000 photons per MeV and a high spatial resolution of ∼30 lp mm−1.134 Besides, for the purpose of being eco-friendly, Chen reported a mechanochemical method for synthesizing [CuI(PPh3)2R] (R = PH, Cu-1 or PH-Br, Cu-2), circumventing the harm brought by diverse organic solvents in traditional chemical experiments.135 Cu-2 demonstrates 2.52 times higher RL intensity than that of the commercial BGO scintillator and 1.52 times larger than that of Cu-1 due to the introduction of PH-Br, which resulted in 37.75% higher PLQY.

Nanoclusters, emerging as a novel class of materials with dimensions intermediate between single atoms and quantum dots, exhibit near-unity PLQY and negligible Stokes shifts. However, key mechanistic aspects, such as the nature of halide ligand coordination dynamics, remain elusive,127 while their prevalent mechanochemical grinding synthesis protocols face scalability challenges compared to solution-based methods commonly employed for other luminescent materials.136

3.4 Rare-earth nanocrystalline scintillators

Rare-earth-doped scintillator materials demonstrate promising applications in nuclear medicine, radiographic imaging, encryption and particle detection. The high energy-transfer efficiency among rare-earth elements enhances the overall performance of the materials, yielding greater light output. Furthermore, doping with rare earth elements enhances the thermal and chemical stability of scintillator materials. Their unique properties also enable a wide range of luminous wavelengths, meeting the diverse needs of various applications.137

In recent efforts, the doping of rare-earth elements into perovskite materials has yielded significant progress. For instance, in 2024, Wang et al. reported a series of Cs2AgxNa1−xInyBi1−yCl6 halide scintillators doped with rare-earth ions.138 Among them, Cs2Ag0.6Na0.4In0.95Bi0.05Cl6:5%Tm has a high PLQY in NIR light of up to 76.3% and a LY of up to 33[thin space (1/6-em)]500 photons per MeV. The flexible scintillators achieved a high spatial resolution of 11.2 lp mm−1 and a low X-ray detection limit of 55.2 nGyair s−1. Moreover, after doping with Tm ions, they achieved clear imaging under bright external lighting by using an infrared camera with a 760 nm long-pass filter in front of the lens (Fig. 6a).


image file: d5qi00671f-f6.tif
Fig. 6 (a) X-ray images of encapsulated springs under various conditions were captured using CsPbBr3, Cs2Ag0.6Na0.4In0.95Bi0.05Cl6 (Tm-free), and Cs2Ag0.6Na0.4In0.95Bi0.05Cl6:5%Tm (Tm-doped) films. (b) Schematic diagram of writing X-ray imaging information and decryption by a layer of the perovskite nanocrystals. (c) Decay curve of the thermal-stimulated luminescence of NaLuF4:Gd3+ (20 mol%) at 314 nm was recorded at room temperature after X-ray excitation was stopped for 5 min. The film was then heated to 80 °C for 160 min. (d) Photograph of the false information EHBUT in the prepared multilayer hydrogel under UV light (365 nm). (e) Photograph of the encrypted information HBU in the multilayer hydrogel after exposure to X-rays. (f) Schematic representation of radio-luminescent functional building units for constructing a thermo-responsive MOF scintillator. (g) Bright field photograph of glass scintillators.

Doping rare-earth ions into fluoride-containing nano-scintillators is also a recognized approach.8,139–144 In 2023, Yang et al. reported high-security X-ray imaging encryption technology by developing an ultra-long RL memory film, namely NaLuF4:Gd3+ or Ce3+, which captured subtle pure UV emission characteristics.145 The encrypted X-ray imaging information can be securely stored in memory films for over seven days and can be optically decoded using perovskite nanocrystals (Fig. 6b and c). Moreover, the as-fabricated flexible memory film exhibits a high spatial resolution of 20 lp mm−1. The varying durations of fluorescence lifetime offer distinct advantages in different application contexts. To address the issue of strong hysteresis scintillation luminescence, Zhang et al. synthesized LiYF4:15Tb/25Gd@LiYF4 core/shell nano-scintillators to inhibit the generation of Frenkel defects.141 In the same year, Yang et al. reported NaLuF4:Gd/Tb/Eu doped with 5%Tb with a high LY of 15[thin space (1/6-em)]313 photons per MeV and a low detection limit of 84.1 nGyair s−1.146 Subsequently, Zhang et al. exhibited Tb3+-doped Na5Lu9F32 with good water dispersibility and highly sensitive luminescence to X-rays.147 Remarkably, they employed multilayer hydrogels for information camouflage and multilayer encryption. The encrypted information can only be identified through X-ray exposure, while false information is revealed under ultraviolet light (Fig. 6d and e). In 2024, Lu et al. proposed a type of heterovalent Cu2+ ion co-doped LiLuF4:Tb,Cu microcrystalline scintillation material with higher RL intensity, long afterglow and thermoluminescence intensity.148 The detection limit of LiLuF4:Tb,Cu can reach 2.7928 nGy s−1 and a spatial resolution of 22 lp mm−1@MTF = 0.2 is achieved.

In the field of scintillators materials, MOFs provide a tunable platform for the integration of X-ray absorption centers and luminescent chromophores in a dense framework manner and rare-earth-doped MOFs are widely studied recently, as it is in perovskite scintillators.149 In 2024, Li et al. proposed thermo-responsive lanthanide MOF scintillators (Tb0.95Eu0.05-BPTC), presenting a low dose rate detection limit of 156.1 nGyair s−1 (ref. 150) (Fig. 6f). Tb0.95Eu0.05-BPTC has a high relative LY of 39[thin space (1/6-em)]000 photons per MeV, imaging spatial resolution of 18 lp mm−1, good irradiation stability and giant color transformation visualization, benefiting the applications. Similarly, Jin et al. reported [1-ethyl-3-methylimidazolium]EuBr3MeOH as scintillators with a LY of 43[thin space (1/6-em)]000 ± 100 photons per MeV and achieved spatial resolution above 10 lp mm−1.151 Additionally, Yang et al. reported Eu-pba and Tb-pba, featuring excellent radiation, hydrolytic, and thermal stabilities. Using them as scintillators can result in a linear response to the X-ray dose rate with detection limits of 4.92 and 3.17 μGy s−1, respectively.152 Recently, Xu et al. firstly proposed organo-lanthanide scintillators that offered diverse emission colors from ultraviolet to near-infrared and enabled precise control of luminescence lifetimes from nanoseconds to microseconds.153 Their materials can efficiently capture the dark triplet excitons generated during the absorption of secondary X-rays, leading to the production of highly stable and efficient RL that surpasses those of commercialized scintillators. Among these scintillators, for instance, remarkably, Eu(NTA)3DPEPO (where NTA refers to 4,4,4-trifluoro-1-(2-naphthyl)-1,3-butanedione and DPEPO refers to bis(2-(diphenylphosphino)phenyl) ether oxide) showed a more than 1000-fold enhancement in RL compared with EuCl3 salts with an equal molar amount of Eu3+. In a recent study, Wang et al. reported the successful synthesis of a novel terbium-based MOF (Tb-MOF-1), the enhanced RL performance of which was attributed to the efficient sensitization effect of the HDOBPDC3− ligand.154 This material demonstrated a remarkably low X-ray detection limit of 1.71 μGy s−1, significantly surpassing conventional scintillation thresholds. Furthermore, the researchers developed a flexible composite scintillator screen through polymer matrix integration, which exhibited exceptional imaging capabilities with a spatial resolution of 7.7 lp mm−1, meeting clinical radiography requirements.

Rare-earth-doped glass scintillators have also attracted significant attention in recent years due to outstanding optical performances such as high transmittance (Fig. 6g) and their ability to absorb high energy radiation and convert the energy into visible light.155,156 In 2023, Sun et al. fabricated Tb3+-doped and Ce3+/Tb3+-co-doped oxyfluoride scintillator (C4T9) glass samples with a spatial resolution of 16 lp mm−1.157 Then in 2024, Li et al. synthesized Tb3+-doped oxyfluoride aluminosilicate glass scintillators with a spatial resolution of 20 lp mm−1 and higher RL of up to 224% compared to the Bi4Ge3O12 crystal.158 Subsequently, Dai et al. reported a transparent glass-ceramic scintillator by embedding Tb3+-doped NaLu2F7 nano-crystals in a glass matrix.159 Importantly, most decreases in performance, due to prolonged exposure to high doses of X-rays, can be completely restored by heat treatment at 350 °C for 30 min.

Additionally, several other rare-earth-doped scintillators have been reported.160,161 For instance, in 2023, Ling et al. prepared YF3:30%Gd3+/10%Tb3+ micro-particles with fine environmental stability and superb X-ray RL, which achieved an imaging spatial resolution of 16.8 lp mm−1.162 Subsequently, in 2024, Wang et al. reported zero-thermal-quenching Lu3Al5O12:Ce3+ transparent ceramics (0.5 mm) with excellent spatial resolution of up to 112 lp mm−1.163 In the same year, Yang et al. synthesized single crystals of rare earth ion-doped ternary chalcogenides, NaGaS2/Eu, with a low detection limit of 250 nGy s−1 and a spatial resolution of 13.2 lp mm−1.164 Doping with rare-earth elements imparts some materials with unexpectedly excellent properties, making the continued exploration of the potential of rare-earth-doped scintillators very promising. Furthermore, Wang et al. recently advanced the frontier of high-temperature scintillator technology through the development of a Tb3+-doped nanocrystalline glass composite system.165 This material possesses a record LY of 54[thin space (1/6-em)]900 photons per MeV and a sensitivity of 635.31 nGyair s−1. Remarkably, the material maintains exceptional thermal stability with 28.1 lp mm−1 spatial resolution at 500 °C and robust moisture resistance, surpassing current high-temperature scintillators.

The utilization of rare earth elements in optical materials has been extensively investigated (Table 4), as their intricate electronic configurations offer a compelling avenue for tailoring material band structures. However, the environmental and biological implications of these elements demand critical scrutiny, as exemplified by nephrotoxicity associated with gadolinium exposure, pulmonary fibrosis risks linked to samarium compounds,166 and inflammatory responses triggered by elevated cerium concentrations.167 Furthermore, the strikingly similar physicochemical properties among rare earth elements complicates purification processes, which objectively inflate raw material costs and ultimately constrain their large-scale industrial implementation.

Table 4 The parameters of rare-earth nanocrystalline scintillators
Material Type Light yield (photons per MeV) Detection limit (nGyair s−1) Spatial resolution (lp mm−1) Ref.
Cs2Ag0.6Na0.4In0.95Bi0.05Cl6:5%Tm Single crystals 33[thin space (1/6-em)]500 55.2 11.2 138
LiYF4:15Tb/25Gd@LiYF4 Nanocrystals 20 141
NaLuF4:Gd/Tb/Eu with 5%Tb-doping Nanocrystals 15[thin space (1/6-em)]313 84.1 8.7 146
Tb3+-doped Na5Lu9F32 Nanocrystals 15[thin space (1/6-em)]800 147
LiLuF4:Tb,Cu Microcrystals 2.7928 22 148
Tb0.95Eu0.05-BPTC MOFs 39[thin space (1/6-em)]000 156.1 18 150
[1-Ethyl-3-methylimidazolium]EuBr3MeOH Organic–inorganic hybrids 43[thin space (1/6-em)]000 405.4 >10 151
Eu-pba/Tb-pba Organic–inorganic hybrids 4920/3170 8/12.6 152
Eu(NTA)3DPEPO Organic–inorganic hybrids 38[thin space (1/6-em)]589 30.8 20 153
Ln-MOF [Tb(HDOBPDC)(DMF)(H2O)]n MOF 1710 7.7 154
SiBNaBaGd–xTb Glass 20 155
Ce3+-doped 55SiO2–20Al2O3–15SrF2–10NaF Glass 18 156
Ce3+/Tb3+-co-doped Glass 343.9 16 157
55SiO2–14.5SrF2–4.5SrO–15LuF3
10Al2O3–0.5Sb2O3−xTb4O7yCeF3 Glass 20 158
55SiO2–20Al2O3–15BaF2–10NaF–xTbF3
Tb3+-doped NaLu2F7 Nanocrystals 20 159
SrS:Ce3+/CaS:Ce3+ Nanocrystals 660   160
2BaO–3SiO2 doped with Eu2O3 Glass 26[thin space (1/6-em)]000 12 161
YF3:Gd3+/Tb3+ Microparticles 16.8 162
Lu3Al5O12:Ce3+ Ceramics 40[thin space (1/6-em)]200 112 163
NaGaS2/Eu Crystal 8188 250 13.2 164
KTb3−xGdxF10 Crystal 54[thin space (1/6-em)]900 635.31 28.1 165


3.5 Organic scintillators

In addition to the scintillators mentioned above, organic scintillators have garnered significant attention168–170 due to their rapid response times, diverse chemical structures, low processing temperatures, and cost-effectiveness (Table 5). During the luminescence phase, the charge carriers recombine to produce singlet states (25%) and triplet states (75%), resulting in light emission. However, triplet excitons may be lost due to the dark state characteristics of fluorescent organic scintillators.171 TADF and room-temperature phosphorescence (RTP) materials have garnered significant attention due to their ability to exhibit bright triplet excitons under X-ray excitation.172,173 Additionally, hybridized local and charge-transfer (HLCT) excited states in organic donor–acceptor systems can produce efficient PL through high-efficiency triplet–singlet reverse intersystem crossing (RISC) transitions, achieving nearly 100% utilization of excitons.174 To enhance efficiency and achieve rapid RISC, a common strategy involves increasing the spin–orbital coupling constant (SOC) in organic emitters by exploiting heavy-atom effects.175 Researchers have proposed innovative methods and applications based on these principles and have identified numerous new, high-performance purely organic scintillators.
Table 5 The parameters of organic scintillators
Material Light yield (photons per MeV) Detection limit (nGyair s−1) Spatial resolution (lp mm−1) Ref.
Py2TTz–I2F4 32[thin space (1/6-em)]583 70.49 26.8 168
(TPPcarz)2MnBr4 44[thin space (1/6-em)]600 32.42 169
Tetra(p-bromophenyl)ethene@poly(vinyltoluene) 14[thin space (1/6-em)]443 13.9 170
4,4′-Bis(9-carbazolyl)biphenyl 25.5 14.3 177
BPA-X (X = Cl, Br, I) 25.95 ± 2.49 20 178
Tp-Au-2 77[thin space (1/6-em)]600 16 179
10-(Pyridine-2-yl)-10H-phenothiazine 278 181
Sulfone-based organic molecules C6 16[thin space (1/6-em)]558 15 182
DCB@C[3]A 77.1 20 183
BiND-OMe 70.68 11 186
OMNI-PTZ 2 97 20 187
1,1,2,2-Tetrakis(4-bromophenyl)ethylene 34[thin space (1/6-em)]600 447 18.69 188


The RL performance of organic fluorescence scintillators could be enhanced by improving carrier transport capabilities.176 Following this guidance, in 2023, Chen et al. introduced a highly efficient organic fluorescent scintillator, 4,4′-bis(9-carbazolyl) biphenyl (CBP), which demonstrated a high PLQY of 61.92% and excellent carrier transport ability (Fig. 7a), and biphenyl (BP) showed an average hole mobility of 0.055 cm2 V−1 s−1 while CBP achieved a hole mobility of 0.094 cm2 V−1 s−1.177


image file: d5qi00671f-f7.tif
Fig. 7 (a) The PLQYs of BP, PCB, and CPB crystalline powders. (b) A reasonable mechanism of radioluminescence in BPA-X and PYI materials. (c) X-ray images of a standard X-ray test pattern plate of four scintillator screens with different thicknesses of 0.05, 0.2, 0.5, and 2 mm. (d) Organic phosphorescent scintillation processes dependent on polymorphism under X-ray irradiation. (e) Commission Internationale de l'Eclairage (CIE) coordinates calculated from the RL spectra of different materials mentioned in this study. (f) CIE chromaticity coordinate diagram of the RL colors of the host–guest materials. (g) Bright-field images and photographs under X-ray exposure of a fish and an electronic circuit board.

In the same year, Chen et al. proposed a molecular design strategy that incorporated supramolecular halogen bonding in fluorescence and room-temperature phosphorescence systems containing pyridine rings.178 Upon exposure to X-ray radiation, the introduction of halogen heavy atoms (X = Cl, Br, I) into 9,10-bis(4-pyridyl)anthracene (BPA) and the formation of halogen bonding interactions in bromophenyl–methylpyridinium iodide (PYI) crystals facilitate the generation of numerous hot electrons and deep holes (Fig. 7b). These charge carriers undergo a series of conversion and transport processes, including electron–electron scattering, Auger processes, and thermalization, resulting in a substantial number of low-energy carriers. Subsequently, these carriers recombined to form electron–hole pairs, leading to the production of 25% singlet and 75% triplet excitons, consistent with spin statistics. During the luminescence process, BPA-X crystals produced rapid fluorescent light emission, primarily utilizing 25% singlet excitons, while PYI crystals exhibited phosphorescent light emission, harnessing both singlet and triplet excitons. Similarly, Chen et al. utilized the strategy of incorporating heavy atoms in their latest publication.179 They developed a new category of organogold(III) complexes, namely Tp-Au-1 and Tp-Au-2, by adopting a through-space interaction motif to achieve high X-ray attenuation efficiency and efficient harvesting of triplet excitons for emission. Gold is a promising transition metal due to its significant spin–orbit coupling constant, derived from its large atomic number (Z = 79) and greater abundance compared to widely used metals like iridium or platinum.180 Consequently, neat films of Tp-Au-1 and Tp-Au-2 exhibit notable TADF properties with significant triplet–singlet charge transfer features, short decay lifetimes of 1.7–2.2 μs and a high PLQY of up to 77%. Remarkably, the emission mechanism of Tp-Au-1 and Tp-Au-2 can be tuned from TADF to phosphorescence by adjusting the polarity of the host matrix. Under X-ray irradiation, Tp-Au-2 demonstrated intense RL and a record-high scintillation LY of 77[thin space (1/6-em)]600 photons per MeV for organic scintillators. The resulting scintillator screens provided high-quality X-ray imaging with a spatial resolution exceeding 16.0 lp mm−1 (Fig. 7c).

In 2024, Dong et al. reported two phenothiazine derivatives exhibiting polymorphism-dependent phosphorescence RL.181 These derivatives demonstrate high radio stability and a low detection limit of 278 nGy s−1. The study not only introduced sulfur atoms to enhance X-ray absorption through the heavy atom effect but also confirmed that incorporating pyrimidine and pyridine groups containing heteroatoms into the phenothiazine skeleton promoted n–π* transitions, improving the intersystem crossing (ISC) process. Additionally, these groups stabilize triplet excitons, thereby enhancing room-temperature phosphorescence through multiple intermolecular interactions in the crystalline state (Fig. 7d). The experiments reveal that molecule stacking significantly affects the non-radiative decay of the triplet excitons of scintillators, which further determines the phosphorescence scintillation performance under X-ray irradiation. In 2024, Zhan et al. adopted a similar approach by developing several proof-of-concept sulfone-based organic molecules (C1–C7) with varying alkoxy chains to control molecular packing modes.182 Sulfone is a well-known electron acceptor unit for constructing organic emitters with TADF or RTP properties. This is primarily due to the n–π* transitions of its nonbonding electrons and the heavy atom effect of the sulfur atom (Z = 16), which enhances SOC. All the resulting molecules exhibited nearly identical PL properties, along with significant TADF and minimal reabsorption features in both dilute solutions and doped films. Among them, C6 demonstrates the highest RL with a maximum LY of 16[thin space (1/6-em)]558 photons per MeV under X-ray stimulation. Additionally, the rigid and flexible C6-based scintillator screens achieve high X-ray imaging resolutions of 15.0 and 10.6 lp mm−1, respectively.

However, heavy atoms could also lead to emission quenching. Consequently, novel strategies have been developed for the luminescence processes of organic scintillator systems by modifying the interactions between donors and acceptors. High-efficient organic TADF or RTP materials can be achieved by synthesizing supramolecular donor macrocycles and introducing acceptor molecules into the macrocycle's cavity, forming an intermolecular donor–acceptor system. In 2024, Zhang et al. synthesized the macrocyclic donor molecule calix[3]acridan (C[3]A), incorporating 1,2-dicyanobenzene (DCB) and 4-bromo-1,2-benzenedicarbonitrile (BrDCB) as guests within the macrocyclic cavities to produce macrocyclic co-crystals, specifically DCB@C[3]A and BrDCB@C[3]A. These co-crystals exhibit green (541 nm, DCB@C[3]A) and orange (633 nm, BrDCB@C[3]A) emissions, respectively (Fig. 7e).183 This demonstrates that adjusting the electron-donating ability of the macrocycle or the electron-accepting ability of the guest molecules can alter the scintillation color of the co-crystals, offering a simplified method for color tuning compared to covalently bonded scintillators. Furthermore, DCB@C[3]A shows the lowest detection limit of 77.1 nGy s−1. In 2024, Wang et al. presented a method for obtaining color-tunable organic scintillator materials, with emission wavelengths ranging from 520 to 682 nm by doping with different guest materials (Fig. 7f), suitable for X-ray imaging through host–guest doping strategy.184 Yang et al. prepared a series of p-terphenyl crystals doped with large-sized conjugated molecules (anthracene, tetracene, pentacene, chrysene, and picene) using the thermal field-elevating Bridgman method, resulting in tunable luminescence ranging from 360 to 680 nm under X-ray radiation.185

In addition to the methods mentioned above, some purely organic materials have been selected as scintillators due to their inherent properties. In 2024, Chen et al. extended the applications of rylene diimides to X-ray imaging, demonstrating superior scintillation properties with reduced non-radiative transitions.186 This improvement is attributed to the increased π–π plane distances, extended slip distances, and reduced π-plane overlap. The organic scintillator exhibits a low detection limit of 70.68 nGy s−1 and a short decay time of 4.37 ns, while achieving a high spatial resolution of 11.0 lp mm−1. Additionally, Zhang et al. demonstrated that the oxidized 1,8-naphthalimide-phenothiazine dyad (OMNI-PTZ 2) with HLCT-excited states showed an enhanced overlap integral between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) on MNI π-orbitals, as well as moderate donor–acceptor electron interactions.187 Consequently, OMNI-PTZ 2(G) exhibits significantly increased RL and greatly reduced decay lifetime, achieving an X-ray dose sensitivity of 97 nGy s−1 and an exceptional spatial resolution of 20 lp mm−1 (Fig. 7g). Furthermore, Du et al. developed a new family of hot exciton scintillators, which were characterized by a large energy gap between the high-lying and lowest triplet states and a small energy difference between triplet and singlet states.188 1,1,2,2-Tetrakis(4-bromophenyl) ethylene demonstrates an ultrafast radiative lifetime of 1.79 ns and a LY of approximately 34[thin space (1/6-em)]600 photons per MeV, showing an excellent combination of high LY and short decay time.

Organic materials, leveraging their structural diversity and tunability, have been extensively investigated for scintillation applications. However, reports on purely organic molecular scintillators remain relatively scarce compared to inorganic or hybrid organic–inorganic systems. This disparity stems from intrinsic limitations: spin-forbidden transitions impose fundamental constraints on LY optimization through molecular design alone, while their inherently low Zeff further limits their efficacy in high-energy radiation absorption.189

3.6 Specialized structure scintillators

The previously discussed advancements in scintillators focus on innovative, high-performance compounds with the goal of synthesizing even more superior scintillators. Furthermore, unique structures have been incorporated into the design of scintillator materials to enhance their performance.190–192 In this regard, Lin et al. proposed the concept of ‘structural engineering’, which, through internal or external design at scales ranging from micro to macro, offered a promising strategy that not only enhanced the performance of scintillators but also broadened their functionality in specific applications such as radiation imaging and therapy, opening up new avenues for candidate materials.193 Likewise, Roques-Carmes et al. proposed a first-principles theory of nanophotonic scintillators and demonstrated order-of-magnitude scintillation enhancements in two different platforms.194 Furthermore, the arraying of scintillators through specialized structures to enhance photon yield or resolution is a common approach. In the design of such structures, three primary objectives are typically pursued: the growth of high-quality scintillators of appropriate thickness, increasing the refractive index contrast between the scintillator and its matrix, and optimizing the photon structure within the scintillator array.195 In 2023, Dierks et al. reported that they synthesized CsPbBr3 nanowires in anodized aluminum oxide and achieved sufficient stability for X-ray micro-tomography under ambient conditions, as well as high spatial resolution.196 In the same year, Yi et al. proposed the design of double-tapered optical-fiber arrays, which could drastically increase the light-collection efficiency and imaging resolution (Fig. 8a).197 They achieved a spatial resolution of 22 lp mm−1 and pixelated gamma-ray imaging with a thickness of 4 mm and ∼10[thin space (1/6-em)]000 pixels under focused 6 MeV irradiation. Moreover, the dose rate (3.5 nGy s−1 to 96 mGy s−1) can be conveniently monitored. Then in 2024, considering that the absence of X-ray energy distinction resulted in insufficient image contrast, Ran et al. presented an innovative multi-energy X-ray linear-array detector based on Cs3Cu2I5.105 Benefiting from negligible self-absorption of the material and side-illuminated scintillation scenarios, the incident X-ray spectra could be reconstructed by analyzing the distribution of scintillator intensity while illuminated from the side, and the outcome error could be kept below 5.63%. Subsequently, Shao et al. fabricated a thick pixelated needle-like array scintillator capable of micrometer resolution (Fig. 8b).198 They achieved ultrahigh spatial resolutions of 60.8 lp mm−1 (the thickness of the scintillators is 0.5 mm) and 51.7 lp mm−1 (the thickness of the scintillators is 1 mm) at a MTF of 0.2 due to isolated light-crosstalk channels and robust light output. More recently, Song et al. proposed an anti-scattering CsPbBr3 scintillator array embedded within a polyurethane acrylate matrix, which could suppress light scattering and enhance the light collection efficiency by nearly two times compared to the monolithic film.199 Due to the large refractive index contrast between the scintillator and matrix, the scintillator array exhibits a low dosage and high resolution.
image file: d5qi00671f-f8.tif
Fig. 8 (a) Schematic representation of scintillator-encapsulated fiber arrays for the detection and imaging of X-rays and gamma rays. Within the fiber array, radioluminescence can propagate over several centimeters. Without the fiber array, however, the radioluminescence becomes attenuated and undetectable due to absorption and scattering within the scintillator layer. (b) Mechanisms of light propagation and the performance of conventional unstructured scintillators versus structured capillary needle-like array scintillators in X-ray imaging. (c) Bright-field (i), stacked (ii), decomposed (iii and iv), and color-reconstructed (v) dual-energy X-ray images of a moth, a beetle, and a leaf, utilizing the top–filter–bottom sandwich structure scintillator. (d) Comparison of traditional X-ray imaging approaches and chromatic X-ray imaging strategies.

In addition to the specialized structures mentioned above, dual-energy X-ray imaging (DEXI) is cutting-edge technology that provides more detailed information about specific materials compared to traditional single-energy X-ray imaging strategies. In 2023, Shao et al. demonstrated a top–filter–bottom sandwich structure scintillator for high-resolution DEXI within a single exposure, achieving an excellent resolution of approximately 18 lp mm−1 on stacked images (Fig. 8c).200 Subsequently, Hu et al. successfully achieved color recognition of objects with different densities by using the scintillators of mutually inactive (BA)2PbBr4:Mn and Cs3Cu2I5:TI, which contributed to the application of color X-ray imaging in real-world scenarios (Fig. 8d).201

4. Conclusion and outlook

In conclusion, we first provide a comprehensive overview of the various parameters and performance standards of scintillators to facilitate the assessment of their efficacy in recent studies. In recent years, the field of scintillators has garnered significant attention and experienced rapid development.

Lead-based perovskite scintillators exhibit exceptional potential for X-ray detection due to their tunable optoelectronic properties, high spatial resolution, and rapid responses. However, their practical application is hindered by intrinsic challenges: environmental and health concerns from lead toxicity, structural instability under exposure to moisture and oxygen, and luminescence degradation under operational stress. Recent strategies, such as polymer encapsulation, metal–organic framework confinement, and defect-passivating ligand systems, have improved the stability by suppressing ion migration and environmental degradation. Concurrently, elemental doping and lattice engineering enhance energy transfer efficiency while mitigating self-absorption. On the other hand, lead-free perovskite scintillators, particularly those that are copper-based, double perovskites, and manganese-based halides, have emerged as promising alternatives to their toxic lead-based counterparts, offering a balance of environmental sustainability and high-performance X-ray detection. For instance, copper-based systems, such as Cs3Cu2I5 and its derivatives, leverage zero-dimensional structures to localize STEs, suppressing non-radiative recombination while enabling tunable emission and flexible film integration. Innovations in doping, composite engineering, and morphology control further enhance their stability and LYs. Despite these advances, challenges persist, including insufficient carrier mobility in the double perovskites, long-term environmental degradation in certain hybrids, and scalability limitations in synthesis protocols. Future research should prioritize synergistic material designs—combining computational modeling with experimental validation—to engineer defect-tolerant lattices, optimize STE dynamics, and develop hydrophobic interfaces. Scalable fabrication techniques, such as inkjet printing and roll-to-roll processing, must be refined to produce large-area, flexible scintillator films for real-world applications. Additionally, exploring multifunctional systems that integrate radiation detection with stimuli-responsive properties could unlock novel applications in low-dose medical imaging, wearable diagnostics, and industrial nondestructive testing. By bridging material innovation with industrial compatibility, lead-free scintillators are poised to redefine radiation detection technologies, aligning high performance with ecological and operational sustainability.

Copper-based cluster scintillators, particularly those centered on CuxIx architectures, represent a groundbreaking advancement in X-ray detection technologies. These materials harness the unique advantages of STE mechanisms, tunable organic–inorganic hybrid frameworks, and efficient radiative recombination, enabling the development of high-performance, low-toxicity scintillators. By optimizing ligand–core charge transfer states and structural engineering, they achieve near-unity PLQYs, ultralow detection limits, and exceptional spatial resolution, while demonstrating remarkable flexibility and environmental stability. Innovations in green synthesis, such as ionic liquid solvents and mechanochemical approaches, further promote lead-free, energy-efficient fabrication processes, laying the foundation for scalable applications. Nevertheless, critical challenges persist. Moisture and oxygen sensitivity in certain materials compromise long-term stability and self-absorption effects and intricate synthesis protocols hinder large-scale production. Future research is expected to prioritize multifunctional solutions: enhancing stability through hydrophobic ligand design, glass-ceramic encapsulation, or MOF-based confinement; leveraging high-throughput computational models to refine STE energy levels and carrier transport dynamics; and advancing scalable manufacturing techniques like inkjet printing and roll-to-roll processing to bridge laboratory innovations with industrial deployment. Furthermore, it is necessary to increase the solubility of materials to reduce the scattering problem of scintillator films. The transformative potential of copper-based nanoclusters lies in their synergistic integration of performance, flexibility and sustainability. Their application in low-dose medical imaging, industrial nondestructive testing, and deep-space radiation monitoring holds immense promise.

Rare-earth-doped scintillator materials have garnered considerable attention as next-generation candidates for radiation detection and imaging applications. The introduction of rare-earth ions, acting as efficient luminescence centers, markedly enhances energy transfer efficiency and scintillation yield, owing to their distinctive 4f–5d electronic transitions. This quantum efficiency improvement translates to superior radiation sensitivity and spectral resolution, which are particularly crucial for low-dose radiography and nuclear medicine diagnostics. The incorporation of rare-earth dopants not only improves radiation hardness but also enhances thermal stability through optimized lattice interactions, ensuring operational reliability in extreme environments ranging from cryogenic detectors to aerospace systems. Furthermore, the engineered energy level configurations of rare-earth ions enable tunable emission spectra spanning near-UV to near-IR wavelengths, facilitating spectral matching with various photodetectors. It is recommended to strengthen mechanistic investigations into multivalent rare-earth activator configurations to elucidate quantitative structure–property relationships between complex energy level structures and optoelectronic characteristics, thereby formulating design principles for targeted applications including medical imaging and particle physics instrumentation. Moreover, exploring longer fluorescence lifetime modulation techniques expands innovative implementations in emerging optoelectronics such as temporal information encryption and anti-counterfeiting. A multidisciplinary approach integrating multiscale simulations with advanced characterization methodologies should be prioritized to overcome current limitations in understanding structure–performance correlations of rare-earth scintillators. It is crucial to emphasize that research on lanthanum-doped scintillator materials requires the systematic reporting of comprehensive performance metrics. The prevalent practice of overemphasizing isolated parameters while neglecting other critical performance indicators constitutes a detrimental research tendency that could impede the healthy development of this field.

Organic scintillators have undergone transformative advancements through strategic manipulation of excited-state dynamics and molecular architectures, positioning them as viable alternatives to conventional inorganic systems. Recent breakthroughs in TADF and RTP materials have effectively addressed the intrinsic limitations of triplet exciton utilization in organic systems. Some innovations, such as introducing heavy-atom effects to amplify spin–orbit coupling, achieve record-breaking LYs while enabling tunable emission mechanisms between TADF and phosphorescence. Despite these achievements, critical challenges persist, including radiation-induced degradation, oxygen quenching of triplet states, and scalability limitations arising from batch-to-batch performance variations. Looking ahead, the field must prioritize the development of hybrid organic–inorganic architectures to reconcile the low-Z limitations of organic matrices with the high stopping power of embedded heavy-element nanostructures. Furthermore, engineering environmentally robust systems, such as oxygen-insensitive RTP scintillators via steric encapsulation or crosslinked polymer networks, will be essential for real-world deployment in medical imaging and aerospace environments. To bridge the gap between laboratory innovation and industrial adoption, concerted efforts must focus on establishing standardized benchmarking protocols and elucidating universal structure–property relationships governing triplet harvesting efficiency. Success in these endeavors could redefine performance paradigms across radiation detection technologies, ultimately enabling organic scintillators to transcend niche applications and compete at the forefront of high-energy physics, next-generation radiography, and portable radiation monitoring systems.

Beyond the development of high-performance scintillators, employing specialized structures to enhance the performance of scintillators has also emerged as a highly viable solution. The integration of scintillator materials with matrix materials featuring unique structures can significantly elevate the LY, stability, and detection resolution of the scintillators. This approach further enhances the potential of scintillators for practical applications. Moreover, DEXI offers a remarkably innovative and promising direction for the advancement of X-ray detection. This design provides the possibility of color recognition between objects of varying densities, thereby promoting the advancement of chromatic X-ray imaging in practical applications. However, studies on the impact of structural factors on optical performance have not been sufficiently developed. Future studies should prioritize the exploration of innovative optical architectures in X-ray scintillator systems, with particular emphasis on harnessing micro–nano-structural engineering to enhance photonic characteristics, thereby advancing scintillation performance and diversifying potential applications across radiation detection domains.

In summary, the shared purposes are to develop scintillators with flexibility, high LY, excellent stability, and high resolution. We aspire for this review to enhance researchers’ understanding of the overall progress in scintillator development, serving as a foundation for the creation of materials with superior performance, and more importantly, offering recommendations for future research directions. The primary purpose of this review is to revisit and update recent advancements in scintillators, while offering reliable recommendations for future developments in this field.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This research was funded by the National Natural Science Foundation of China (No. 62274024), the Natural Science Foundation Program of Sichuan Province (No. 2024NSFSC0232), a Key Teacher Start-Up Research Grant from UESTC (No. Y030202059018069) and the Xiaomi Young Scholars Fellowship.

References

  1. P. Büchele, M. Richter, S. F. Tedde, G. J. Matt, G. N. Ankah, R. Fischer, M. Biele, W. Metzger, S. Lilliu, O. Bikondoa, J. E. Macdonald, C. J. Brabec, T. Kraus, U. Lemmer and O. Schmidt, X-ray imaging with scintillator-sensitized hybrid organic photodetectors, Nat. Photonics, 2015, 9, 843–848 CrossRef.
  2. X. Xu, W. Qian, S. Xiao, J. Wang, S. Zheng and S. Yang, Halide perovskites: A dark horse for direct X–ray imaging, EcoMat, 2020, 2, e12064 CrossRef CAS.
  3. M. J. Yaffe and J. A. Rowlands, X-ray detectors for digital radiography, Phys. Med. Biol., 1997, 42, 1–39 CrossRef CAS PubMed.
  4. B. G. M. Durie and S. E. Salmon, High Speed Scintillation Autoradiography, Science, 1976, 190, 1093–1095 CrossRef PubMed.
  5. P. Lecoq, Development of new scintillators for medical applications, Nucl. Instrum. Methods Phys. Res., Sect. A, 2016, 809, 130–139 CrossRef CAS.
  6. W. C. Röntgen, On a New Kind of Rays, Science, 1896, 3, 227–231 CrossRef.
  7. Q. Ma, Y. Cao, X. Ge, Z. Zhang, S. Gao and J. Song, X–Ray Excited Luminescence Materials for Cancer Diagnosis and Theranostics, Laser Photonics Rev., 2023, 18, 2300565 CrossRef.
  8. Z. Hong, Z. Chen, Q. Chen and H. Yang, Advancing X-ray Luminescence for Imaging, Biosensing, and Theragnostics, Acc. Chem. Res., 2022, 56, 37–51 CrossRef PubMed.
  9. H. Hu, G. Niu, Z. Zheng, L. Xu, L. Liu and J. Tang, Perovskite semiconductors for ionizing radiation detection, EcoMat, 2022, 4, e12258 CrossRef CAS.
  10. M. Nikl and A. Yoshikawa, Recent R&D Trends in Inorganic Single–Crystal Scintillator Materials for Radiation Detection, Adv. Opt. Mater., 2015, 3, 463–481 CrossRef CAS.
  11. J. D. Valentine, D. K. Wehe, G. F. Knoll and C. E. Moss, Temperature dependence of CsI (T1) absolute scintillation yield, IEEE Trans. Nucl. Sci., 1993, 40, 1267–1274 CAS.
  12. W. Mengesha, T. D. Taulbee, B. D. Rooney and J. D. Valentine, Light yield nonproportionality of CsI(Tl), CsI(Na), and YAP, IEEE Trans. Nucl. Sci., 1998, 45, 456–461 CAS.
  13. H. Grassmann, E. Lorenz and H.-G. Moser, Properties of CsI(TI)—Renaissance of an old scintillation material, Nucl. Instrum. Methods Phys. Res., Sect. A, 1985, 228, 323–326 CrossRef CAS.
  14. M. K. a, J. P. b and M. Moszyński, Comparison of YAP and BGO for high-resolution PET detectors, Nucl. Instrum. Methods Phys. Res., Sect. A, 1998, 404, 413–417 CrossRef.
  15. M. J. Weber and R. R. Monchamp, Luminescence of Bi4Ge3O12: Spectral and decay properties, J. Appl. Phys., 1973, 44, 5495–5499 CrossRef CAS.
  16. Y. Zhou, J. Chen, O. M. Bakr and O. F. Mohammed, Metal Halide Perovskites for X-ray Imaging Scintillators and Detectors, ACS Energy Lett., 2021, 6, 739–768 CrossRef CAS.
  17. C. Dujardin, E. Auffray, E. Bourret-Courchesne, P. Dorenbos, P. Lecoq, M. Nikl, A. N. Vasil'ev, A. Yoshikawa and R. Y. Zhu, Needs, Trends, and Advances in Inorganic Scintillators, IEEE Trans. Nucl. Sci., 2018, 65, 1977–1997 CAS.
  18. B. K. Cha, J. Y. Kim, T. J. Kim, C. Sim and G. Cho, Fabrication and imaging characterization of high sensitive CsI(Tl) and Gd2O2S(Tb) scintillator screens for X-ray imaging detectors, Radiat. Meas., 2010, 45, 742–745 CrossRef CAS.
  19. M. Laval, M. Moszyński, R. Allemand, E. Cormoreche, P. Guinet, R. Odru and J. Vacher, Barium fluoride—Inorganic scintillator for subnanosecond timing, Nucl. Instrum. Methods Phys. Res., 1983, 206, 169–176 CrossRef CAS.
  20. C. D. Brandle, Czochralski growth of oxides, J. Cryst. Growth, 2004, 264, 593–604 CrossRef CAS.
  21. L. Qin, H. Li, S. Lu, D. Ding and G. Ren, Growth and characteristics of LYSO (Lu2(1−x−y)Y2xSiO5:Cey) scintillation crystals, J. Cryst. Growth, 2005, 281, 518–524 CrossRef CAS.
  22. M. J. Weber, Inorganic scintillators: today and tomorrow, J. Lumin., 2002, 100, 35–45 CrossRef CAS.
  23. F. Maddalena, L. Tjahjana, A. Xie, Arramel, S. Zeng, H. Wang, P. Coquet, W. Drozdowski, C. Dujardin, C. Dang and M. D. Birowosuto, Inorganic, Organic, and Perovskite Halides with Nanotechnology for High–Light Yield X- and γ-ray Scintillators, Crystals, 2019, 9, 88 CrossRef.
  24. Q. Chen, J. Wu, X. Ou, B. Huang, J. Almutlaq, A. A. Zhumekenov, X. Guan, S. Han, L. Liang, Z. Yi, J. Li, X. Xie, Y. Wang, Y. Li, D. Fan, D. B. L. Teh, A. H. All, O. F. Mohammed, O. M. Bakr, T. Wu, M. Bettinelli, H. Yang, W. Huang and X. Liu, All-inorganic perovskite nanocrystal scintillators, Nature, 2018, 561, 88–93 CrossRef CAS PubMed.
  25. Y. Wang, M. Li, Z. Chai, Y. Wang and S. Wang, Perovskite Scintillators for Improved X–ray Detection and Imaging, Angew. Chem., Int. Ed., 2023, 62, e202304638 CrossRef CAS PubMed.
  26. Q. Zhou, W. Li, J. Xiao, A. Li and X. Han, Low–Dimensional Metal Halide for High Performance Scintillators, Adv. Funct. Mater., 2024, 34, 2402902 CrossRef CAS.
  27. J. Fan, H. Li, W. Liu and G. Ouyang, Fabrication Strategies of Mn2+-Based Scintillation Screens for X-Ray Detection and Imaging, Angew. Chem., Int. Ed., 2025, 64, e202425661 CrossRef CAS PubMed.
  28. W. Zhao, X. Huang, S. Hu, F. Yang and J. Zhong, Rare-earth nanocrystalline scintillators for biomedical application: A review, Ceram. Int., 2024 DOI:10.1016/j.ceramint.2024.08.149.
  29. C. Xie, P. You, Z. Liu, L. Li and F. Yan, Ultrasensitive broadband phototransistors based on perovskite/organic-semiconductor vertical heterojunctions, Light: Sci. Appl., 2017, 6, e17023–e17023 CrossRef PubMed.
  30. F. Zhou, Z. Li, W. Lan, Q. Wang, L. Ding and Z. Jin, Halide Perovskite, a Potential Scintillator for X-Ray Detection, Small Methods, 2020, 4, 2000506 CrossRef CAS.
  31. J. H. Heo, D. H. Shin, J. K. Park, D. H. Kim, S. J. Lee and S. H. Im, High-Performance Next-Generation Perovskite Nanocrystal Scintillator for Nondestructive X-Ray Imaging, Adv. Mater., 2018, 30, 1801743 CrossRef PubMed.
  32. D. Rutstrom, L. Stand, M. Koschan, C. L. Melcher and M. Zhuravleva, Europium concentration effects on the scintillation properties of Cs4SrI6:Eu and Cs4CaI6:Eu single crystals for use in gamma spectroscopy, J. Lumin., 2019, 216, 116740 CrossRef CAS.
  33. Y. Zhang, R. Sun, X. Ou, K. Fu, Q. Chen, Y. Ding, L.-J. Xu, L. Liu, Y. Han, A. V. Malko, X. Liu, H. Yang, O. M. Bakr, H. Liu and O. F. Mohammed, Metal Halide Perovskite Nanosheet for X-ray High-Resolution Scintillation Imaging Screens, ACS Nano, 2019, 13, 2520–2525 CrossRef CAS PubMed.
  34. A. Shinde, R. Gahlaut and S. Mahamuni, Low-Temperature Photoluminescence Studies of CsPbBr 3 Quantum Dots, J. Phys. Chem. C, 2017, 121, 14872–14878 CrossRef CAS.
  35. S. M. Ferro, M. Wobben and B. Ehrler, Rare-earth quantum cutting in metal halide perovskites – a review, Mater. Horiz., 2021, 8, 1072–1083 RSC.
  36. T. Cai, J. Wang, W. Li, K. Hills-Kimball, H. Yang, Y. Nagaoka, Y. Yuan, R. Zia and O. Chen, Mn(2+)/Yb(3+) Codoped CsPbCl(3) Perovskite Nanocrystals with Triple-Wavelength Emission for Luminescent Solar Concentrators, Adv. Sci., 2020, 7, 2001317 CrossRef CAS PubMed.
  37. L. Zi, J. Song, N. Wang, T. Wang, W. Li, H. Zhu, W. Xu and H. Song, X–Ray Quantum Cutting Scintillator Based on CsPbClx,Br3−x:Yb3+ Single Crystals, Laser Photonics Rev., 2023, 17, 2200852 CrossRef CAS.
  38. X. Wu, Z. Guo, S. Zhu, B. Zhang, S. Guo, X. Dong, L. Mei, R. Liu, C. Su and Z. Gu, Ultrathin, Transparent, and High Density Perovskite Scintillator Film for High Resolution X–Ray Microscopic Imaging, Adv. Sci., 2022, 9, 2200831 CrossRef CAS.
  39. S. Chen, W. Liu, M. Xu, P. Shi and M. Zhu, Electrospray prepared flexible CsPbBr 3 perovskite film for efficient X-ray detection, J. Mater. Chem. C, 2023, 11, 8431–8437 RSC.
  40. A. Erroi, F. Carulli, F. Cova, I. Frank, M. L. Zaffalon, J. Llusar, S. Mecca, A. Cemmi, I. Di Sarcina, F. Rossi, L. Beverina, F. Meinardi, I. Infante, E. Auffray and S. Brovelli, Ultrafast Nanocomposite Scintillators Based on Cd-Enhanced CsPbCl 3 Nanocrystals in Polymer Matrix, ACS Energy Lett., 2024, 9, 2333–2342 CrossRef CAS.
  41. B. Zhang, G. Lyu, E. A. Kelly and R. C. Evans, Forster Resonance Energy Transfer in Luminescent Solar Concentrators, Adv. Sci., 2022, 9, e2201160 CrossRef PubMed.
  42. H. Dierks, Z. Zhang, N. Lamers and J. Wallentin, 3D X-ray microscopy with a CsPbBr3 nanowire scintillator, Nano Res., 2023, 16, 1084–1089 CrossRef CAS.
  43. N. Z. Galunov, I. F. Khromiuk and O. A. Tarasenko, Features of pulse shape discrimination capability of organic heterogeneous scintillators, Nucl. Instrum. Methods Phys. Res., Sect. A, 2020, 949, 162870 CrossRef CAS.
  44. K. Kagami, M. Koshimizu, Y. Fujimoto, S. Kishimoto, R. Haruki, F. Nishikido and K. Asai, High-energy X-ray detection capabilities of Hf-loaded plastic scintillators synthesized by sol–gel method, J. Mater. Sci.: Mater. Electron., 2020, 31, 896–902 CrossRef CAS.
  45. F. Cao, X. Xu, D. Yu and H. Zeng, Lead-free halide perovskite photodetectors spanning from near-infrared to X-ray range: a review, Nanophotonics, 2021, 10, 2221–2247 CrossRef CAS.
  46. G. C. Papavassiliou, Optical properties of small inorganic and organic metal particles, Prog. Solid State Chem., 1979, 12, 185–271 CrossRef CAS.
  47. M. D. Birowosuto, D. Cortecchia, W. Drozdowski, K. Brylew, W. Lachmanski, A. Bruno and C. Soci, X-ray Scintillation in Lead Halide Perovskite Crystals, Sci. Rep., 2016, 6, 37254 CrossRef CAS PubMed.
  48. P. Y. D. Maulida, S. Hartati, D. Kowal, L. J. Diguna, M. A. Kuddus Sheikh, M. H. Mahyuddin, I. Mulyani, D. Onggo, F. Maddalena, A. Bachiri, M. E. Witkowski, M. Makowski, W. Drozdowski, Arramel and M. D. Birowosuto, Organic Chain Length Effect on Trap States of Lead Halide Perovskite Scintillators, ACS Appl. Energy Mater., 2023, 6, 5912–5922 CrossRef CAS.
  49. H. Chen, M. Lin, Y. Zhu, D. Zhang, J. Chen, Q. Wei, S. Yuan, Y. Liao, F. Chen, Y. Chen, M. Lin and X. Fang, Halogen–bonding boosting the high performance X–ray imaging of organic scintillators, Small, 2024, 20, 2307277 CrossRef CAS.
  50. D. Yu, P. Wang, F. Cao, Y. Gu, J. Liu, Z. Han, B. Huang, Y. Zou, X. Xu and H. Zeng, Two-dimensional halide perovskite as β-ray scintillator for nuclear radiation monitoring, Nat. Commun., 2020, 11, 3395 CrossRef CAS.
  51. W. Shao, X. Wang, Z. Zhang, J. Huang, Z. Han, S. Pi, Q. Xu, X. Zhang, X. Xia and H. Liang, Highly Efficient and Flexible Scintillation Screen Based on Manganese(II) Activated 2D Perovskite for Planar and Nonplanar High–Resolution X–Ray Imaging, Adv. Opt. Mater., 2022, 10, 2102282 CrossRef CAS.
  52. M. Xia, Z. Xie, H. Wang, T. Jin, L. Liu, J. Kang, Z. Sang, X. Yan, B. Wu, H. Hu, J. Tang and G. Niu, Sub–Nanosecond 2D Perovskite Scintillators by Dielectric Engineering, Adv. Mater., 2023, 35, 2211769 CrossRef CAS.
  53. E. Mosconi, J. M. Azpiroz and F. De Angelis, Ab Initio Molecular Dynamics Simulations of Methylammonium Lead Iodide Perovskite Degradation by Water, Chem. Mater., 2015, 27, 4885–4892 CrossRef CAS.
  54. A. Jana, S. Cho, S. A. Patil, A. Meena, Y. Jo, V. G. Sree, Y. Park, H. Kim, H. Im and R. A. Taylor, Perovskite: Scintillators, direct detectors, and X-ray imagers, Mater. Today, 2022, 55, 110–136 CrossRef CAS.
  55. F. Cao, D. Yu, W. Ma, X. Xu, B. Cai, Y. M. Yang, S. Liu, L. He, Y. Ke, S. Lan, K.-L. Choy and H. Zeng, Shining Emitter in a Stable Host: Design of Halide Perovskite Scintillators for X-ray Imaging from Commercial Concept, ACS Nano, 2020, 14, 5183–5193 CrossRef CAS PubMed.
  56. B. Wang, J. Peng, X. Yang, W. Cai, H. Xiao, S. Zhao, Q. Lin and Z. Zang, Template Assembled Large–Size CsPbBr 3 Nanocomposite Films toward Flexible, Stable, and High–Performance X–Ray Scintillators, Laser Photonics Rev., 2022, 16, 2100736 CrossRef CAS.
  57. H. Lv, Q. Hao, N. Yan, L. Ma and M. Chen, Large-area in situ growth of a flexible perovskite scintillator film for X-ray indirect detection applications, J. Mater. Chem. C, 2024, 12, 8970–8976 RSC.
  58. W. Zheng, H. Zhang, X. Wang, X. Zhang, T. Long, H. Wang, W. W. Yu and C. Zhou, Dual Function Polymer Ligands of Perovskite Nanocrystals for Extraordinary Water Resistance and X–Ray Imaging Scintillators, Adv. Opt. Mater., 2023, 12, 2301241 CrossRef.
  59. J. Shen, R. Jia, Y. Hu, W. Zhu, K. Yang, M. Li, D. Zhao, J. Shi and J. Lian, Cold-Sintered All-Inorganic Perovskite Bulk Composite Scintillators for Efficient X-ray Imaging, ACS Appl. Mater. Interfaces, 2024, 16, 24703–24711 CrossRef CAS PubMed.
  60. J. Chen, G. Jiang, E. Hamann, H. Mescher, Q. Jin, I. Allegro, P. Brenner, Z. Li, N. Gaponik, A. Eychmuller and U. Lemmer, Organosilicon-Based Ligand Design for High-Performance Perovskite Nanocrystal Films for Color Conversion and X-ray Imaging, ACS Nano, 2024, 18, 10054–10062 CrossRef CAS PubMed.
  61. H. Wu, P. Ran, L. Yao, H. Cai, W. Cao, Y. Cui, Y. Michael Yang, D. Yang and G. Qian, Confinement of methylammonium lead bromide nanocrystals in metal–organic frameworks as a stable scintillator for high-performance X-ray imaging, Chem. Eng. J., 2024, 491, 152098 CrossRef CAS.
  62. R. Li, W. Zhu, H. Wang, Y. Jiao, Y. Gao, R. Gao, R. Wang, H. Chao, A. Yu and X. Liu, Ultrastable and flexible glass−ceramic scintillation films with reduced light scattering for efficient X−ray imaging, npj Flexible Electron., 2024, 8, 31 CrossRef CAS.
  63. Z. Xiang, H. Cao, Y. Wei, W. Wang and X. Ouyang, Bivalent metal ion co-doped methylammonium lead tribromide perovskite nanocrystal scintillator for X-ray imaging, Chem. Eng. J., 2024, 495, 153432 CrossRef CAS.
  64. J. Qiu, H. Zhao, Z. Mu, J. Chen, H. Gu, C. Gu, G. Xing, X. Qin and X. Liu, Turning Nonemissive CsPb2Br5 Crystals into High-Performance Scintillators through Alkali Metal Doping, Nano Lett., 2024, 24, 2503–2510 CrossRef CAS.
  65. R. Subagyo, M. Eid, N. Safitri, R. Hanifah, A. A. Afkauni, L. Zhang, M. Makowski, M. A. K. Sheikh, M. H. Mahyuddin, D. Kowal, M. Witkowski, W. Drozdowski, M. D. Birowosuto, A. Arramel and Y. Kusumawati, Liquid-based cationic ligand engineering in one-dimensional bismuth bromide perovskites: A-site influence on scintillation properties, J. Mater. Chem. C, 2025 10.1039/D4TC04390A.
  66. J. A. Lai, P. Wang, B. Zheng, T. Xuan, D. Wu, Z. Wang, Y. Wang, W. Zhang, J. Du, P. He, K. An and X. Tang, Enhanced Performance in Cesium Tellurium Chlorine by Hafnium Alloying for X–Ray Computed Tomography Imaging, Adv. Opt. Mater., 2024, 12, 2303297 CrossRef CAS.
  67. F. Zhang, Y. Zhou, Z. Chen, X. Niu, H. Wang, M. Jia, J. Xiao, X. Chen, D. Wu, X. Li, Z. Shi and C. Shan, Large–Area X–Ray Scintillator Screen Based on Cesium Hafnium Chloride Microcrystals Films with High Sensitivity and Stability, Laser Photonics Rev., 2023, 17, 2200848 CrossRef CAS.
  68. W. Ke and M. G. Kanatzidis, Prospects for low-toxicity lead-free perovskite solar cells, Nat. Commun., 2019, 10, 965 CrossRef.
  69. T. Cai, W. Shi, S. Hwang, K. Kobbekaduwa, Y. Nagaoka, H. Yang, K. Hills-Kimball, H. Zhu, J. Wang, Z. Wang, Y. Liu, D. Su, J. Gao and O. Chen, Lead-Free Cs(4)CuSb(2)Cl(12) Layered Double Perovskite Nanocrystals, J. Am. Chem. Soc., 2020, 142, 11927–11936 CrossRef CAS PubMed.
  70. M. H. Miah, M. U. Khandaker, M. Aminul Islam, M. Nur-E-Alam, H. Osman and M. H. Ullah, Perovskite materials in X-ray detection and imaging: recent progress, challenges, and future prospects, RSC Adv., 2024, 14, 6656–6698 RSC.
  71. H. Tang, Y. Xu, X. Hu, Q. Hu, T. Chen, W. Jiang, L. Wang and W. Jiang, Lead-Free Halide Double Perovskite Nanocrystals for Light-Emitting Applications: Strategies for Boosting Efficiency and Stability, Adv. Sci., 2021, 8, 2004118 CrossRef CAS.
  72. J. C. Dahl, W. T. Osowiecki, Y. Cai, J. K. Swabeck, Y. Bekenstein, M. Asta, E. M. Chan and A. P. Alivisatos, Probing the Stability and Band Gaps of Cs2AgInCl6 and Cs2AgSbCl6 Lead-Free Double Perovskite Nanocrystals, Chem. Mater., 2019, 31, 3134–3143 CrossRef CAS.
  73. S. Ghosh, H. Shankar and P. Kar, Recent developments of lead-free halide double perovskites: a new superstar in the optoelectronic field, Mater. Adv., 2022, 3, 3742–3765 RSC.
  74. N. Varnakavi, R. Rajavaram, K. Gupta, P. R. Cha and N. Lee, Scintillation Performance of Mn(II)–Doped Cs2NaBiCl6 Double Perovskite Nanocrystals for X–Ray Imaging Applications, Adv. Opt. Mater., 2024, 12, 2301868 CrossRef CAS.
  75. N. J. Cherepy, R. D. Sanner, P. R. Beck, E. L. Swanberg, T. M. Tillotson, S. A. Payne and C. R. Hurlbut, Bismuth- and lithium-loaded plastic scintillators for gamma and neutron detection, Nucl. Instrum. Methods Phys. Res., Sect. A, 2015, 778, 126–132 CrossRef CAS.
  76. H. Chen, Q. Han, H. Qin, Y. Shen, H. Lv, Y. Liu, L. Du, Y. Wang, Y. He and W. Ning, Highly stable bismuth-chloride perovskite X-ray direct detectors with an ultralow detection limit, Chem. Sci., 2025, 16, 4768–4774 RSC.
  77. J. Li, Q. Hu, J. Xiao and Z. G. Yan, High-stability double perovskite scintillator for flexible X-ray imaging, J. Colloid Interface Sci., 2024, 671, 725–731 CrossRef CAS.
  78. V. Naresh, P.-R. Cha and N. Lee, Cs2NaGdCl6:Tb3+–A Highly Luminescent Rare-Earth Double Perovskite Scintillator for Low-Dose X-ray Detection and Imaging, ACS Appl. Mater. Interfaces, 2024, 16, 19068–19080 CrossRef CAS PubMed.
  79. M. Wang, X. Qing, T. Du, C. Wu and X. Han, Te4+-doped Cs2SnCl6 scintillator for flexible and efficient X-ray imaging screens, J. Mater. Chem. C, 2024, 12, 2241–2246 RSC.
  80. A. Datta, J. Fiala and S. Motakef, 2D perovskite-based high spatial resolution X-ray detectors, Sci. Rep., 2021, 11, 22897 CrossRef CAS PubMed.
  81. C.-F. Wang, C.-L. Fang, Y. Feng, S.-Y. Liu, Y. Zhang, H.-Y. Ye, L.-P. Miao and L. Liu, One-dimensional lead-free perovskite single crystals with high X-ray response grown by liquid phase diffusion, J. Mater. Chem. C, 2023, 11, 134–140 RSC.
  82. M. Zhou, H. Jiang, T. Hou, S. Hou, J. Li, X. Chen, C. Di, J. Xiao, H. Li and D. Ju, Inch-size and thickness-adjustable hybrid manganese halide single-crystalline films for high resolution X-ray imaging, Chem. Eng. J., 2024, 490, 151823 CrossRef CAS.
  83. L. Liu, H. Hu, W. Pan, H. Gao, J. Song, X. Feng, W. Qu, W. Wei, B. Yang and H. Wei, Robust Organogel Scintillator for Self-healing and Ultra-flexible X-ray Imaging, Adv. Mater., 2024, 36, e2311206 CrossRef PubMed.
  84. S. Wang, H. Chen, Y. Xu, G. Peng, H. Wang, Q. Li, X. Zhou, Z. Li, Q. Wang and Z. Jin, Organic Cation Modulation in Manganese Halides to Optimize Crystallization Process and X-Ray Response Toward Large-Area Scintillator Screen, Small, 2024, e2403234,  DOI:10.1002/smll.202403234.
  85. W. Zhang, P. Sui, W. Zheng, L. Li, S. Wang, P. Huang, W. Zhang, Q. Zhang, Y. Yu and X. Chen, Pseudo-2D Layered Organic-Inorganic Manganese Bromide with a Near-Unity Photoluminescence Quantum Yield for White Light-Emitting Diode and X-Ray Scintillator, Angew. Chem., Int. Ed., 2023, 62, e202309230 CrossRef CAS PubMed.
  86. S. B. Xiao, X. Zhang, X. Mao, H. J. Yang, Z. N. Chen and L. J. Xu, Ultrahigh X–Ray Imaging Spatial Resolution Enabled by an 0D Mn(II) Hybrid Scintillator, Adv. Funct. Mater., 2024, 34, 2404003 CrossRef CAS.
  87. J. Zhang, X. Wang, W. Q. Wang, X. Deng, C. Y. Yue, X. W. Lei and Z. Gong, Near-Unity Green Luminescent Hybrid Manganese Halides as X-ray Scintillators, Inorg. Chem., 2024, 63, 2647–2654 CrossRef CAS PubMed.
  88. W. Li, Y. Li, Y. Wang, Z. Zhou, C. Wang, Y. Sun, J. Sheng, J. Xiao, Q. Wang, S. Kurosawa, M. Buryi, D. John, K. Paurová, M. Nikl, X. OuYang and Y. Wu, Highly Efficient and Flexible Scintillation Screen Based on Organic Mn(II) Halide Hybrids toward Planar and Nonplanar X–Ray Imaging, Laser Photonics Rev., 2023, 18, 2300860 CrossRef.
  89. Z. Gong, J. Zhang, X. Deng, M. P. Ren, W. Q. Wang, Y. J. Wang, H. Cao, L. Wang, Y. C. He and X. W. Lei, Near–unity broadband emissive hybrid manganese bromides as highly–efficient radiation scintillators, Aggregate, 2024, 5, e574 CrossRef CAS.
  90. H. Wang, W. Li, K. Wang, H. Yin, X. Mao and R. Zhang, Turning self-trapped exciton emission via Sb3+ doping in zero-dimensional hybrid indium chloride for X-ray scintillation, J. Lumin., 2024, 269, 120507 CrossRef CAS.
  91. H. Huang, Y. Yang, S. Qiao, X. Wu, Z. Chen, Y. Chao, K. Yang, W. Guo, Z. Luo, X. Song, Q. Chen, C. Yang, Y. Yu and Z. Zou, Accommodative Organoammonium Cations in A–Sites of Sb–In Halide Perovskite Derivatives for Tailoring BroadBand Photoluminescence with X–Ray Scintillation and White–Light Emission, Adv. Funct. Mater., 2024, 34, 2309112 CrossRef CAS.
  92. M. Wang, Z. Zhang, J. Lyu, J. Qiu, C. Gu, H. Zhao, T. Wang, Y. Ren, S. W. Yang, G. Qin Xu and X. Liu, Overcoming Thermal Quenching in X-ray Scintillators through Multi-Excited State Switching, Angew. Chem., Int. Ed., 2024, 63, e202401949 CrossRef CAS PubMed.
  93. W. F. Wang, M. J. Xie, P. K. Wang, J. Lu, B. Y. Li, M. S. Wang, S. H. Wang, F. K. Zheng and G. C. Guo, Thermally Activated Delayed Fluorescence (TADF)-active Coinage-metal Sulfide Clusters for High-resolution X-ray Imaging, Angew. Chem., Int. Ed., 2024, 63, e202318026 CrossRef CAS PubMed.
  94. S. Yuan, G. Zhang, F. Chen, J. Chen, Y. Zhang, Y. Di, Y. Chen, Y. Zhu, M. Lin and H. Chen, Thermally Activated Delayed Fluorescent Ag(I) Complexes for Highly Efficient Scintillation and High–Resolution X–Ray Imaging, Adv. Funct. Mater., 2024, 34, 2400436 CrossRef CAS.
  95. J. O'Neill, J. Ghosh, S. Alghamdi, I. Braddock, C. Crean, R. Dorey, R. Mulholland, S. Richards, M. Wilson, H. Salway, M. Anaya, J. Reiss, D. Wolfe and P. Sellin, Hydrothermal and Mechanosynthesis of Mixed–Cation Double Perovskite Scintillators for Radiation Detection, Adv. Opt. Mater., 2024, 12, 2301335 CrossRef.
  96. C. Wang, Z.-G. Yan, Y. Wang, J. Zhu, C. Peng, Y. Qu, F. Yang, J. Xiao and X. Han, All-Inorganic Ruddlesden–Popper Perovskite Cs2CdCl4:Mn for Low-Dose and Flexible X-ray Imaging, ACS Mater. Lett., 2024, 6, 1429–1438 CrossRef CAS.
  97. Z. Wang, J. Chen, X. Xu, T. Bai, Q. Kong, H. Yin, Y. Yang, W. W. Yu, R. Zhang, X. Liu and K. Han, B(I)-Site Alkali Metal Engineering of Lead-Free Perovskite Nanocrystals for Efficient X-Ray Detection and Imaging, Adv. Opt. Mater., 2024, 12, 2302617 CrossRef CAS.
  98. H. Cui, W. Zhu, Y. Deng, T. Jiang, A. Yu, H. Chen, S. Liu and Q. Zhao, Lead-free organic–inorganic hybrid scintillators for X-ray detection, Aggregate, 2024, 5, e454 CrossRef CAS.
  99. L. Lian, M. Zheng, W. Zhang, L. Yin, X. Du, P. Zhang, X. Zhang, J. Gao, D. Zhang, L. Gao, G. Niu, H. Song, R. Chen, X. Lan, J. Tang and J. Zhang, Efficient and Reabsorption–Free Radioluminescence in Cs 3 Cu 2 I 5 Nanocrystals with Self–Trapped Excitons, Adv. Sci., 2020, 7, 2000195 CrossRef CAS PubMed.
  100. Q. Yao, J. Li, X. Li, Y. Ma, H. Song, Z. Li, Z. Wang and X. Tao, Achieving a Record Scintillation Performance by Micro-Doping a Heterovalent Magnetic Ion in Cs3Cu2I5 Single-Crystal, Adv. Mater., 2023, 35, 2304938 CrossRef CAS PubMed.
  101. G. Peng, F. Qiu, Z. Li, H. Wang, Q. Li and Z. Jin, Single–Source Evaporated High–Quality Single–Phase Cs3Cu2I5 Scintillator Films for High Performance X–Ray Imaging, Adv. Funct. Mater., 2024, 34, 2403052 CrossRef CAS.
  102. Z. Wang, L. Wang, J. Xie, Y. Yang, Y. Song, G. Xiao, Y. Fu, L. Zhang, Y. Fang, D. Yang and Q. Dong, HCOO(-) Doping-Induced Multiexciton Emissions in Cs(3)Cu(2)I(5) Crystals for Efficient X-Ray Scintillation, Small, 2024, 20, e2309922 CrossRef PubMed.
  103. I. K. Moon, S. Yoo, J. Choi, H. K. Kim and Y. Kang, Flexible Wood–Based X–Ray Scintillator Film Using Lead–Free Cs3Cu2I5 Perovskite Nanoparticles, Small Struct., 2024, 5, 2400043 CrossRef CAS.
  104. X. Hao, L. Nie, X. Zhu, G. Zeng, C. Liu, Z. Teng, H. Liu, Y. Yue, X. Yu and T. Wang, High-Resolution X-ray Image from Copper-Based Perovskite Hybrid Polymer, ACS Appl. Mater. Interfaces, 2024, 16, 29210–29216 CrossRef CAS.
  105. P. Ran, Q. Yao, J. Hui, Y. Su, L. Yang, C. Kuang, X. Liu and Y. Yang, Multi–Energy X–Ray Linear–Array Detector Enabled by the Side–Illuminated Metal Halide Scintillator, Laser Photonics Rev., 2024, 18, 2300587 CrossRef CAS.
  106. W. Xiang, D. Shen, X. Zhang, X. Li, Y. Liu and Y. Zhang, Transparent and Planar Cs(3)Cu(2)Cl(5) Crystals for Micrometer-Resolution X-ray Imaging Screen, ACS Appl. Mater. Interfaces, 2024, 16, 4918–4924 CrossRef CAS PubMed.
  107. P. Ran, X. Chen, Z. Chen, Y. Su, J. Hui, L. Yang, T. Liu, X. Tang, H. Zhu, X. J. She and Y. Yang, Metal Halide CsCu2I3 Flexible Scintillator with High Photodiode Spectral Compatibility for X–Ray Cone Beam Computed Tomography (CBCT) Imaging, Laser Photonics Rev., 2024, 18, 2300743 CrossRef CAS.
  108. H. Wu, Q. Wang, A. Zhang, G. Niu, M. Nikl, C. Ming, J. Zhu, Z. Zhou, Y.-Y. Sun, G. Nan, G. Ren, Y. Wu and J. Tang, One-dimensional scintillator film with benign grain boundaries for high-resolution and fast X-ray imaging, Sci. Adv., 2023, 9, eadh1789 CrossRef CAS PubMed.
  109. A. Zhang, J. Pang, H. Wu, Q. Tan, Z. Zheng, L. Xu, J. Tang and G. Niu, Event-based X-ray imager with ghosting-free scintillator film, Optica, 2024, 11, 606–611 CrossRef CAS.
  110. C. Shu, L. Lei, S. Tie, X. Lu, S. Dong, Z. Fan, N. Yang, R. Yuan, J. Zhu and X. Zheng, Strain-Released Cs5Cu3Cl6I2 Scintillators by Rb+ Doping for High-Resolution X-Ray Imaging, J. Phys. Chem. C, 2024, 128, 6106–6113 CrossRef CAS.
  111. Y. Hu, J. Jin, K. Han and Z. Xia, Unveiling The Role of Cu+ Doping in Rb2AgBr3 Scintillators toward Enhanced Photoluminescence Quantum Efficiency and Light Yield, Adv. Opt. Mater., 2024, 12, 2302063 CrossRef CAS.
  112. J. Yao, D. Huang, X. Hu, H. Cheng, D. Wang, X. Li, W. Yang and R. Xie, One-Dimensional Copper-Doped Rb(2)AgI(3) with Efficient Sky-Blue Emission as a High-Performance X-ray Scintillator, ACS Omega, 2024, 9, 28969–28977 CrossRef CAS PubMed.
  113. Q. Wu, X. Zhou, W. Gui, L. Yao, Y. Wang and C. L. Wang, Rb2AgBr3:Cu scintillators for high performance gamma and X-ray detection, Chem. Eng. J., 2025, 507, 160815 CrossRef CAS.
  114. Z. Wang, Y. Wei, C. Liu, Y. Liu and M. Hong, Mn2+-Activated Cs3Cu2I5 Nano-Scintillators for Ultra-High Resolution Flexible X-Ray Imaging, Laser Photonics Rev., 2023, 17, 2200851 CrossRef CAS.
  115. P. Ran, X. Chen, Z. Chen, Y. Su, J. Hui, L. Yang, T. Liu, X. Tang, H. Zhu, X.-J. She and Y. Yang, Metal Halide CsCu2I3 Flexible Scintillator with High Photodiode Spectral Compatibility for X-Ray Cone Beam Computed Tomography (CBCT) Imaging, Laser Photonics Rev., 2024, 18, 2300743 CrossRef CAS.
  116. P. Mao, Y. Tang, B. Wang, D. Fan and Y. Wang, Organic-Inorganic Hybrid Cuprous Halide Scintillators for Flexible X-ray Imaging, ACS Appl. Mater. Interfaces, 2022, 14, 22295–22301 CrossRef CAS PubMed.
  117. J. Ni, Q. Cao, K. Xiao, K. Gang, S. Liu, X. Liu and Q. Zhao, Colloidal Copper(I) Iodide Cluster-Based Scintillators for High-Resolution X-Ray Imaging, Laser Photonics Rev., 2025, 19, 2400963 CrossRef CAS.
  118. Y. Wang, T. Zhang, W. Zhao, W. Xu, Z. Wu, Y. D. Suh, Y. Zhang, X. Liu and W. Huang, Machine Learning-Guided Discovery of Copper(I)-Iodide Cluster Scintillators for Efficient X-ray Luminescence Imaging, Angew. Chem., 2025, 64, e202413672 CrossRef CAS PubMed.
  119. J. A. Lai, C. Li, Z. Wang, L. Guo, Y. Wang, K. An, S. Cao, D. Wu, Z. Liu, Z. Hu, Y. Leng, J. Du, P. He and X. Tang, Photoluminescence-Tunable organic phosphine cuprous halides clusters for X-ray scintillators and white light emitting diodes, Chem. Eng. J., 2024, 494, 153077 CrossRef CAS.
  120. S. Zhao, J. Zhao, S. M. H. Qaid, D. Liang, K. An, W. Cai, Q. Qian and Z. Zang, White emission metal halides for flexible and transparent X-ray scintillators, Appl. Phys. Rev., 2024, 11, 011408 CAS.
  121. N. Zhang, L. Qu, S. Dai, G. Xie, C. Han, J. Zhang, R. Huo, H. Hu, Q. Chen, W. Huang and H. Xu, Intramolecular charge transfer enables highly-efficient X-ray luminescence in cluster scintillators, Nat. Commun., 2023, 14, 2901 CrossRef CAS.
  122. Q.-H. Zou, W.-H. Yang, L.-K. Wu, L.-L. Jiang, S.-H. Wang, L. Liu, R.-F. Li, H.-Y. Ye and J.-R. Li, Ionothermal synthesis of a stable three-dimensional [Cu4I4] cluster scintillator with near-unity quantum efficiency and weak thermal quenching, Inorg. Chem. Front., 2025, 12, 692–700 RSC.
  123. Q. Zou, W. Yang, L. Wu, L. Jiang, S. Wang, L. Liu, R. Li, H. Ye and J. Li, Ionothermal synthesis of a hybrid cuprous(I) iodide scintillator with efficient cyan emission and high antiwater stability, Chem. Eng. J., 2025, 506, 159971 CrossRef CAS.
  124. H. Wang, J. X. Wang, X. Song, T. He, Y. Zhou, O. Shekhah, L. Gutierrez-Arzaluz, M. Bayindir, M. Eddaoudi, O. M. Bakr and O. F. Mohammed, Copper Organometallic Iodide Arrays for Efficient X-ray Imaging Scintillators, ACS Cent. Sci., 2023, 9, 668–674 CrossRef CAS PubMed.
  125. N. Lin, R. C. Wang, S. Y. Zhang, Z. H. Lin, X. Y. Chen, Z. N. Li, X. W. Lei, Y. Y. Wang and C. Y. Yue, 0D Hybrid Cuprous Halide as an Efficient Light Emitter and X–Ray Scintillator, Laser Photonics Rev., 2023, 17, 2300427 CrossRef CAS.
  126. H.-J. Yang, W. Xiang, X. Zhang, J.-Y. Wang, L.-J. Xu and Z.-N. Chen, High spatial resolution X-ray scintillators based on a 2D copper(I) iodide hybrid, J. Mater. Chem. C, 2024, 12, 438–442 RSC.
  127. Q. Kong, X. Jiang, Y. Sun, J. Zhu and X. Tao, Yellow-emissive organic copper(I) halide single crystals with [Cu4I4] cubane unit as efficient X-ray scintillators, Inorg. Chem. Front., 2024, 11, 3028–3035 RSC.
  128. H. Lv, W. Shao, H. Chen, G. Zhu, Y. Wang, Z. Zhang and H. Liang, Ultra-Broad Emission Copper Halide Scintillator-Based X-Ray Imager, Adv. Sci., 2025, 12, 2405995 CrossRef PubMed.
  129. T. He, Y. Zhou, P. Yuan, J. Yin, L. Gutiérrez-Arzaluz, S. Chen, J.-X. Wang, S. Thomas, H. N. Alshareef, O. M. Bakr and O. F. Mohammed, Copper Iodide Inks for High-Resolution X-ray Imaging Screens, ACS Energy Lett., 2023, 8, 1362–1370 CrossRef CAS.
  130. B. Li, J. Jin, X. Liu, M. Yin, X. Zhang, Z. Xia and Y. Xu, Multiphase Transformation in Hybrid Copper(I)-Based Halides Enable Improved X-ray Scintillation and Real-Time Imaging, ACS Mater. Lett., 2024, 6, 1542–1548 CrossRef CAS.
  131. R. Gu, K. Han, J. Jin, H. Zhang and Z. Xia, Surfactant-Assisted Synthesis of Hybrid Copper(I) Halide Nanocrystals for X-ray Scintillation Imaging, Chem. Mater., 2024, 36, 2963–2970 CrossRef CAS.
  132. Q. C. Peng, Y. B. Si, J. W. Yuan, Q. Yang, Z. Y. Gao, Y. Y. Liu, Z. Y. Wang, K. Li, S. Q. Zang and B. Zhong Tang, High Performance Dynamic X-ray Flexible Imaging Realized Using a Copper Iodide Cluster-Based MOF Microcrystal Scintillator, Angew. Chem., Int. Ed., 2023, 62, e202308194 CrossRef CAS PubMed.
  133. Y. V. Demyanov, Z. Ma, Z. Jia, M. I. Rakhmanova, G. M. Carignan, I. Y. Bagryanskaya, V. S. Sulyaeva, A. A. Globa, V. K. Brel, L. Meng, H. Meng, Q. Lin, J. Li and A. V. Artemev, Copper(I)–Arsine Clusters with a Near–Unity Phosphorescence Quantum Yield for X–Ray Scintillation and LED Applications, Adv. Opt. Mater., 2024, 12, 2302904 CrossRef CAS.
  134. P. Yuan, T. He, Y. Zhou, J. Yin, H. Zhang, Y. Zhang, X. Yuan, C. Dong, R. Huang, W. Shao, S. Chen, X. Song, R. Zhou, N. Zheng, M. Abulikemu, M. Eddaoudi, M. Bayindir, O. F. Mohammed and O. M. Bakr, Hybrid Thermally Activated Nanocluster Fluorophores for X-ray Scintillators, ACS Energy Lett., 2023, 8, 5088–5097 CrossRef CAS.
  135. Y. C. Chen, S. Q. Yuan, G. Z. Zhang, Y. M. Di, Q. W. Qiu, X. Yang, M. J. Lin, Y. N. Zhu and H. M. Chen, Mechanochemical Synthesis of Cuprous Complexes for X-ray Scintillation and Imaging, Inorg. Chem., 2024, 63, 3572–3577 CrossRef CAS PubMed.
  136. Y.-C. Chen, S.-Q. Yuan, G.-Z. Zhang, Y.-M. Di, Q.-W. Qiu, X. Yang, M.-J. Lin, Y.-N. Zhu and H.-M. Chen, Mechanochemical Synthesis of Cuprous Complexes for X-ray Scintillation and Imaging, Inorg. Chem., 2024, 63, 3572–3577 CrossRef CAS PubMed.
  137. X. Wu, Z. Guo, S. Zhu, B. Zhang, S. Guo, X. Dong, L. Mei, R. Liu, C. Su and Z. Gu, Ultrathin, Transparent, and High Density Perovskite Scintillator Film for High Resolution X–Ray Microscopic Imaging, Adv. Sci., 2022, 9, 2200831 CrossRef CAS PubMed.
  138. C. Wang, J. Xiao, S. Ma, Y. Li and Z. Yan, Efficient Near–Infrared Emitting Halide Scintillators with Mitigating External Light Crosstalk for Portable X–Ray Imaging, Adv. Funct. Mater., 2024, 34, 2401995 CrossRef CAS.
  139. T. Yang, P. Wu, R. Chen, J. Ni, C. Xu, X. Liu, S. Wang and S. Liu, Enhanced luminescence of Tb doping nanoscintillators based on NaLuF4 host matrix for X-ray imaging, J. Lumin., 2024, 269, 120498 CrossRef CAS.
  140. P. Qiao, J. Yang, H. Ma and L. Lei, Tuning scintillation property of CsLu2F7: Pr crystals for X-ray imaging, Opt. Mater., 2024, 148, 114812 CrossRef CAS.
  141. A. Zhang, S. Xu and L. Lei, Strongly reduced hysteresis scintillation luminescence in lanthanide doped fluoride nanoscintillators, J. Alloys Compd., 2023, 952, 169845 CrossRef CAS.
  142. H. Zou, M. Yi, S. Xu and L. Lei, A hollow NaBiF4:Tb nanoscintillator with ultra-weak afterglow for high-resolution X-ray imaging, Chem. Commun., 2023, 59, 11732–11735 RSC.
  143. M. Bungla, M. Tyagi, A. K. Ganguli and P. N. Prasad, A NaBiF4:Gd/Tb nanoscintillator for high-resolution X-ray imaging, J. Mater. Chem. C, 2024, 12, 9595–9605 RSC.
  144. K. Yang, Y. Yang, D. Sun, S. Li, X. Song and H. Yang, Designing highly UV-emitting lanthanide nanoscintillators for in vivo X-ray-activated tumor therapy, Sci. China Mater., 2023, 66, 4090–4099 CrossRef CAS.
  145. Z. Yang, P. Zhang, X. Chen, Z. Hong, J. Gong, X. Ou, Q. Wu, W. Li, X. Wang, L. Xie, Z. Zhang, Z. Yu, X. Qin, J. Tang, H. Zhang, Q. Chen, S. Han and H. Yang, High–Confidentiality X–Ray Imaging Encryption Using Prolonged Imperceptible Radioluminescence Memory Scintillators, Adv. Mater., 2023, 35, 2309413 CrossRef CAS PubMed.
  146. T. Yang, L. Guo, H. Wang, X. Xu, P. Wu, N. Zhang, X. Liu, S. Liu and Q. Zhao, Enhanced radioluminescence of NaLuF4:Eu3+ nanoscintillators by terbium sensitization for X-ray imaging, Inorg. Chem. Front., 2023, 10, 3974–3982 RSC.
  147. H. Zhang, B. Zhang, C. Cai, K. Zhang, Y. Wang, Y. Wang, Y. Yang, Y. Wu, X. Ba and R. Hoogenboom, Water-dispersible X-ray scintillators enabling coating and blending with polymer materials for multiple applications, Nat. Commun., 2024, 15, 2055 CrossRef CAS PubMed.
  148. H. Lu, X. Xu, Y. Li, G. Feng, J. Yang, X. Huang, Q. Luo, S. Wang and S. Wu, Heterovalent Co–Doped LiLuF4 Composite Flexible Films as Persistent Luminescence Scintillator for X–Ray High–Resolution Extension Imaging, Adv. Opt. Mater., 2024, 12, 2303134 CrossRef CAS.
  149. X. Liu, R. Li, X. Xu, Y. Jiang, W. Zhu, Y. Yao, F. Li, X. Tao, S. Liu, W. Huang and Q. Zhao, Lanthanide(III)–Cu4I4 Organic Framework Scintillators Sensitized by Cluster–Based Antenna for High–Resolution X–ray Imaging, Adv. Mater., 2023, 35, 2206741 CrossRef CAS PubMed.
  150. H. Li, Y. Li, L. Zhang, E. Hu, D. Zhao, H. Guo and G. Qian, A Thermo–Responsive MOFs for X–Ray Scintillator, Adv. Mater., 2024, 36, 2405535 CrossRef CAS PubMed.
  151. J. Jin, K. Han, Y. Wang and Z. Xia, Bandgap Narrowing in Europium(II)-Based Bromide Hybrids toward Improved X-ray Scintillation and Imaging, Chem. Mater., 2024, 36, 4813–4820 CrossRef CAS.
  152. J. Yang, H. Lu, L. Yang, Y. Yao, Z. Wei, M. Chen, H. Qi, Y. Ren, Y. Wang, J. Qiu and J. Lin, Lanthanide Organic–Inorganic Hybrids for X-ray Scintillation and Imaging, Inorg. Chem., 2024, 63, 3642–3647 CrossRef CAS PubMed.
  153. J. Xu, R. Luo, Z. Luo, J. Xu, Z. Mu, H. Bian, S. Y. Chan, B. Y. H. Tan, D. Chi, Z. An, G. Xing, X. Qin, C. Gong, Y. Wu and X. Liu, Ultrabright molecular scintillators enabled by lanthanide-assisted near-unity triplet exciton recycling, Nat. Photonics, 2025, 19, 71–78 CrossRef CAS.
  154. P.-K. Wang, W.-F. Wang, B.-Y. Li, R.-X. Qian, Z.-X. Gao, S.-H. Wang, F.-K. Zheng and G.-C. Guo, Highly sensitive terbium-based metal–organic framework scintillators applied in flexible X-ray imaging, Inorg. Chem. Front., 2025, 12, 1040–1048 RSC.
  155. L. Li, J. Chen, X. Peng, T. Jiang, L. Lei and H. Guo, High-resolution Tb3+-doped Gd-based oxyfluoride glass scintillators for X-ray imaging, J. Mater. Chem. C, 2023, 11, 11664–11670 RSC.
  156. J. Guo, L. Li, J. Chen, H. Li and H. Guo, Ce3+-doped oxyfluoride glass scintillator: optimized radioluminescence and application in X-ray imaging, J. Alloys Compd., 2024, 980, 173670 CrossRef CAS.
  157. Z. Sun, X. Huang, J. Yang, S. Wang and S. Wu, Preparation and scintillation properties of Ce3+/Tb3+-co-doped oxyfluoride glass for high-resolution X-ray imaging, Ceram. Int., 2023, 49, 15500–15506 CrossRef CAS.
  158. L. Li, J. Chen, Z. Wen, J. Guo, Q. Wang and H. Guo, X-ray imaging scintillator: Tb3+-doped oxyfluoride aluminosilicate glass, Ceram. Int., 2024, 50, 757–763 CrossRef CAS.
  159. W. Dai, Q. Zhang, G. A. Ashraf, H. Gong, R. Wei, H. Guo and F. Hu, Novel transparent NaLu2F7:Tb3+ glass-ceramics scintillator for highly resolved X-ray imaging, Ceram. Int., 2024, 50, 21878–21888 CrossRef CAS.
  160. Y. Liu, K. Han, S. Zhao, X. Zhou, S. Gao, Y. Wang, Z. Xia and S. Jiang, Ce3+-doped alkaline-earth sulfide nanocrystals for X-ray scintillation imaging screens, J. Mater. Chem. C, 2024, 12, 3221–3227 RSC.
  161. D. Wang, H. Li, J. Chen, D. Hou, X. Liu, Z. Yu, S. Lv, J. Qiu and S. Zhou, Congruent Glass Composite Scintillator for Efficient High-Energy Ray Detection, Adv. Mater., 2025, 37, 2412661 CrossRef CAS.
  162. Y. Ling, X. Zhao, P. Hao, Y. Song, J. Liu, L. Zhao, Y. Qian and C. Guo, Gd3+-sensitized rare earth fluoride scintillators for High-resolution flexible X-ray imaging, Chem. Eng. J., 2023, 476, 146790 CrossRef CAS.
  163. T. Wang, S. Hu, T. Ji, X. Zhu, G. Zeng, L. Huang, A. N. Yakovlev, J. Qiu, X. Xu and X. Yu, High–Temperature X–Ray Imaging with Transparent Ceramics Scintillators, Laser Photonics Rev., 2024, 18, 2300892 CrossRef CAS.
  164. J. Yang, X. Huang, X. Xu, H. Lu, S. Wang and S. Wu, Layered Chalcogenide Scintillators Enabled by Reversible Hydrous-Induced Phase Transformation for High-Resolution X-ray Imaging, ACS Appl. Mater. Interfaces, 2024, 16, 15050–15058 CrossRef CAS PubMed.
  165. S. Wang, J. Yang, C. Liu, W. Li, X. Zheng, X. Qiao, X. Sun, S. Qian, J. Han, J. Wu, X. Xu, J. Ren and J. Zhang, KTb3−GdF10 Nano-Glass Composite Scintillator with Excellent Thermal Stability and Record X-Ray Imaging Resolution, Laser Photonics Rev., 2025, 19, 2401611 CrossRef CAS.
  166. H. Li, P. Wang, G. Lin and J. Huang, The role of rare earth elements in biodegradable metals: A review, Acta Biomater., 2021, 129, 33–42 CrossRef CAS PubMed.
  167. D. A. Gkika, M. Chalaris and G. Z. Kyzas, Review of Methods for Obtaining Rare Earth Elements from Recycling and Their Impact on the Environment and Human Health, Processes, 2024, 12, 1235 CrossRef CAS.
  168. Y.-H. Chen, G.-Z. Zhang, F.-H. Chen, S.-Q. Zhang, X. Fang, H.-M. Chen and M.-J. Lin, Halogen-bonded charge-transfer co-crystal scintillators for high-resolution X-ray imaging, Chem. Sci., 2024, 15, 7659–7666 RSC.
  169. T. F. Manny, T. B. Shonde, H. Liu, M. S. Islam, O. J. Olasupo, J. Viera, S. Moslemi, M. Khizr, C. Chung, J. S. R. V. Winfred, L. M. Stand, E. F. Hilinski, J. Schlenoff and B. Ma, Efficient X-Ray Scintillators Based on Facile Solution Processed 0D Organic Manganese Bromide Hybrid Films, Adv. Funct. Mater., 2025, 35, 2413755 CrossRef CAS.
  170. S. He, P. Wan, H. Li, Z. Bao, X. Sui, G. Zheng, H. Yin, J. Pang, T. Jin, S. Yuan, L. Xu, X. Ouyang, Z. Zheng, L. Liu, G. Niu and J. Tang, Hot Exciton-Based Plastic Scintillator Engineered for Efficient Fast Neutron Detection and Imaging, Adv. Funct. Mater., 2025, 2503688 CrossRef.
  171. W. Zhao, Z. He and B. Z. Tang, Room-temperature phosphorescence from organic aggregates, Nat. Rev. Mater., 2020, 5, 869–885 CrossRef CAS.
  172. W. Ma, Y. Su, Q. Zhang, C. Deng, L. Pasquali, W. Zhu, Y. Tian, P. Ran, Z. Chen, G. Yang, G. Liang, T. Liu, H. Zhu, P. Huang, H. Zhong, K. Wang, S. Peng, J. Xia, H. Liu, X. Liu and Y. M. Yang, Thermally activated delayed fluorescence (TADF) organic molecules for efficient X-ray scintillation and imaging, Nat. Mater., 2021, 21, 210–216 CrossRef.
  173. X. Wang, H. Shi, H. Ma, W. Ye, L. Song, J. Zan, X. Yao, X. Ou, G. Yang, Z. Zhao, M. Singh, C. Lin, H. Wang, W. Jia, Q. Wang, J. Zhi, C. Dong, X. Jiang, Y. Tang, X. Xie, Y. Yang, J. Wang, Q. Chen, Y. Wang, H. Yang, G. Zhang, Z. An, X. Liu and W. Huang, Organic phosphors with bright triplet excitons for efficient X-ray-excited luminescence, Nat. Photonics, 2021, 15, 187–192 CrossRef CAS.
  174. X. Chen, D. Ma, T. Liu, Z. Chen, Z. Yang, J. Zhao, Z. Yang, Y. Zhang and Z. Chi, Hybridized Local and Charge-Transfer Excited-State Fluorophores through the Regulation of the Donor–Acceptor Torsional Angle for Highly Efficient Organic Light-Emitting Diodes, CCS Chem., 2022, 4, 1284–1294 CrossRef CAS.
  175. Y. X. Hu, J. Miao, T. Hua, Z. Huang, Y. Qi, Y. Zou, Y. Qiu, H. Xia, H. Liu, X. Cao and C. Yang, Efficient selenium-integrated TADF OLEDs with reduced roll-off, Nat. Photonics, 2022, 16, 803–810 CrossRef CAS.
  176. C. Dong, X. Wang, W. Gong, W. Ma, M. Zhang, J. Li, Y. Zhang, Z. Zhou, Z. Yang, S. Qu, Q. Wang, Z. Zhao, G. Yang, A. Lv, H. Ma, Q. Chen, H. Shi, Y. Yang and Z. An, Influence of Isomerism on Radioluminescence of Purely Organic Phosphorescence Scintillators, Angew. Chem., Int. Ed., 2021, 60, 27195–27200 CrossRef CAS.
  177. H. Chen, M. Lin, C. Zhao, D. Zhang, Y. Zhang, F. Chen, Y. Chen, X. Fang, Q. Liao, H. Meng and M. Lin, Highly Efficient, Low–Dose, and Ultrafast Carbazole X–Ray Scintillators, Adv. Opt. Mater., 2023, 11, 2300365 CrossRef CAS.
  178. H. Chen, M. Lin, Y. Zhu, D. Zhang, J. Chen, Q. Wei, S. Yuan, Y. Liao, F. Chen, Y. Chen, M. Lin and X. Fang, Halogen–bonding boosting the high performance X–ray imaging of organic scintillators, Small, 2024, 20, 2307277 CrossRef CAS PubMed.
  179. T. Chen, Y. Xu, A. Ying, C. Yang, Q. Lin and S. Gong, Through–Space Charge–Transfer Organogold(III) Complexes Enable High–Performance X–ray Scintillation and Imaging, Angew. Chem., Int. Ed., 2024, 63, e202401833 CrossRef CAS PubMed.
  180. V. W.-W. Yam, S. W.-K. Choi, T.-F. Lai and W.-K. Lee, Syntheses, crystal structures and photophysics of organogold(III) diimine complexes, J. Chem. Soc., Dalton Trans., 1993, 1001–1002 RSC.
  181. M. Dong, A. Lv, X. Zou, N. Gan, C. Peng, M. Ding, X. Wang, Z. Zhou, H. Chen, H. Ma, L. Gu, Z. An and W. Huang, Polymorphism–Dependent Organic Room Temperature Phosphorescent Scintillation for X–Ray Imaging, Adv. Mater., 2024, 36, 2310663 CrossRef CAS.
  182. L. Zhan, Y. Xu, T. Chen, Y. Tang, C. Zhong, Q. Lin, C. Yang and S. Gong, Organic molecules with dual triplet–harvesting channels enable efficient X–ray scintillation and imaging, Aggregate, 2024, 5, e485 CrossRef CAS.
  183. G. Zhang, F. Chen, Y. Di, S. Yuan, Y. Zhang, X. Quan, Y. Chen, H. Chen and M. Lin, Guest–Induced Thermally Activated Delayed Fluorescence Organic Supramolcular Macrocycle Scintillators for High–Resolution X–Ray Imaging, Adv. Funct. Mater., 2024, 34, 2404123 CrossRef CAS.
  184. H. Wang, C. Peng, M. Chen, Y. Xiao, T. Zhang, X. Liu, Q. Chen, T. Yu and W. Huang, Wide–Range Color–Tunable Organic Scintillators for X–Ray Imaging Through Host–Guest Doping, Angew. Chem., Int. Ed., 2024, 63, e202316190 CrossRef CAS.
  185. W. Yang, P. Han, S. Zhu, Z. Cui, Z. Li, S. Wu, W. Xu, Z. Gao, T. Ba, Y. Liang, H. Jiang and W. Hu, Bulk Organic Doped p-Terphenyl Crystals for X-ray Scintillators with Low Detection Limit, Nanoseconds Decay Time, and Tunable Fluorescence, ACS Appl. Electron. Mater., 2024, 6, 4223–4231 CrossRef CAS.
  186. J. Chen, G. Liu, F. Chen, Y. Chen, X. Fang, H. Chen and M.-J. Lin, Binaphthol diimide scintillators for X-ray imaging, Sci. China Mater., 2024, 67, 2583–2589 CrossRef CAS.
  187. Y. Zhang, M. Chen, X. Wang, M. Lin, H. Wang, W. Li, F. Chen, Q. Liao, H. Chen, Q. Chen, M. Lin and H. Yang, Efficient and Fast X-Ray Luminescence in Organic Phosphors Through High-Level Triplet-Singlet Reverse Intersystem Crossing, CCS Chem., 2024, 6, 334–341 CrossRef CAS.
  188. X. Du, S. Zhao, L. Wang, H. Wu, F. Ye, K.-H. Xue, S. Peng, J. Xia, Z. Sang, D. Zhang, Z. Xiong, Z. Zheng, L. Xu, G. Niu and J. Tang, Efficient and ultrafast organic scintillators by hot exciton manipulation, Nat. Photonics, 2024, 18, 162–169 CrossRef.
  189. T. J. Hajagos, C. Liu, N. J. Cherepy and Q. Pei, High-Z Sensitized Plastic Scintillators: A Review, Adv. Mater., 2018, 30, 1706956 CrossRef PubMed.
  190. J. Hui, P. Ran, Y. Su, L. Yang, X. Xu, T. Liu and Y. M. Yang, High-Resolution X-ray Imaging Based on Solution-Phase-Synthesized Cs3Cu2I5 Scintillator Nanowire Array, J. Phys. Chem. C, 2022, 126, 12882–12888 CrossRef CAS.
  191. Z. Zhang, H. Dierks, N. Lamers, C. Sun, K. Nováková, C. Hetherington, I. G. Scheblykin and J. Wallentin, Single-Crystalline Perovskite Nanowire Arrays for Stable X-ray Scintillators with Micrometer Spatial Resolution, ACS Appl. Nano Mater., 2021, 5, 881–889 CrossRef.
  192. J. Hui, P. Ran, Y. Su, L. Yang, X. Xu, T. Liu, Y. Gu, X. She and Y. Yang, Stacked Scintillators Based Multispectral X-Ray Imaging Featuring Quantum-Cutting Perovskite Scintillators With 570 nm Absorption-Emission Shift, Adv. Mater., 2025, 37, 2416360 CrossRef CAS PubMed.
  193. Z. Lin, S. Lv, Z. Yang, J. Qiu and S. Zhou, Structured Scintillators for Efficient Radiation Detection, Adv. Sci., 2021, 9, 2102439 CrossRef PubMed.
  194. C. Roques-Carmes, N. Rivera, A. Ghorashi, S. E. Kooi, Y. Yang, Z. Lin, J. Beroz, A. Massuda, J. Sloan, N. Romeo, Y. Yu, J. D. Joannopoulos, I. Kaminer, S. G. Johnson and M. Soljačić, A framework for scintillation in nanophotonics, Science, 2022, 375, eabm9293 CrossRef CAS PubMed.
  195. X. Xu, Y. M. Xie, H. Shi, Y. Wang, X. Zhu, B. X. Li, S. Liu, B. Chen and Q. Zhao, Light Management of Metal Halide Scintillators for High–Resolution X–Ray Imaging, Adv. Mater., 2024, 36, 2303738 CrossRef CAS.
  196. H. Dierks, Z. Zhang, N. Lamers and J. Wallentin, 3D X-ray microscopy with a CsPbBr3 nanowire scintillator, Nano Res., 2022, 16, 1084–1089 CrossRef.
  197. L. Yi, B. Hou, H. Zhao, H. Q. Tan and X. Liu, A double-tapered fibre array for pixel-dense gamma-ray imaging, Nat. Photonics, 2023, 17, 494–500 CrossRef CAS.
  198. W. Shao, T. He, L. Wang, J. X. Wang, Y. Zhou, B. Shao, E. Ugur, W. Wu, Z. Zhang, H. Liang, S. De Wolf, O. M. Bakr and O. F. Mohammed, Capillary Manganese Halide Needle–Like Array Scintillator with Isolated Light Crosstalk for Micro–X–Ray Imaging, Adv. Mater., 2024, 36, 2312053 CrossRef CAS PubMed.
  199. J. Song, Y. He, H. Hu, M. Li, C. Li, B. Yang and H. Wei, Anti-Scattering Perovskite Scintillator Arrays for High-Resolution Computed Tomography Imaging, Adv. Mater., 2025, 37, 2417248 CrossRef CAS PubMed.
  200. W. Shao, T. He, J.-X. Wang, Y. Zhou, P. Yuan, W. Wu, Z. Zhang, O. M. Bakr, H. Liang and O. F. Mohammed, Transparent Organic and Metal Halide Tandem Scintillators for High-Resolution Dual-Energy X-ray Imaging, ACS Energy Lett., 2023, 8, 2505–2512 CrossRef CAS.
  201. X. Hu, S. Luo, J. Leng, C. Wang, Y. Chen, J. Chen, X. Li and H. Zeng, Density-discriminating chromatic X-ray imaging based on metal halide nanocrystal scintillators, Sci. Adv., 2023, 9, eadh5081 CrossRef CAS PubMed.

Footnote

These authors contributed equally to this work.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.