Next Article in Journal
Monodentate Ligands in X-Cu(I)-Y Complexes—Structural Aspects
Previous Article in Journal
Late Transition Metal Olefin Polymerization Catalysts Derived from 8-Arylnaphthylamines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging New-Generation Semiconductor Single Crystals of Metal Halide Perovskites for Radiation Detection

College of Engineering and Applied Sciences, State Key Laboratory of Analytical Chemistry for Life Science, National Laboratory of Microstructures, Nanjing University, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Inorganics 2024, 12(11), 278; https://doi.org/10.3390/inorganics12110278
Submission received: 9 October 2024 / Revised: 27 October 2024 / Accepted: 28 October 2024 / Published: 30 October 2024
(This article belongs to the Section Inorganic Solid-State Chemistry)

Abstract

:
Radiation detection uses semiconductor materials to convert high-energy photons into charge (direct detection) or low-energy photons (indirect detection), and it has a wide range of applications in nuclear physics, medical imaging, astronomical detection, homeland security, and other fields. Metal halide perovskites have the advantages of high frequency number, high carrier mobility, high defect tolerance, low defect density, adjustable band gap, and fast light response, and they have wide application prospects in the field of radiation detection. However, the research is still in its infancy stage, and it is far from meeting the requirements of industrial application. This paper focuses on the advantages of metal halide perovskite single-crystal materials in both semiconductors-based direct conversion detection and scintillator-based indirect detection as well as the latest progress in this promising field. This paper not only introduces the latest application of lead halide perovskite monocrystalline materials in high-energy electromagnetic radiation detection (X-ray and γ-rays), but it also introduces the latest development of α-particle/β-particle/neutron detection. Finally, this paper points out the challenges and future prospects of metal halide perovskite single-crystal materials in radiation detection.

1. Introduction

Ionizing radiation, encompassing high-energy electromagnetic radiation (X-rays and γ-rays) and particle radiation (alpha particles, beta particles, and neutrons), plays a pivotal role in numerous research fields: physics [1,2], materials [3,4], chemistry [4,5], crystallography [6], molecular biology [7], and astronomy [8]. X-rays, γ-rays, and neutrons serve as powerful tools for medical imaging [9], industrial inspection [10], and homeland security [11]. As such, high-performance radiation detectors are vital for extracting the valuable information carried by high-energy photons and particles.
Radiation detectors can be classified by their working mechanisms into direct conversion detectors (solid-state detectors) and indirect detectors (scintillators) [12,13]. Direct conversion detectors function by applying a bias voltage to gather electron–hole pairs generated by interactions with high-energy photons or particles. To improve charge transport and ensure efficient collection, the detector materials must have a high mobility-lifetime product (μτ product), which allows electrons or holes to travel through the detector and reach the electrodes. Additionally, to reduce leakage current under high electric fields, the material should have a wide bandgap and a low defect density. Consequently, direct detectors typically employ large-sized semiconductors with low defect density. Scintillators, on the other hand, function by transforming radiation into low-energy visible light through the radiative recombination of electron–hole pairs. The light emitted is then detected by devices such as charge-coupled devices (CCDs), photomultiplier tubes (PMTs), or silicon photomultipliers (SiPMs) [14,15]. To be effective, scintillators must fulfill several criteria: They should have a high light yield, indicating the number of photons produced per unit of absorbed radiation energy; a fast decay time, significant for applications requiring rapid detection; and an emission wavelength that aligns well with the sensitivity of optical detectors to ensure efficient coupling.
X-rays and γ-rays interact with matter through three primary mechanisms: photoelectric absorption, Compton scattering, and pair production. Photoelectric absorption refers to the complete absorption of high-energy photons by inner shell electrons, resulting in the production of kinetic energy. Compton scattering involves the scattering of incident photons by outer shell electrons, which gain kinetic energy in the process. Pair production occurs only when the photon energy exceeds 1.022 MeV, leading to the generation of electron–positron pairs. For charged particles, interactions primarily involve Coulomb forces and electron scattering. The energy of the scattered electrons is ultimately converted into electron–hole pairs. While α particles transfer less energy per collision, they can interact with multiple electrons. In contrast, β particles exhibit greater penetration depth and may also produce bremsstrahlung radiation. Neutron detection is more complex, relying primarily on elastic scattering. Thermal neutrons can be effectively captured by specific nuclei, while fast neutrons interact with hydrogen atoms, transferring energy to electrons through the recoil nucleus.
Metal halide perovskites have gained extensive interest in the fields of materials [12], physics [16], and optoelectronic research [17] due to their cost-effective solution processability, tunable bandgap, structural flexibility, and excellent optoelectronic properties [18,19]. The general chemical formula of halide perovskites is ABX3, where A represents a small organic cation or an inorganic cation (e.g., MA+, FA+, Cs+, or Rb+), B is a metal cation (e.g., Pb2+, Cu+, Sn2+, or Bi3+), and X is a halide anion (Cl, Br, or I) [20]. Metal halide perovskites typically exhibit low trap densities, high carrier mobility, and high photoluminescence quantum yields (PLQYs). These exceptional optical and electronic properties make them promising candidates for applications in light-emitting diodes (LEDs) [21], photodetectors [22], radiation detectors [23,24], and lasers [25]. Moreover, metal halide perovskites are developing into promising semiconductors for high-energy radiation detection, with potential applications for detecting X-rays [26], γ-rays [27], alpha particles [28], beta particles [29], and neutrons [30]. Currently, traditional semiconductors such as Si, α-Se, and CdZnTe are widely used in direct detectors. However, these materials have inherent challenges. For example, Si with low X-ray stopping power is limited in the potential applications above 50 keV [31]. Additionally, complex fabrication processes, high operating voltages, and high manufacturing costs create significant obstacles to their practical use. Traditional scintillators such as CsI, Bi4Ge3O12 (BGO), and (Lu,Y)2SiO5 (LYSO) have also been successfully commercialized due to their energy resolution and high light yield. However, these materials also face a number of challenges, such as non-tunable scintillation resulting from fixed transition energies, radiation-induced afterglow, and the complexity of their fabrication processes [32].
Metal halide perovskites offer advantages for radiation detection owing to their high-atomic-number ions, which provide strong radiation absorption, and their high mobility-lifetime (μτ) product, enabling sensitive detection. Various forms, including single crystals and nanocrystals, have been applied in direct radiation detectors [33]. Additionally, their high PLQYs, tunable optical bandgaps, and fast photoluminescence lifetimes make them promising materials for scintillation [34]. Recent years have produced significant advancements in developing efficient halide perovskite scintillators.
In this review, we begin by summarizing recent advancements in various perovskite single crystals for X-ray detectors, energy-resolved γ-ray detectors, and particle radiation detectors, encompassing organic–inorganic lead halide perovskites, all-inorganic lead halide perovskites, and all-inorganic lead-free perovskites. We then focus on the performance of direct detectors based on perovskite single crystals, discussing key parameters such as the mobility-lifetime (μτ) product, sensitivity, and detection limit (Table 1). Additionally, we provide an overview of the scintillation properties of perovskite single crystals, including light yield, emission wavelength, and decay time (Table 2), as well as their applications in alpha particle, beta particle, and neutron detection (Table 3). Finally, we address the challenges faced in this field and offer insightful perspectives on future research directions.

2. Perovskite Single Crystals for Direct Detection of X-Ray and γ-Ray

Semiconductor materials used in direct detectors must incorporate heavy elements for effective interaction with X-rays and γ-rays while also fulfilling several essential criteria. Firstly, they must have a large μτ product to ensure efficient charge collection. Secondly, minimizing the trap density is crucial to reduce dark current, and thirdly, high resistivity is essential to ensure stability during high electric field operation. Perovskite single crystals generally fulfill these requirements, including organic–inorganic perovskite SCs, all-inorganic lead-based perovskite SCs, and all-inorganic lead-free perovskite SCs.

2.1. Organic–Inorganic Lead Halide Perovskite Single Crystals for Direct Detection of X-Ray and γ-Ray

Organic–inorganic perovskites can be fabricated using low-temperature solution methods for radiation detection. Yakunin et al. first used perovskite polycrystalline films to create direct X-ray detectors in photodiode and photoconductive structures. Using spin-coated MAPbI3 polycrystalline films (<1 μm) in a photodiode structure, they achieved a X-ray sensitivity of specifically 25 μC mGy1 cm3 [26]. In another study, they developed gamma-ray detectors using MAPbI3, FAPbI3, and iodine-treated MAPbBr3, which detected gamma rays from 137Cs and 241Am sources [27]. Among these, FAPbI3 could resolve the 241Am spectrum at 59.6 keV with a 35% energy resolution. Despite its good performance, FAPbI3 crystals experienced a phase transition from the α-phase to the wide bandgap δ-phase within few days, limiting their long-term stability and practical use.
In contrast, MA-based organic–inorganic perovskites exhibit greater structural stability [61]. Recently, Liu et al. recently used a vacuum evaporation crystallization (VEC) method to grow MAPbBr3 single crystals (SCs), which demonstrated a narrower rocking curve of 0.00922°, a longer photoluminescence lifetime of 1150 ns, a lower trap density of 2.28 × 109 cm3, and a higher carrier mobility of 130 cm2 V1 s1. These MAPbBr3 SCs achieved a low detection limit of 54 nGy s1 and a high sensitivity of 24,552 μC Gy1 cm2 in X-ray detectors, making them the most sensitive MAPbBr3 X-ray detectors reported to date [36]. Moreover, even after 200 days without encapsulation, the detector maintained 97.1% of its initial photocurrent response at a dose rate of 21.09 mGy s1 [36].
Organic–inorganic perovskites also face a significant challenge: ion migration. Ion migration in perovskites often results in detector instability, dark current drift, and hysteresis, which limits their application in mature technologies. Liu et al. reported a series of MAPbI3 SC X-ray detectors with different deuteration levels (DxMAPbI3, where x = 0, 0.15, 0.75, and 0.99) [35] (Figure 1a,d). By controlling the deuterium (D) content in the organic cation, ion migration in MAPbI3 could be effectively suppressed, allowing for control over the detector’s sensitivity, detection limit, ion migration, and resistivity, thus enhancing its performance. Deuteration of MAPbI3 crystals affects exciton behavior, hydrogen bonding, and crystal stability, potentially enhancing photovoltaic efficiency. The lower zero-point energy and shorter bond length of C–D compared to C–H reduce vibration amplitude, possibly aiding exciton separation and migration. The shorter N–D⋯I hydrogen bonds strengthen the crystal lattice, increasing structural stability and modifying interlayer forces. These effects may raise activation barriers in processes like charge transfer and defect migration, thereby enhancing thermal and dynamic stability. As a result, the D0.99MAPbI3 SC detector exhibited over fivefold enhancement, achieving a record-high μτ product of 5.39 × 102 cm2 V1, ultra-high sensitivity of 2.18 × 106 μC Gy1 cm2 under 120 keV X-rays, a low detection limit of 4.8 nGyair s−1, and long-term stability.
MA-based perovskites have also been studied for γ-ray detection. In the following study, Wei et al. showed that doping bromide crystals with chloride (MAPbBr2.94Cl0.06) improved the typical resolution of γ-ray detectors at 662 keV to 12%, with the best-performing device achieving a resolution of 6.5% [43]. The incorporation of Cl ions significantly lowered extrinsic conductivity, resulting in a high resistivity of 3.6 × 109 Ω·cm, approaching that of intrinsic semiconductors. Additionally, He et al. enhanced a detector using MAPbI3, achieving an energy resolution of 6.8% (57Co 122 keV) [44] (Figure 1e). To address the low resistivity, they employed asymmetric Schottky-type contacts to minimize dark current, which is essential for better energy resolution. However, MAPbI3 crystals continue to face stability issues due to voltage-induced ion migration. After 1 h of continuous exposure to gamma ray (57Co 122 keV), the number of channels decreased by 1.7%, the counting rate dropped by 18%, and the energy resolution worsened from 6.8% to 8.7% showed in Figure 1f.

2.2. All-Inorganic Lead Halide Perovskite Single Crystals for Direct Detection of X-Ray and γ-Ray

Despite the rapid development of organic–inorganic perovskite-based radiation detectors, they face challenges with long-term stability, necessitating further research to address this issue. In contrast, all-inorganic metal halide perovskites, such as CsPbBr3 and CsPbCl3, show superior thermal stability owing to the lack of organic molecules. At the initial stage, Stoumpos et al. proposed using CsPbBr3 single crystals for radiation detection in 2013 [3]. These CsPbBr3 crystals, prepared using the vertical Bridgman method, demonstrated a resistivity from 109 to 1011 Ω·cm, which exceeds that of organic–inorganic perovskites (108 to 109 Ω·cm). The detector successfully resolved the Kα and Kβ peaks from a silver X-ray source, although no γ-ray spectrum was obtained.
Subsequently, He et al. employed an optimized Bridgman growth method to produce large, crack-free single-crystal ingots with very high purity in which the total impurities were lower than 10 ppm, with low defect density. By utilizing a specialized device design with asymmetric electrode materials, the dark leakage was effectively suppressed, enabling CsPbBr3 detectors to achieve excellent energy resolution. Specifically, CsPbBr3 detectors successfully resolved spectra for 662 keV (137Cs γ-rays), with the best spectral resolution reaching 3.8%. (Figure 2a,b) [24] This exceptional performance was attributed to an unprecedentedly slow hole lifetime (exceeding 25 μs) and an excellent hole μτ product (1.34 × 103 cm2 V1). In addition to excellent energy resolution, CsPbBr3 γ-ray detectors exhibited good spectral linearity across a broad energy range of 32.3 to 662 keV (Figure 2c) and maintained continuous operation for over 120 h, significantly outperforming their organic–inorganic perovskite counterparts.
The same group reported significant progress in developing next-generation CsPbBr3-based γ-ray detectors, achieving improved performance that could potentially surpass CZT detectors. They fabricated CsPbBr3 ingots with a 1.5-inch diameter [45], where the increased diameter led to an enhanced hole μτ product (8 × 103 cm2 V1) and an extended lifetime of 296 μs. Enhanced crystal growth conditions led to an energy resolution of 1.4% for 662 keV γ-rays, rivaling that of cutting-edge CZT detectors. In addition, CsPbBr3 detectors demonstrated outstanding thermal and long-term stability (18 months; 2 °C to 70 °C).
In addition to the vertical Bridgman method for growing CsPbBr3 single crystals, Hua et al. developed a new atmosphere-controlled Edge-defined Film-fed Growth (EFG) technique to address the issues of low resistivity and significant ion migration in CsPbBr3 SCs [37]. This method offers several advantages, including rapid growth, low production cost, high crystal quality, and low trap density. Compared to CsPbBr3 SCs grown by other methods, EFG-CsPbBr3 exhibits higher resistivity (1.61 × 1010 Ω·cm), an energy barrier of 0.378 eV, and a large μτ product (8.11 × 104 cm2 V1). These properties contribute to an impressively low detection limit of 10.81 nGyair s1. Under an applied electric field of 5000 V cm1, the EFG-CsPbBr3 detector achieved a high sensitivity of 46,180 μC Gyair1 cm2 for 120 keV X-rays while maintaining stable dark current at high voltages. Furthermore, the device demonstrated long-term detection capabilities, operating for over 30 days without encapsulation.
Ion doping has also been instrumental in improving the low resistivity and mitigating severe ion migration in CsPbBr3, both of which negatively impact X-ray detector performance by increasing detection limits and causing current drift. Zhang et al. investigated the electrical properties and performance of X-ray detection of CsPbBr3−nIn SCs by doping iodine atoms into melt-grown CsPbBr3 [38]. As a result, the resistivity of CsPbBr3−nIn increased from 3.6 × 109 Ω·cm to 2.2 × 1011 Ω·cm (CsPbBr3 to CsPbBr2I), effectively suppressing leakage current and lowering the detection limit. Furthermore, CsPbBr3−nIn single crystals exhibited stable dark current due to their high ion migration activation energy. Under an electric field of 5000 V cm1, CsPbBr3−nIn SCs achieved a high sensitivity of 6.3 × 104 μC Gy1 cm2 (CsPbBr2.9I0.1) and a low detection limit of 54 nGy s1 (CsPbBr2I) for 120 keV hard X-rays. The CsPbBr2.9I0.1 detector demonstrated a stable current response with a dark current density of 0.58 μA cm2, providing clear imaging for 120 keV X-rays over 30 days under ambient conditions.
Additionally, the same group fabricated Cs1−mRbmPbBr3 single crystals by partially substituting Cs with Rb to enhance lattice distortion in CsPbBr3 and suppress ion migration. (Figure 2d–f) [40]. The Cs0.7Rb0.3PbBr3 single-crystal detector operated under 120 kVp X-rays. The Cs0.7Rb0.3PbBr3 detector showed a low dark current density of 1.01 μA cm2, a high sensitivity of 33,631 μC Gy1 cm2, and a low detection limit of 148 nGy s1. Furthermore, the detector demonstrated stability over 30 days highlighting its potential for X-ray detection applications.

2.3. All-Inorganic Lead-Free Perovskite Single Crystals for Direct Detection of X-Ray and γ-Ray

Although lead halide perovskites like CsPbBr3 hold significant promise for radiation detection, the presence of lead raises serious concerns regarding toxicity to human health and the environment [62]. To address this issue, elements such as Sn, Bi, and Cu can be used as substitutes for Pb. These materials exhibit γ-ray absorption capabilities similar to those of lead-based perovskites and have suitable band gaps for generating electron–hole pairs, making them potential alternatives for perovskite radiation detectors. Cs2AgBiBr6, like other halide perovskite systems, suffers from poor reproducibility and significant fluctuations in electrical properties. Yin et al. optimized the growth of Cs2AgBiBr6 to obtain crystals with smooth surfaces and high resistivity, ranging from 6.10 × 109 to 3.31 × 1010 Ω·cm, compared to unoptimized Cs2AgBiBr6, which showed a resistivity range of 6.04 × 107 to 5.61 × 109 Ω·cm. Under an electric field of 50 V mm1, the sensitivity of the X-ray detector was measured at 1974 µC Gy1 cm2, which is still not as capable as lead halide perovskite detectors [41].
Wei et al. prepared high-quality Cs3Cu2I5 SCs and Li-doped Cs3Cu2I5 SCs using the Bridgman method [42]. Li+ doping improved the carrier mobility from 6.49 to 9.52 cm2 V1 s1 by extending the carrier lifetime, resulting in an increase in the μτ product from 1.4 × 104 to 2.9 × 104 cm2 V1. This enhancement led to improved optoelectronic performance of Cs3Cu2I5 SCs. A vertical device structure composed of Au/Cs3Cu2I5 SC/PCBM/Au was fabricated as a high-sensitivity direct X-ray detector, achieving a sensitivity of 831.1 µC Gy1 cm2 and a low detection limit of 34.8 nGyairs1. This low-dimensional copper-based halide single crystal, typically regarded as an excellent scintillator for indirect detection, demonstrated its potential for direct detection in this study (Figure 3a–f).

3. Perovskite Single-Crystal Scintillators for Indirect Detection of X-Ray and γ-Ray

Beyond the requirement for high-Z materials, scintillators must satisfy several additional performance criteria to effectively detect γ-rays with high spectral resolution. Among these, light yield is the most critical parameter, defined as the number of visible photons generated per MeV of absorbed γ-ray energy. Light yield has a direct influence on energy resolution, with the theoretical limit being inversely proportional to the square root of the light yield. For example, at 662 keV, a light yield of 10,000 photons/MeV corresponds to a theoretical energy resolution limit of 2.5% [63].
Perovskite scintillators are available in various forms, such as nanocrystal and bulk crystal scintillators. In this context, we focus on the application of metal halide perovskite single crystals as scintillators in comparison to traditional bulk scintillators. To the best of our knowledge, LaBr3:Ce currently represents the highest-performing traditional scintillator, characterized by a light yield exceeding 60,000 ph/MeV, an energy resolution below 3%, and an exceptionally fast decay time of 35 ns [64]. However, this material suffers from high hygroscopicity, rendering it extremely unstable in ambient conditions, in addition to its intrinsic radioactive background and substantial production cost.
CsPbBr3 SCs, which are commonly used in direct conversion detectors, do not exhibit exceptional scintillation performance at room temperature; they require low temperatures (7 K) to achieve a light yield of 50,000 ph/MeV for 12 keV X-rays [47], which limits their practical applications. Recently, Cs3Cu2I5 has attracted significant attention due to its high quantum yield, positioning it as a promising scintillator material. Cs3Cu2I5 exhibits commendable scintillation properties along with excellent air stability, facilitating crystal processing and device fabrication. Furthermore, Cs3Cu2I5 contains only non-radioactive elements, thus avoiding the radioactive background associated with LaBr3:Ce. Lian et al. demonstrated the potential of Cs3Cu2I5 as an X-ray scintillator, noting its high PLQY of 73.7%, zero self-absorption and an emission peak that perfectly matches the SiPM response peak [65]. Cheng et al. grew Cs3Cu2I5 SCs from melt using the Bridgman method, reporting a scintillation emission peak at 440 nm and a primary decay time of 967 ns. Cs3Cu2I5 SC displayed a high scintillation yield of approximately 32,000 ph/MeV under X-ray radiation. Under γ-ray, they achieved a light yield of about 29,000 ph/MeV with an energy resolution of 3.4% at 662 keV [50].
Yao et al. successfully grew large, high-quality Cs3Cu2I5 SCs using a solution method, further improving the PLQY to 97.76%. The emission peak remained unchanged, ensuring compatibility with photomultiplier tubes (PMTs) and silicon photomultipliers (SiPMs). As a scintillator, Cs3Cu2I5 demonstrated excellent energy resolution and light yield under various γ-ray sources, highlighting its significant potential for practical applications. Additionally, a major breakthrough was achieved in the scintillation decay time, with a significantly increased proportion of the fast decay (39 ns) reaching 82%. Researchers concluded that ion doping in bulk single crystals could further enhance the scintillation properties [51].
Building on this, the doping of Cs3Cu2I5 SCs with ions such as In, Tl, Mn, and others has been explored [52,53,54,55]. Yao et al. reported a high-quality bulk Cs3Cu2I5 SC scintillator with an ultra-high light yield (95,772 ph/MeV for 137Cs), excellent energy resolution (3.79% at 662 keV), and ultra-fast scintillation decay time (3 ns) (Figure 4a–f) [55]. Mechanistic studies revealed that trace doping (at ppm levels) of heterovalent magnetic ions effectively modulated the emission dynamics of self-trapped excitons (STE). Firstly, the decay mechanism of Cs3Cu2I5 indicates that the emission of singlet STE requires electrons to be excited into the conduction band. However, under low-energy UV excitation (280 nm), electrons are trapped in the forbidden band due to the wide bandgap (3.63 eV) and high exciton binding energy (860.5 meV). The distorted zero-dimensional structure captures these excited-state electrons, forming STEs, while strong electron–phonon coupling alters their spin direction, leading to a triplet state. In contrast, gamma-ray excitation elevates electrons to higher energy levels, necessitating inelastic scattering and thermalization before they relax to the conduction band. These high-energy electrons migrate toward Mn2+, which facilitates directional migration due to its positive charge. When the electrons approach the magnetic atom (Mn2+), they can change their spin direction through exchange interactions or spin–orbit coupling. In this situation, the STEs contain a significant proportion of singlet states, resulting in a large percentage of ultrafast scintillation. Considering the comprehensive performance of the trace-doped Cs3Cu2I5 SC, including its outstanding scintillation properties, practical physical characteristics, and cost advantages, this material emerges as an ideal scintillator with significant commercial potential.

4. Perovskite Single Crystal for Detection of α-Particle/β-Particle/Neutron

Charged particles primarily interact with matter through Coulomb scattering with electrons in the atoms of a material. The scattered electrons convert their energy into electron–hole pairs. Due to the large mass of α-particles relative to electrons, they transfer only a small portion of their total kinetic energy in each collision. However, α-particles have a large cross-section, allowing them to interact with many electrons over a short distance, resulting in rapid energy loss and a short penetration depth. The interaction of β-particles is similar to that of α-particles, but owing to their smaller cross-section, β-particles have a longer penetration depth. Additionally, high-energy β-particles can lose energy through bremsstrahlung, emitting high-energy photons that may interact with detectors. For all charged particles, the cross-section increases with the atomic number (Z) of the material. However, since charged particles have lower penetration capabilities compared to high-energy photons, materials with a lower Z can also be used in detectors.
Detecting neutrons is more challenging than detecting charged particles because neutrons do not interact with matter through Coulomb forces. However, thermal neutrons can be effectively captured through nuclear reactions involving certain nuclei, such as 3He, 6Li, 10B, and 157Gd. These reactions produce secondary radiation, which enables the detection of thermal neutrons.
Previously, perovskite single crystals such as CsPbBr3 and Cs3Bi2I9 have been used for the direct conversion detection of α-particles (Figure 5a–c) [39,57,58,60]. For instance, Zhang et al. designed and fabricated three typical metal-semiconductor–metal nanostructured CsPbBr3 α-particle detectors. These detectors achieved an energy resolution of 5.70% and demonstrated long-term stability for over 8 h of continuous detection. This approach is simple and highly reproducible, offering a promising strategy to enhance the CsPbBr3 α-particle detectors.
Wang et al. demonstrated the significant potential of zero-dimensional Cs3Cu2I5 single crystals as discriminating scintillators for charged particles [60]. They showed that an exciton-capture strategy could greatly enhance the scintillation yield and improve α/β-ray discrimination ability (Figure 5d–f). Tl doping introduced Tl-bound excitons, effectively suppressing exciton–exciton quenching. As a result, Cs3Cu2I5 achieved a high scintillation yield of 26,000 photons/MeV under α-particle excitation, with an improved α/β pulse shape discrimination (PSD) figure of merit (FoM) of 2.64. For β-ray excitation, Cs3Cu2I5 showed a scintillation yield of 14,000 ph/MeV with an α/β PSD FoM of 2.08. Experimental results indicated that Cs3Cu2I5 SCs have the ability to detect radioactive 220Rn, showing great potential for environmental monitoring applications. With advantages such as non-hygroscopicity, scalability, low cost, and high performance, Cs3Cu2I5 emerges as a promising next-generation scintillator material for α/β detection and discrimination.

5. Conclusions and Future Perspectives

In summary, the field of perovskite radiation detectors has advanced rapidly, covering detectors for X-rays, γ-rays, α-particles, β-particles, and neutrons. Various perovskite materials—including organic–inorganic, all-inorganic, lead-based, and lead-free metal halide perovskites—have been preliminarily explored. Additionally, large-scale growth of high-quality single crystals using solution methods and the Bridgman method has been investigated.
Among these materials, organic–inorganic perovskites, despite offering advantages such as relatively good energy resolution for γ-rays (6.5% at 662 keV) and high sensitivity for X-rays (2.18 × 106 μC Gy1 cm2), suffer from inherent stability issues, which severely limit their potential for further development as radiation detectors. In contrast, all-inorganic perovskites, particularly CsPbBr3 single crystals, have emerged as the most promising materials for gamma ray detection. The energy resolution of γ-ray spectra for fully optimized CsPbBr3 detectors has improved to 1.4%, making them comparable to state-of-the-art commercial CZT detectors [45]. For future advancements in perovskite single-crystal radiation detectors, addressing the challenges of large-scale, high-quality crystal growth remains crucial. During the growth process, crystals may exhibit inhomogeneous purity or localized defects, leading to variations in device performance. To enhance X-ray detection sensitivity and crystal stability, researchers have explored the growth of perovskite single crystals through cation mixing (Cs+ and Rb+) and halide mixing (I, Br, and Cl). Furthermore, doping inorganic perovskites has been found to enhance carrier transport [38,43,66], suggesting that the incorporation of novel dopants could further improve the performance of perovskite detectors. Recent studies have shown that controlling the deuterium (D) content in organic cations can adjust detector sensitivity, detection limits, ion migration, and resistivity, thereby enhancing detector performance. This approach presents a direct strategy for developing ultrasensitive X-ray detection and imaging systems based on perovskite single crystals [35].
Concerns over environmental toxicity have driven the development of lead-free perovskite materials. Preliminary tests for radiation detection have been conducted using several lead-free inorganic metal halides, including Cs2AgBiBr6, Cs3Bi2I9, and Cs3Cu2I5 [39,41,42]. These materials exhibit stopping capabilities comparable to those of lead-based perovskites and offer higher stability, though further efforts are needed to utilize these materials effectively for radiation detection. In recent years, the copper-based halide perovskite Cs3Cu2I5 has gained significant attention as a scintillator for radiation detection, with research covering undoped crystals as well as those doped with In, Tl, and Mn. While the development of a highly bright, ultrafast, and low-cost scintillator remains a challenge in the field, a high-quality bulk Cs3Cu2I5:Mn SC scintillator has emerged with remarkable properties. This material demonstrates an exceptionally high light yield (95,772 ph/MeV for 137Cs γ-rays)—the highest reported to date—along with excellent energy resolution (3.79% at 662 keV) and an ultrafast scintillation decay time (3 ns, 81.5%). The combination of these outstanding scintillation properties, practical physical characteristics, and low production costs makes this single crystal an ideal candidate for commercial scintillation applications, warranting further exploration for commercialization. However, the time cost associated with the crystal growth process is significantly high, and the reproducibility of the experiments remains to be verified. These factors are substantial enough to impact the widespread application and commercialization of Cs3Cu2I5:Mn, which remain challenging.
There has also been progress in halide perovskite detectors for particle radiation. Compared to traditional detectors, halide perovskites offer advantages such as tunable compositions, flexible synthesis methods, low costs, and potential for large-scale production. However, perovskite-based particle radiation detectors still lag behind conventional materials, with energy resolutions for α, β, and neutron detectors significantly lower than the state-of-the-art and detection efficiencies below those of existing materials, necessitating further efforts to develop practical particle radiation detectors. Direct conversion perovskite α-particle detectors have demonstrated the ability to resolve energy spectra, though their energy resolution remains well below that of the best diamond-based detectors. Since β-particle detection is similar to X-ray/γ-ray detection, perovskites that have proven effective for X-ray/γ-ray detection could also serve as effective β-particle detectors. Currently, perovskite thermal neutron detectors are primarily focused on indirect conversion solid-state detectors, which achieve reasonable detection efficiency; however, future efforts should prioritize the development of direct conversion detectors with higher detection efficiencies. Recent studies have also shown that Cs3Cu2I5 single crystals can provide an effective exciton trapping strategy, significantly improving scintillation yield and α/β-ray discrimination ability and thereby opening new possibilities for achieving α/β particle detection and identification using lead-free perovskite scintillators [60].

Author Contributions

Conceptualization, Z.D. and G.L.; writing—original draft preparation, G.L.; writing—review and editing, M.P. and Z.Y.; supervision, C.P.C. and Z.D.; funding acquisition, Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (22075129).

Acknowledgments

The authors are grateful to Tao Zhang and Jian He from Nanjing University for their valuable discussions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nagaosa, N.; Tokura, Y. Topological properties and dynamics of magnetic skyrmions. Nat. Nanotechnol. 2013, 8, 899–911. [Google Scholar] [CrossRef] [PubMed]
  2. Bruno, A.; Bazilevskaya, G.A.; Boezio, M.; Christian, E.R.; Nolfo, G.A.d.; Martucci, M.; Merge’, M.; Mikhailov, V.V.; Munini, R.; Richardson, I.G.; et al. Solar Energetic Particle Events Observed by the PAMELA Mission. Astrophys. J. 2018, 862, 97. [Google Scholar] [CrossRef]
  3. Stoumpos, C.C.; Malliakas, C.D.; Peters, J.A.; Liu, Z.; Sebastian, M.; Im, J.; Chasapis, T.C.; Wibowo, A.C.; Chung, D.Y.; Freeman, A.J.; et al. Crystal Growth of the Perovskite Semiconductor CsPbBr3: A New Material for High-Energy Radiation Detection. Cryst. Growth Des. 2013, 13, 2722–2727. [Google Scholar] [CrossRef]
  4. Kim, J.; Pearton, S.J.; Fares, C.; Yang, J.; Ren, F.; Kim, S.; Polyakov, A.Y. Radiation damage effects in Ga2O3 materials and devices. J. Mater. Chem. C 2019, 7, 10–24. [Google Scholar] [CrossRef]
  5. Dionigi, F.; Zeng, Z.; Sinev, I.; Merzdorf, T.; Deshpande, S.; Lopez, M.B.; Kunze, S.; Zegkinoglou, I.; Sarodnik, H.; Fan, D.; et al. In-situ structure and catalytic mechanism of NiFe and CoFe layered double hydroxides during oxygen evolution. Nat. Commun. 2020, 11, 2522. [Google Scholar] [CrossRef]
  6. Chen, V.B.; Arendall, W.B., 3rd; Headd, J.J.; Keedy, D.A.; Immormino, R.M.; Kapral, G.J.; Murray, L.W.; Richardson, J.S.; Richardson, D.C. MolProbity: All-atom structure validation for macromolecular crystallography. Acta Crystallogr. D Biol. Crystallogr. 2010, 66, 12–21. [Google Scholar] [CrossRef]
  7. Kang, Y.; Zhou, X.E.; Gao, X.; He, Y.; Liu, W.; Ishchenko, A.; Barty, A.; White, T.A.; Yefanov, O.; Han, G.W.; et al. Crystal structure of rhodopsin bound to arrestin by femtosecond X-ray laser. Nature 2015, 523, 561–567. [Google Scholar] [CrossRef]
  8. D’Avanzo, P. Short gamma-ray bursts: A review. J. High Energy Astrophys. 2015, 7, 73–80. [Google Scholar] [CrossRef]
  9. Khan, A.I.; Shah, J.L.; Bhat, M.M. CoroNet: A deep neural network for detection and diagnosis of COVID-19 from chest x-ray images. Comput. Methods Programs Biomed. 2020, 196, 105581. [Google Scholar] [CrossRef]
  10. Wildenschild, D.; Sheppard, A.P. X-ray imaging and analysis techniques for quantifying pore-scale structure and processes in subsurface porous medium systems. Adv. Water Resour. 2013, 51, 217–246. [Google Scholar] [CrossRef]
  11. Wells, K.; Bradley, D.A. A review of X-ray explosives detection techniques for checked baggage. Appl. Radiat. Isot. 2012, 70, 1729–1746. [Google Scholar] [CrossRef] [PubMed]
  12. Zhou, Y.; Chen, J.; Bakr, O.M.; Mohammed, O.F. Metal Halide Perovskites for X-ray Imaging Scintillators and Detectors. ACS Energy Lett. 2021, 6, 739–768. [Google Scholar] [CrossRef]
  13. Liu, F.; Wu, R.; Wei, J.; Nie, W.; Mohite, A.D.; Brovelli, S.; Manna, L.; Li, H. Recent Progress in Halide Perovskite Radiation Detectors for Gamma-Ray Spectroscopy. ACS Energy Lett. 2022, 7, 1066–1085. [Google Scholar] [CrossRef]
  14. Burrows, D.N.; Hill, J.E.; Nousek, J.A.; Kennea, J.A.; Wells, A.; Osborne, J.P.; Abbey, A.F.; Beardmore, A.; Mukerjee, K.; Short, A.D.T.; et al. The Swift X-Ray Telescope. Space Sci. Rev. 2005, 120, 165–195. [Google Scholar] [CrossRef]
  15. Nemallapudi, M.V.; Gundacker, S.; Lecoq, P.; Auffray, E.; Ferri, A.; Gola, A.; Piemonte, C. Sub-100 ps coincidence time resolution for positron emission tomography with LSO:Ce codoped with Ca. Phys. Med. Biol. 2015, 60, 4635–4649. [Google Scholar] [CrossRef]
  16. Zhou, G.; Su, B.; Huang, J.; Zhang, Q.; Xia, Z. Broad-band emission in metal halide perovskites: Mechanism, materials, and applications. Mater. Sci. Eng. R Rep. 2020, 141, 100548. [Google Scholar] [CrossRef]
  17. Steele, J.A.; Puech, P.; Keshavarz, M.; Yang, R.; Banerjee, S.; Debroye, E.; Kim, C.W.; Yuan, H.; Heo, N.H.; Vanacken, J.; et al. Giant Electron–Phonon Coupling and Deep Conduction Band Resonance in Metal Halide Double Perovskite. ACS Nano 2018, 12, 8081–8090. [Google Scholar] [CrossRef] [PubMed]
  18. Manser, J.S.; Christians, J.A.; Kamat, P.V. Intriguing Optoelectronic Properties of Metal Halide Perovskites. Chem. Rev. 2016, 116, 12956–13008. [Google Scholar] [CrossRef]
  19. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with Wide Color Gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef]
  20. Akkerman, Q.A.; Manna, L. What Defines a Halide Perovskite? ACS Energy Lett. 2020, 5, 604–610. [Google Scholar] [CrossRef]
  21. Stranks, S.D.; Snaith, H.J. Metal-halide perovskites for photovoltaic and light-emitting devices. Nat. Nanotechnol. 2015, 10, 391–402. [Google Scholar] [CrossRef]
  22. Bao, C.; Yang, J.; Bai, S.; Xu, W.; Yan, Z.; Xu, Q.; Liu, J.; Zhang, W.; Gao, F. High Performance and Stable All-Inorganic Metal Halide Perovskite-Based Photodetectors for Optical Communication Applications. Adv. Mater. 2018, 30, 1803422. [Google Scholar] [CrossRef] [PubMed]
  23. Steele, J.A.; Pan, W.; Martin, C.; Keshavarz, M.; Debroye, E.; Yuan, H.; Banerjee, S.; Fron, E.; Jonckheere, D.; Kim, C.W.; et al. Photophysical Pathways in Highly Sensitive Cs2AgBiBr6 Double-Perovskite Single-Crystal X-Ray Detectors. Adv. Mater. 2018, 30, 1804450. [Google Scholar] [CrossRef] [PubMed]
  24. He, Y.; Matei, L.; Jung, H.J.; McCall, K.M.; Chen, M.; Stoumpos, C.C.; Liu, Z.; Peters, J.A.; Chung, D.Y.; Wessels, B.W.; et al. High spectral resolution of gamma-rays at room temperature by perovskite CsPbBr3 single crystals. Nat. Commun. 2018, 9, 1609. [Google Scholar] [CrossRef] [PubMed]
  25. Dong, H.; Zhang, C.; Liu, X.; Yao, J.; Zhao, Y.S. Materials chemistry and engineering in metal halide perovskite lasers. Chem. Soc. Rev. 2020, 49, 951–982. [Google Scholar] [CrossRef]
  26. Yakunin, S.; Sytnyk, M.; Kriegner, D.; Shrestha, S.; Richter, M.; Matt, G.J.; Azimi, H.; Brabec, C.J.; Stangl, J.; Kovalenko, M.V.; et al. Detection of X-ray photons by solution-processed organic-inorganic perovskites. Nat. Photonics 2015, 9, 444–449. [Google Scholar] [CrossRef]
  27. Yakunin, S.; Dirin, D.N.; Shynkarenko, Y.; Morad, V.; Cherniukh, I.; Nazarenko, O.; Kreil, D.; Nauser, T.; Kovalenko, M.V. Detection of gamma photons using solution-grown single crystals of hybrid lead halide perovskites. Nat. Photonics 2016, 10, 585–589. [Google Scholar] [CrossRef]
  28. Xu, Q.; Wei, H.; Wei, W.; Chuirazzi, W.; DeSantis, D.; Huang, J.; Cao, L. Detection of charged particles with a methylammonium lead tribromide perovskite single crystal. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2017, 848, 106–108. [Google Scholar] [CrossRef]
  29. Yu, D.; Wang, P.; Cao, F.; Gu, Y.; Liu, J.; Han, Z.; Huang, B.; Zou, Y.; Xu, X.; Zeng, H. Two-dimensional halide perovskite as beta-ray scintillator for nuclear radiation monitoring. Nat. Commun. 2020, 11, 3395. [Google Scholar] [CrossRef]
  30. El Bouanani, L.; Keating, S.E.; Avila-Avendano, C.; Reyes-Banda, M.G.; Pintor-Monroy, M.I.; Singh, V.; Murillo, B.L.; Higgins, M.; Quevedo-Lopez, M.A. Solid-State Neutron Detection Based on Methylammonium Lead Bromide Perovskite Single Crystals. ACS Appl. Mater. Interfaces 2021, 13, 28049–28056. [Google Scholar] [CrossRef]
  31. Su, Y.; Ma, W.; Yang, Y. Perovskite semiconductors for direct X-ray detection and imaging. J. Semicond. 2020, 41, 051204. [Google Scholar] [CrossRef]
  32. Chen, Q.; Wu, J.; Ou, X.; Huang, B.; Almutlaq, J.; Zhumekenov, A.A.; Guan, X.; Han, S.; Liang, L.; Yi, Z.; et al. All-inorganic perovskite nanocrystal scintillators. Nature 2018, 561, 88–93. [Google Scholar] [CrossRef]
  33. Wei, H.; Huang, J. Halide lead perovskites for ionizing radiation detection. Nat. Commun. 2019, 10, 1066. [Google Scholar] [CrossRef] [PubMed]
  34. Moseley, O.D.I.; Doherty, T.A.S.; Parmee, R.; Anaya, M.; Stranks, S.D. Halide perovskites scintillators: Unique promise and current limitations. J. Mater. Chem. C Mater. 2021, 9, 11588–11604. [Google Scholar] [CrossRef]
  35. Liu, L.; Xu, M.; Xu, X.; Tao, X.; Gao, Z. High Sensitivity X-ray Detectors with Low Degradation Based on Deuterated Halide Perovskite Single Crystals. Adv. Mater. 2024, 2406443. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, D.; Sun, X.; Jiang, L.; Jiang, X.; Chen, H.; Cui, F.; Zhang, G.; Wang, Y.; Lu, Y.-B.; Wu, Z.; et al. Ultralow detection limit and high sensitivity X-ray detector of high-quality MAPbBr3 perovskite single crystals. J. Mater. Chem. A 2024, 12, 12467–12474. [Google Scholar] [CrossRef]
  37. Hua, Y.; Zhang, G.; Sun, X.; Zhang, P.; Hao, Y.; Xu, Y.; Yang, Y.; Lin, Q.; Li, X.; Zhai, Z.; et al. Suppressed ion migration for high-performance X-ray detectors based on atmosphere-controlled EFG-grown perovskite CsPbBr3 single crystals. Nat. Photonics 2024, 18, 870–877. [Google Scholar] [CrossRef]
  38. Zhang, P.; Hua, Y.; Xu, Y.; Sun, Q.; Li, X.; Cui, F.; Liu, L.; Bi, Y.; Zhang, G.; Tao, X. Ultrasensitive and Robust 120 keV Hard X-ray Imaging Detector based on Mixed-Halide Perovskite CsPbBr3−nIn. Adv. Mater. 2022, 34, 2106562. [Google Scholar] [CrossRef]
  39. Sun, Q.; Xu, Y.; Zhang, H.; Xiao, B.; Liu, X.; Dong, J.; Cheng, Y.; Zhang, B.; Jie, W.; Kanatzidis, M.G. Optical and electronic anisotropies in perovskitoid crystals of Cs3Bi2I9 studies of nuclear radiation detection. J. Mater. Chem. A 2018, 6, 23388–23395. [Google Scholar] [CrossRef]
  40. Zhang, P.; Huang, L.; Li, L.; Chen, Z.; Zhang, G.; Tao, X. High-Flux Hard X-ray Cs1–mRbmPbBr3 Single-Crystal Detector with Suppressed Ion Migration. ACS Mater. Lett. 2024, 6, 4810–4818. [Google Scholar] [CrossRef]
  41. Yin, L.; Wu, H.; Pan, W.; Yang, B.; Li, P.; Luo, J.; Niu, G.; Tang, J. Controlled Cooling for Synthesis of Cs2AgBiBr6 Single Crystals and Its Application for X-ray Detection. Adv. Opt. Mater. 2019, 7, 1900491. [Google Scholar] [CrossRef]
  42. Wei, Q.; Fan, X.; Xiang, P.; Qin, L.; Liu, W.; Shi, T.; Yin, H.; Cai, P.; Tong, Y.; Tang, G.; et al. Cs3Cu2I5 Single Crystal for Efficient Direct X-ray Detection. Adv. Opt. Mater. 2023, 11, 2300247. [Google Scholar] [CrossRef]
  43. Wei, H.; DeSantis, D.; Wei, W.; Deng, Y.; Guo, D.; Savenije, T.J.; Cao, L.; Huang, J. Dopant compensation in alloyed CH3NH3PbBr3−xClx perovskite single crystals for gamma-ray spectroscopy. Nat. Mater. 2017, 16, 826–833. [Google Scholar] [CrossRef] [PubMed]
  44. He, Y.; Ke, W.; Alexander, G.C.B.; McCall, K.M.; Chica, D.G.; Liu, Z.; Hadar, I.; Stoumpos, C.C.; Wessels, B.W.; Kanatzidis, M.G. Resolving the Energy of γ-Ray Photons with MAPbI3 Single Crystals. ACS Photonics 2018, 5, 4132–4138. [Google Scholar] [CrossRef]
  45. He, Y.; Petryk, M.; Liu, Z.; Chica, D.G.; Hadar, I.; Leak, C.; Ke, W.; Spanopoulos, I.; Lin, W.; Chung, D.Y.; et al. CsPbBr3 perovskite detectors with 1.4% energy resolution for high-energy γ-rays. Nat. Photonics 2021, 15, 36–42. [Google Scholar] [CrossRef]
  46. He, Y.; Stoumpos, C.C.; Hadar, I.; Luo, Z.; McCall, K.M.; Liu, Z.; Chung, D.Y.; Wessels, B.W.; Kanatzidis, M.G. Demonstration of Energy-Resolved γ-Ray Detection at Room Temperature by the CsPbCl3 Perovskite Semiconductor. J. Am. Chem. Soc. 2021, 143, 2068–2077. [Google Scholar] [CrossRef]
  47. Mykhaylyk, V.B.; Kraus, H.; Kapustianyk, V.; Kim, H.J.; Mercere, P.; Rudko, M.; Da Silva, P.; Antonyak, O.; Dendebera, M. Bright and fast scintillations of an inorganic halide perovskite CsPbBr3 crystal at cryogenic temperatures. Sci. Rep. 2020, 10, 8601. [Google Scholar] [CrossRef]
  48. Shen, J.; Jia, R.; Hu, Y.; Zhu, W.; Yang, K.; Li, M.; Zhao, D.; Shi, J.; Lian, J. Cold-Sintered All-Inorganic Perovskite Bulk Composite Scintillators for Efficient X-ray Imaging. ACS Appl. Mater. Interfaces 2024, 16, 24703–24711. [Google Scholar] [CrossRef]
  49. Wang, C.; Yan, Z.-G.; Wang, Y.; Zhu, J.; Peng, C.; Qu, Y.; Yang, F.; Xiao, J.; Han, X. All-Inorganic Ruddlesden–Popper Perovskite Cs2CdCl4:Mn for Low-Dose and Flexible X-ray Imaging. ACS Mater. Lett. 2024, 6, 1429–1438. [Google Scholar] [CrossRef]
  50. Cheng, S.; Beitlerova, A.; Kucerkova, R.; Nikl, M.; Ren, G.; Wu, Y. Zero-Dimensional Cs3Cu2I5 Perovskite Single Crystal as Sensitive X-ray and γ-Rray Scintillator. Phys. Status Solidi (RRL)-Rapid Res. Lett. 2020, 14, 2000374. [Google Scholar] [CrossRef]
  51. Yao, Q.; Li, J.; Li, X.; Zheng, X.; Wang, Z.; Tao, X. High-Quality Cs3Cu2I5 Single-Crystal is a Fast-Decaying Scintillator. Adv. Opt. Mater. 2022, 10, 2201161. [Google Scholar] [CrossRef]
  52. Wang, Q.; Zhou, Q.; Nikl, M.; Xiao, J.; Kucerkova, R.; Beitlerova, A.; Babin, V.; Prusa, P.; Linhart, V.; Wang, J.; et al. Highly Resolved X-ray Imaging Enabled by In(I) Doped Perovskite-Like Cs3Cu2I5 Single Crystal Scintillator. Adv. Opt. Mater. 2022, 10, 2200304. [Google Scholar] [CrossRef]
  53. Yuan, D. Air-Stable Bulk Halide Single-Crystal Scintillator Cs3Cu2I5 by Melt Growth: Intrinsic and Tl Doped with High Light Yield. ACS Appl. Mater. Interfaces 2020, 12, 38333–38340. [Google Scholar] [CrossRef] [PubMed]
  54. Cheng, S.; Nikl, M.; Beitlerova, A.; Kucerkova, R.; Du, X.; Niu, G.; Jia, Y.; Tang, J.; Ren, G.; Wu, Y. Ultrabright and Highly Efficient All-Inorganic Zero-Dimensional Perovskite Scintillators. Adv. Opt. Mater. 2021, 9, 2100460. [Google Scholar] [CrossRef]
  55. Yao, Q.; Li, J.; Li, X.; Ma, Y.; Song, H.; Li, Z.; Wang, Z.; Tao, X. Achieving a Record Scintillation Performance by Micro-Doping a Heterovalent Magnetic Ion in Cs3Cu2I5 Single-Crystal. Adv. Mater. 2023, 35, 2304938. [Google Scholar] [CrossRef]
  56. Shimaoka, T.; Kaneko, J.H.; Tsubota, M.; Shimmyo, H.; Watanabe, H.; Chayahara, A.; Umezawa, H.; Shikata, S.-i. High-performance diamond radiation detectors produced by lift-off method. EPL (Europhys. Lett.) 2016, 113, 62001. [Google Scholar] [CrossRef]
  57. He, Y.; Liu, Z.; McCall, K.M.; Lin, W.; Chung, D.Y.; Wessels, B.W.; Kanatzidis, M.G. Perovskite CsPbBr3 single crystal detector for alpha-particle spectroscopy. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2019, 922, 217–221. [Google Scholar] [CrossRef]
  58. Zhang, M.; Yang, Y.; Xia, G.; Zou, J.; Deng, W.; Tian, F.; Deng, J. Metal–Semiconductor–Metal-Nanostructured CsPbBr3 Crystal Detector for Long-Term Stable α-Particle Detection. ACS Appl. Nano Mater. 2022, 5, 16039–16044. [Google Scholar] [CrossRef]
  59. McCall, K.M.; Liu, Z.; Trimarchi, G.; Stoumpos, C.C.; Lin, W.; He, Y.; Hadar, I.; Kanatzidis, M.G.; Wessels, B.W. α-Particle Detection and Charge Transport Characteristics in the A3M2I9 Defect Perovskites (A = Cs, Rb; M = Bi, Sb). ACS Photonics 2018, 5, 3748–3762. [Google Scholar] [CrossRef]
  60. Wang, Q.; Wang, C.; Shi, H.; Chen, J.; Yang, J.; Beitlerova, A.; Kucerkova, R.; Zhou, Z.; Li, Y.; Nikl, M.; et al. Exciton-harvesting enabled efficient charged particle detection in zero-dimensional halides. Light Sci. Appl. 2024, 13, 190. [Google Scholar] [CrossRef]
  61. Nayak, P.K.; Moore, D.T.; Wenger, B.; Nayak, S.; Haghighirad, A.A.; Fineberg, A.; Noel, N.K.; Reid, O.G.; Rumbles, G.; Kukura, P.; et al. Mechanism for rapid growth of organic-inorganic halide perovskite crystals. Nat. Commun. 2016, 7, 13303. [Google Scholar] [CrossRef] [PubMed]
  62. Li, J.; Cao, H.L.; Jiao, W.B.; Wang, Q.; Wei, M.; Cantone, I.; Lu, J.; Abate, A. Biological impact of lead from halide perovskites reveals the risk of introducing a safe threshold. Nat. Commun. 2020, 11, 310. [Google Scholar] [CrossRef] [PubMed]
  63. McGregor, D.S. Materials for Gamma-Ray Spectrometers: Inorganic Scintillators. Annu. Rev. Mater. Res. 2018, 48, 245–277. [Google Scholar] [CrossRef]
  64. van Loef, E.V.D.; Dorenbos, P.; van Eijk, C.W.E.; Krämer, K.; Güdel, H.U. High-energy-resolution scintillator: Ce3+ activated LaBr3. Appl. Phys. Lett. 2001, 79, 1573–1575. [Google Scholar] [CrossRef]
  65. Lian, L.; Zheng, M.; Zhang, W.; Yin, L.; Du, X.; Zhang, P.; Zhang, X.; Gao, J.; Zhang, D.; Gao, L.; et al. Efficient and Reabsorption-Free Radioluminescence in Cs3Cu2I5 Nanocrystals with Self-Trapped Excitons. Adv. Sci. 2020, 7, 2000195. [Google Scholar] [CrossRef]
  66. Li, J.; Du, X.; Niu, G.; Xie, H.; Chen, Y.; Yuan, Y.; Gao, Y.; Xiao, H.; Tang, J.; Pan, A.; et al. Rubidium Doping to Enhance Carrier Transport in CsPbBr3 Single Crystals for High-Performance X-Ray Detection. ACS Appl. Mater. Interfaces 2020, 12, 989–996. [Google Scholar] [CrossRef]
Figure 1. (ac) High–quality MAPbBr3 single crystals by vacuum evaporation crystallization method with a superior detection sensitivity of 24,552 mCGy−1 cm−2 and a low detection limit of 54 nGy s−1. (a) Photographs of DxMAPbI3 SCs with different degrees of deuteration. (b) Comparison of sensitivities of the DxMAPbI3 detectors. (c) X-ray dose−rate−dependent SNR at 0 bias and 120 keV hard X-rays. (d) Evolution in surface potential along marked lines in KPFM images of DxMAPbI3 SCs. Reprinted with permission from ref. [35]. Copyright 2024 Wiley. (e) Images of MAPbI3 single crystals and Schottky-type MAPbI3 detector achieves an excellent energy resolution of 6.8% for 57Co 122 keV gamma ray. (f) Comparison of 57Co γ-ray spectra collected over durations of 90 and 3600 s. Reprinted with permission from ref. [44]. Copyright 2018 American Chemical Society.
Figure 1. (ac) High–quality MAPbBr3 single crystals by vacuum evaporation crystallization method with a superior detection sensitivity of 24,552 mCGy−1 cm−2 and a low detection limit of 54 nGy s−1. (a) Photographs of DxMAPbI3 SCs with different degrees of deuteration. (b) Comparison of sensitivities of the DxMAPbI3 detectors. (c) X-ray dose−rate−dependent SNR at 0 bias and 120 keV hard X-rays. (d) Evolution in surface potential along marked lines in KPFM images of DxMAPbI3 SCs. Reprinted with permission from ref. [35]. Copyright 2024 Wiley. (e) Images of MAPbI3 single crystals and Schottky-type MAPbI3 detector achieves an excellent energy resolution of 6.8% for 57Co 122 keV gamma ray. (f) Comparison of 57Co γ-ray spectra collected over durations of 90 and 3600 s. Reprinted with permission from ref. [44]. Copyright 2018 American Chemical Society.
Inorganics 12 00278 g001
Figure 2. (a) As-grown single crystal ingot Φ11 mm and the single crystal wafers; (b) spectrum of 137Cs γ-ray source with 662 keV; (c) with different radiation sources, CsPbBr3 detectors show highly linear detection response. Reprinted with permission from ref. [24]. Copyright 2018 Springer Nature. (d) Photographs of as-grown Cs1−mRbmPbBr3 SCs; (e) illustration of Cs0.7Rb0.3PbBr3 detector (f). Comparison of the time-dependent responses of Br3, Rb0.1, Rb0.2, and Rb0.3 detectors to 120 kVp X-rays. Reprinted with permission from ref. [40]. Copyright 2024 American Chemical Society.
Figure 2. (a) As-grown single crystal ingot Φ11 mm and the single crystal wafers; (b) spectrum of 137Cs γ-ray source with 662 keV; (c) with different radiation sources, CsPbBr3 detectors show highly linear detection response. Reprinted with permission from ref. [24]. Copyright 2018 Springer Nature. (d) Photographs of as-grown Cs1−mRbmPbBr3 SCs; (e) illustration of Cs0.7Rb0.3PbBr3 detector (f). Comparison of the time-dependent responses of Br3, Rb0.1, Rb0.2, and Rb0.3 detectors to 120 kVp X-rays. Reprinted with permission from ref. [40]. Copyright 2024 American Chemical Society.
Inorganics 12 00278 g002
Figure 3. (a) X-ray response of the Cs3Cu2I5 and Cs3Cu2I5:Li detectors. Current densities of (b) Cs3Cu2I5 and (c) Cs3Cu2I5:Li detectors. (d) SNR of the detectors. (e) Stability of the Cs3Cu2I5:Li detector. (f) Optical and X-ray images of a metallic key. Reprinted with permission from ref. [42]. Copyright 2023 Wiley.
Figure 3. (a) X-ray response of the Cs3Cu2I5 and Cs3Cu2I5:Li detectors. Current densities of (b) Cs3Cu2I5 and (c) Cs3Cu2I5:Li detectors. (d) SNR of the detectors. (e) Stability of the Cs3Cu2I5:Li detector. (f) Optical and X-ray images of a metallic key. Reprinted with permission from ref. [42]. Copyright 2023 Wiley.
Inorganics 12 00278 g003
Figure 4. (a) Photograph of Cs3Cu2I5:Mn SC. (b,c) Scintillation decay curves of Cs3Cu2I5:Mn SC. (d) The light yield of Cs3Cu2I5:Mn SC under different sources. (e,f) Spectra of Cs3Cu2I5:Mn SC. (g) Decay mechanism. Reprinted with permission from ref. [55]. Copyright 2024 Wiley.
Figure 4. (a) Photograph of Cs3Cu2I5:Mn SC. (b,c) Scintillation decay curves of Cs3Cu2I5:Mn SC. (d) The light yield of Cs3Cu2I5:Mn SC under different sources. (e,f) Spectra of Cs3Cu2I5:Mn SC. (g) Decay mechanism. Reprinted with permission from ref. [55]. Copyright 2024 Wiley.
Inorganics 12 00278 g004
Figure 5. (a) Spectra of Cs3Bi2I9 devices under 241Am 5.5 MeV α-particle source. Reprinted with permission from ref. [59]. Copyright 2018 American Chemical Society. (b) Energy spectra of detectors Ti/Ni/CsPbBr3/Ni/Ti at various voltages for 600 s. Reprinted with permission from ref. [58]. Copyright 2022 American Chemical Society. (c) 241Am α-particle and γ-ray spectrum by In/CsPbBr3/Au detector. Reprinted with permission from ref. [57]. Copyright 2024 Elsevier. (d) Ionization density of Cs3Cu2I5 under α- and β-ray excitation. (e) Energy spectra. (f) Scintillation pulse waveforms. Reprinted with permission from ref. [60]. Copyright 2024 Springer Nature.
Figure 5. (a) Spectra of Cs3Bi2I9 devices under 241Am 5.5 MeV α-particle source. Reprinted with permission from ref. [59]. Copyright 2018 American Chemical Society. (b) Energy spectra of detectors Ti/Ni/CsPbBr3/Ni/Ti at various voltages for 600 s. Reprinted with permission from ref. [58]. Copyright 2022 American Chemical Society. (c) 241Am α-particle and γ-ray spectrum by In/CsPbBr3/Au detector. Reprinted with permission from ref. [57]. Copyright 2024 Elsevier. (d) Ionization density of Cs3Cu2I5 under α- and β-ray excitation. (e) Energy spectra. (f) Scintillation pulse waveforms. Reprinted with permission from ref. [60]. Copyright 2024 Springer Nature.
Inorganics 12 00278 g005
Table 1. Performance of Direct X-ray/γ-ray Detectors Based on Metal Halide Perovskite Single Crystals.
Table 1. Performance of Direct X-ray/γ-ray Detectors Based on Metal Halide Perovskite Single Crystals.
MaterialsGrowth Method αDevice Structureμτ Product
(cm2 V−1)
Sensitivity
(μC Gy−1 cm−2)
Detection Limit
(nGyair s−1)
Voltage or Electric FieldEnergy ResolutionRef.
X-ray detectors (120 keV)
DxMAPbI3ITC methodAu/DxMAPbI3/Au5.39 × 10−22.18 × 1064.840 V/mm [35]
MAPbBr3ITC methodPt/MAPbBr3/Pt1.49 × 10−424,5225480 V/cm [36]
CsPbBr3EFG methodEGaIn/CsPbBr3/Au8.11 × 10−446,18010.815000 V/cm [37]
CsPbBr2.9I0.1BridgmanAu/CsPbBr2.9I0.1/Au5.06 × 10−362,7481175000 V/cm [38]
Cs3Bi2I9BridgmanAu/Cs3Bi2I9/Au2.03 × 10−5 111.9 2800 V/cm [39]
Cs0.7Rb0.3PbBr3BridgmanInGa/Cs0.7Rb0.3PbBr3/Au 33,6311485000 V/cm [40]
Cs2AgBiBr6Solution Au/Cs2AgBiBr6/Au5.95 × 10−3194745.7500 V/cm [41]
Cs3Cu2I5:LiBridgmanAu/Cs3Cu2I5:Li/PCBM/Au2.9 × 10−4831.134.8450 V/cm [42]
γ-ray detectors
MAPbBr2.94Cl0.06ITC methodCr/MAPbBr2.94Cl0.06/C60/BCP/Cr1.8 × 10−2 10 V6.5% (137Cs 662 keV)[43]
MAPbI3TRM methodGa/MAPbI3/Au8.1 × 10−4 (h) 70 V,
460 V/cm
6.8%
(57Co 122 keV)
[44]
7.4 × 10−4 (e)12%
(241Am 59.5 keV)
CsPbBr3BridgmanGa/CsPbBr3/Au1.34 × 10−3 150 V3.9%
(57Co 122 keV)
[24]
900 V3.8%
(137Cs 662 keV)
CsPbBr3BridgmanEGaIn/CsPbBr3/Au8 × 10−3 5500 V/cm1.4%
(137Cs 662 keV)
[45]
CsPbCl3BridgmanGa/CsPbCl3/Au3.2 × 10−4 300 V~16%
(57Co 122 keV)
[46]
α ITC, inverse-temperature crystallization; TRM, temperature-rising method; EFG, edge-defined film-fed growth method; BCP, bathocuproine; PCBM, [6,6]-phenyl-C61-butyric acid methyl ester; EGaIn, eutectic gallium–indium.
Table 2. Scintillation Properties of Metal Halide Scintillator Single Crystals.
Table 2. Scintillation Properties of Metal Halide Scintillator Single Crystals.
MaterialsGrowth MethodMaximum Emission
(nm)
Light Yield
(Photons/MeV)
Decay Time
(ns)
Detection Limit
(nGyair s−1)
Energy Resolution or Spatial Resolution
(X-Ray)
(lp mm−1)
Radiation
Source
Ref.
CsPbBr3Bridgman535, 545~50,000
(7 K)
~1 X-ray[47]
CsPbBr3/Cs4PbBr6Cold sintering51533,8009.87919.3%241Am 59.6 keV[48]
8.9 lp mm−1X-ray
Cs2CdCl4:MnHydrothermal method58588,138 31.0416.1 lp mm−1X-ray[49]
Cs3Cu2I5Bridgman44032,000967 X-ray
137Cs 662 keV
[50]
29,0003.4%
Cs3Cu2I5Solution method44532,69539 4.45%137Cs 662 keV[51]
3.03%60Co 1332 keV
8.15%152Eu 334 keV
Cs3Cu2I5:InBridgman460, 62053,000370096.218 lp mm−1X-ray[52]
44,000137Cs 662 keV
Cs3Cu2I5:TlBridgman~450, 51751,000893 4.5%137Cs 662 keV[53]
Cs3Cu2I5:TlBridgman440, 51087,00071766.33.4%137Cs 662 keV[54]
Cs3Cu2I5:MnSolution44595,7723 3.79%137Cs 662 keV[55]
2.63%60Co 1332 keV
Table 3. α-particle/β-particle/neutron detection.
Table 3. α-particle/β-particle/neutron detection.
Direct ConversionGrowth MethodDevice Structureμτ Product
(cm2 V−1)
Energy ResolutionRadiation SourceRef.
DiamondCVD βAl/diamond/TiC/Au3.1 × 10−4 (h)0.3%α-particle (241Am 5.5 MeV)[56]
MAPbI3TRM methodGa/MAPbI3/Au8.1 × 10−4 (h) α-particle (241Am 5.5 MeV)[44]
7.4 × 10−4 (e)
CsPbBr3BridgmanIn/CsPbBr3/Au4.5 × 10−4 (e) α-particle (241Am 5.5 MeV)[57]
9.5 × 10−4 (h)
CsPbBr3BridgmanTi/Ni/CsPbBr3/Ni/Ti6.4 × 10−3 (e)5.70% (700 V)α-particle (241Am 5.5 MeV)[58]
Cs3Bi2I9BridgmanAu/Cs3Bi2I9/Au5.4 × 10−5 (e) α-particle (241Am 5.5 MeV)[59]
MAPbBr3ITC methodAg/In/Ga2O3/MAPbBr3/Au Efficiency 3.92%Neutron(252Cf)[30]
ScintillatorGrowth MethodLight Yield
(Photons/MeV)
Wavelength
(nm)
Decay Time
(ns)
Energy ResolutionRadiation SourceRef.
STA2PbBr4:MnSolution~24,000610500 β-particle(63Ni 66.7 keV)[29]
Cs3Cu2I5:TlBridgman26,000445, 51070810.5%α-particle (241Am 5.5 MeV)[60]
β CVD, chemical vapor deposition.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luo, G.; Peng, M.; Yang, Z.; Chu, C.P.; Deng, Z. Emerging New-Generation Semiconductor Single Crystals of Metal Halide Perovskites for Radiation Detection. Inorganics 2024, 12, 278. https://doi.org/10.3390/inorganics12110278

AMA Style

Luo G, Peng M, Yang Z, Chu CP, Deng Z. Emerging New-Generation Semiconductor Single Crystals of Metal Halide Perovskites for Radiation Detection. Inorganics. 2024; 12(11):278. https://doi.org/10.3390/inorganics12110278

Chicago/Turabian Style

Luo, Guigen, Min Peng, Zhibin Yang, Chungming Paul Chu, and Zhengtao Deng. 2024. "Emerging New-Generation Semiconductor Single Crystals of Metal Halide Perovskites for Radiation Detection" Inorganics 12, no. 11: 278. https://doi.org/10.3390/inorganics12110278

APA Style

Luo, G., Peng, M., Yang, Z., Chu, C. P., & Deng, Z. (2024). Emerging New-Generation Semiconductor Single Crystals of Metal Halide Perovskites for Radiation Detection. Inorganics, 12(11), 278. https://doi.org/10.3390/inorganics12110278

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop