Fact-checked by Grok 2 weeks ago

X-ray detector

An detector is a device that captures and converts incoming photons into a measurable signal, typically electrical charge or light, to produce images or spectra for analysis. These detectors operate on principles such as photoelectric , where an photon interacts with a to eject electrons, generating a current or charge proportional to the photon's energy. Common types include gas-filled proportional counters, which use ionized gas to amplify signals in chambers with electrodes; detectors like charge-coupled devices (CCDs) or cadmium-zinc-telluride (CdZnTe) that directly create electron-hole pairs; and scintillator-based systems that convert X-rays to visible light before detection. In medical applications, traditional screen-film detectors use fluorescent screens to expose , while modern digital flat-panel detectors (FPDs) employ indirect conversion with cesium iodide () scintillators or direct conversion with photoconductors for real-time digital output. X-ray detectors vary in performance metrics, including (the fraction of photons detected, often high for at soft X-ray energies around 8 keV but decreasing at higher energies), (down to 100–150 µm in FPDs or 110 µm in pixel array detectors), and energy resolution (as fine as 270 eV in spectroscopic models). Advanced integrating detectors, such as the Cornell-SLAC Pixel Array Detector (CSPAD), achieve frame rates up to 120 Hz, while photon-counting detectors like PILATUS exceed 1 MHz per pixel count rates, enabling ultrafast imaging. Materials like , CdTe, and scintillators (e.g., ) are selected based on the X-ray energy range, from soft X-rays (<10 keV) in medical radiography to hard X-rays (>20 keV) in . These detectors are essential across fields: in for , , and computed (CT) to visualize anatomy with reduced patient dose via digital systems; in astronomy for satellites like and to study cosmic phenomena; and in scientific research for protein crystallography, analysis, and time-resolved studies at synchrotrons and free-electron lasers. As of 2025, advancements have achieved higher efficiency for hard X-rays, readout speeds exceeding 1000 Hz in detectors, and improved energy resolution, with emerging materials enabling lower-dose imaging to support applications like nanoscale imaging and dynamic material studies.

Principles of Operation

X-ray Interactions with Matter

X-rays interact with matter through several fundamental processes that determine the and of photons in detector materials, leading to energy deposition that can be detected. These interactions are probabilistic, governed by cross-sections that depend on the and the of the material. In the context of X-ray detectors, understanding these processes is essential for optimizing material selection and detector efficiency, particularly in the diagnostic energy range of 10-150 keV used in . The is a primary interaction where an incident is completely absorbed by an , transferring its to an inner-shell , which is then ejected as a photoelectron. The excess , after overcoming the electron's , becomes the of the photoelectron, while the is left in an , often leading to the emission of characteristic or electrons as it relaxes. The probability of this interaction, described by the photoelectric cross-section \sigma_{pe}, scales approximately as \sigma_{pe} \propto Z^4 / E^{3.5}, where Z is the of the material and E is the ; this strong dependence on Z makes high-Z materials like lead or iodine particularly effective for absorbing low-energy . Compton scattering, or inelastic scattering, occurs when an collides with a loosely bound outer-shell , transferring part of its to the ( ) and at an with reduced . This dominates in lower-Z materials like or at energies above about 30 keV, contributing to in detectors due to the scattered 's deviation from the original path. The wavelength shift of the scattered is given by the Compton formula: \Delta \lambda = \frac{h}{m_e c} (1 - \cos \theta) where h is Planck's constant, m_e is the , c is the , and \theta is the scattering angle; the Compton wavelength h/(m_e c) \approx 0.00243 nm quantifies the maximum shift at \theta = 180^\circ. The cross-section for is proportional to Z / E, decreasing with increasing energy. Coherent scattering, also known as , involves elastic interaction of the photon with the entire atom, resulting in a change of direction without energy loss to the atom. This process is coherent because the scattered photon's phase remains related to the incident wave, and it contributes minimally to signal generation in detectors since no net energy is deposited. Its cross-section scales with Z^2 and is significant only at low energies (<30 keV) in high-Z materials, but it is generally negligible compared to other interactions in typical detector applications. Pair production becomes possible only when the X-ray photon energy exceeds 1.022 MeV, the rest mass energy of an electron-positron pair (twice 0.511 MeV), converting the into an electron-positron pair in the field of the . Above this , the cross-section increases logarithmically with energy and scales with Z^2, but this interaction is irrelevant for diagnostic detectors, as typical energies (10-150 keV) are well below the . The overall attenuation of an X-ray beam through a is described by the Beer-Lambert law: I = I_0 e^{-\mu x} where I is the transmitted intensity, I_0 is the incident intensity, \mu is the (sum of contributions from all interactions), and x is the thickness; the \mu/\rho normalizes for density \rho and similarly sums the mass cross-sections for photoelectric, , coherent, and processes. In detector , \mu varies strongly with energy: photoelectric absorption dominates at lower energies (e.g., <50 keV), enhancing in high-Z absorbers, while prevails at higher energies (50-150 keV), reducing efficiency in low-Z like . This energy dependence influences detector design, as higher energies require thicker or higher-Z layers to achieve sufficient absorption.

Signal Generation and Readout

In X-ray detectors, signal generation begins with the conversion of absorbed energy into detectable forms through , indirect, or modes. conversion occurs in materials where X-rays create electron-hole pairs that are directly collected as electrical charge, bypassing intermediate production. Indirect conversion involves scintillators that first transform X-ray energy into visible photons, which are then detected by photodiodes or similar devices to generate charge. modes, such as in photostimulable phosphor systems, trap energy in the form of a during exposure, which is later released as upon stimulation for readout. Charge collection in direct-conversion semiconductor detectors relies on the drift and diffusion of electron-hole pairs generated by X-ray interactions, primarily through photoelectric absorption. Electrons and holes migrate under an applied electric field, with efficiency determined by the mobility-lifetime product (μτ), which quantifies how far carriers travel before recombination; higher μτ values, such as 10^{-3} to 10^{-2} cm²/V in materials like CdTe, enable better charge transport and reduced trapping losses. In indirect detectors, charge arises from photoelectrons produced by scintillator-emitted light interacting with photodetectors, though collection is influenced by optical coupling and quantum efficiency rather than direct carrier mobility. The process in indirect detectors involves excitation of atoms by energy deposition, followed by de-excitation that emits light photons. Light yield is proportional to the energy deposited (E), often expressed as N = Y · E, where N is the number of photons and Y is the yield in photons per keV, typically ranging from 10 to 100 photons/keV for common scintillators like CsI:Tl. This proportionality ensures that the optical signal scales with incident intensity, though factors like non-radiative recombination can reduce efficiency. Readout electronics in digital X-ray detectors, such as those using (TFT) arrays, capture and process generated charges by sequentially addressing pixels during integration periods, converting analog signals to digital via analog-to-digital converters (ADCs). Noise sources include dark current from thermal generation, readout noise from TFT switching, and from quantum statistics, which degrade performance; the (SNR) is given by SNR = S / √Var, where S is the signal and Var is the total variance from these sources. Integration times, often milliseconds for flat-panel systems, balance sensitivity against . Amplification stages enhance weak signals in certain detectors; image intensifiers employ multi-stage acceleration with factors up to 10^4 through electrostatic focusing and phosphor screens, while in semiconductors like or perovskites provides internal via , achieving factors of 10 to 1000 to overcome electronic noise limits. Historically, X-ray detector readout evolved from analog systems, such as film-screen or chains in image intensifiers, which suffered from geometric distortion and limited , to modern digital sampling using high-speed ADCs (e.g., 12-16 bits at kHz rates) in TFT-based arrays, enabling precise quantization and post-processing since the .

Detectors for X-ray Imaging

Screen-Film Systems

Screen-film systems represent the traditional analog method for X-ray detection in radiography, utilizing to capture and record images. These systems typically employ a double-emulsion film sandwiched between two intensifying screens, which enhance sensitivity by converting into visible light. The intensifying screens are commonly made of materials such as calcium tungstate (CaWO₄), which fluoresce upon absorption, emitting light photons that expose the film's emulsion layers on both sides. This configuration reduces the required exposure by a factor of 50 to 100 compared to direct film exposure alone, thereby minimizing patient dose. The operational mechanism relies on the formation of a in the film's crystals, primarily (AgBr), which are embedded in a . When s interact with the intensifying screens, they produce light that sensitizes the grains; additionally, direct absorption in the can generate photoelectrons that reduce silver ions to form metallic silver specks. This remains invisible until chemical development, where unexposed silver halides are removed, leaving metallic silver grains that create areas of opacity corresponding to the exposure pattern. The system's response to exposure is characterized by the Hurter and Driffield () curve, which plots optical density against the logarithm of exposure, illustrating the film's (low exposure), linear (useful range), and (high exposure) regions, with average determining . Sensitivity and speed vary across film types, such as par-speed (medium , relative speed ~100 with calcium screens) and ultra-speed (higher , relative speed ~400), quantified by exposure indices that reflect the reciprocity between film and screen. The of these systems, measured as (DQE), is approximately 20-30%, limited by incomplete absorption and light diffusion in the screens. reaches 10-15 line pairs per millimeter (lp/mm), enabling fine detail visualization, though thicker screens for faster speeds reduce this to favor . Advantages of screen-film systems include high suitable for detecting subtle anatomical details and relatively low cost for initial setup and operation. However, they require wet chemical processing, which generates silver-laden waste and demands facilities, while the limited —typically a of 50:1—necessitates precise exposure control to avoid underexposure or overexposure artifacts. The gradual replacement by systems began in the 1980s with early computed radiography prototypes and accelerated through the 1990s and 2000s due to improved workflow efficiency and reduced environmental impact; by the 2010s, adoption was widespread in developed regions, though screen-film remains in use in some developing areas as of the 2020s for its simplicity and affordability.

Photostimulable Phosphor Systems

Photostimulable phosphor systems, also known as computed radiography (CR), represent an early form of digital imaging that bridges analog and fully digital technologies by using reusable phosphor plates to capture and store latent images. These systems were pioneered by Fuji in 1983, with the introduction of the Fuji Computed Radiography (FCR) system, which utilized scanning laser-stimulated to read out stored energy. The technology enables the conversion of exposure into a without the need for traditional processing, offering a wide latitude for exposure variations. The core material in these systems is a photostimulable , typically fluorobromide doped with (BaFBr:Eu²⁺), which forms a layered capable of trapping charge carriers. This is coated onto a flexible plate, often ~200-300 μm thick for standard use, with the Eu²⁺ activator creating color centers that enhance storage efficiency. Variations such as BaF(Br,I):Eu²⁺ or BaFI:Eu²⁺ may be employed to optimize across energy ranges. In operation, incident s generate -hole pairs within the lattice, where are trapped in F-centers ( traps associated with vacancies), forming a proportional to the local radiation dose. A scanning , commonly a helium-neon (HeNe) at 633 nm , stimulates the trapped , promoting them to the conduction band for recombination with holes at Eu²⁺ centers, resulting in () emission around 390 nm. The intensity is directly proportional to the trapped charge and thus the X-ray exposure dose, allowing quantitative readout. The system comprises an imaging plate housed in a light-tight cassette, a dedicated reader unit, and an erasure mechanism. In the reader, the plate is scanned line-by-line by the laser beam (spot size ~50-100 μm), with emitted PSL collected via a fiber optic light guide and detected by a photomultiplier tube (PMT) to produce an analog signal that is digitized into a pixel array (typically 10-bit or higher depth). Post-readout, residual trapped charge is erased by flooding the plate with intense white light (~450-700 nm), discharging any remaining centers for reuse. Performance characteristics include a exceeding 10,000:1, enabling detection of exposures from ~0.01 mR to over 100 mR without saturation, far surpassing screen-film systems. typically ranges from 2.5 to 5 lp/mm for general , with high-resolution plates achieving up to 10 lp/mm for applications like , limited primarily by spot size and layer thickness. Compared to traditional screen-film systems, photostimulable plates offer reusability for thousands of cycles and eliminate wet chemical processing, reducing environmental impact and workflow time. However, they exhibit lower than modern direct digital detectors and are susceptible to plate wear from repeated handling and mechanical stress.

Image Intensifier Systems

Image intensifier systems are vacuum-tube devices that enable real-time X-ray imaging by converting low-intensity X-ray patterns into bright visible images, primarily used in dynamic procedures such as . These systems amplify the signal electronically, allowing for continuous observation without the high radiation doses required for direct visual fluoroscopy. Introduced in the late 1940s, they revolutionized by facilitating procedures like and , with peak adoption occurring from the 1970s to the 1990s. The core structure consists of an input phosphor layer, typically cesium iodide (CsI) about 300-500 µm thick, which absorbs X-rays and converts them to visible light with up to 70% efficiency. This light strikes a photocathode, often cesium-antimony (Cs₃Sb), generating photoelectrons that are accelerated and focused by electron optics at 20-30 kV toward a smaller output phosphor, usually zinc-cadmium sulfide (ZnCdS:Ag) 4-8 µm thick, where they produce a intensified visible image. The output is coupled to a television camera for video display or recorded on film, enabling real-time monitoring at frame rates up to 30 fps for adequate temporal resolution in motion imaging. Performance is characterized by a brightness of up to 10,000, calculated as the product of minification (from the reduction in size, e.g., input diameter squared over output diameter squared) and flux (from increasing photons per ). reaches about 1-2 line pairs per millimeter (lp/mm) at the periphery, though limited by pincushion distortion arising from the curved input mapping to a flat output, which stretches peripheral features. Veiling , caused by scattered within the tube, reduces contrast by 10-20%. In applications like and labs, these systems provide essential real-time visualization of vascular structures with contrast agents, supporting precise guidance during interventions. However, their bulky design (input fields of 15-57 cm), higher doses compared to modern alternatives, and artifacts like glare and distortion have led to replacement by flat-panel detectors since the .

Direct Conversion Detectors

Direct conversion detectors are semiconductor-based devices that absorb photons and generate electrical charge carriers directly through the , without the intermediate production of visible light. This approach leverages the high () materials in semiconductors to achieve efficient X-ray absorption, particularly for diagnostic energies in the 10–100 keV range, enabling compact designs with minimal signal loss. Common materials for these detectors include amorphous selenium (a-Se), (CdTe), and mercuric iodide (HgI₂). Amorphous selenium, with selenium's Z=34, offers good absorption for low-energy applications while being compatible with large-area deposition techniques like thermal evaporation. CdTe, benefiting from higher Z values (Cd Z=48, Te Z=52), provides superior for higher-energy s due to its and bandgap properties.00568-4) Mercuric iodide (HgI₂) is another high-Z compound ( Z=80, I Z=53) valued for its room-temperature operation and direct bandgap, though it is less commonly used in commercial systems owing to fabrication challenges. In operation, an incident is absorbed in the layer, creating electron-hole pairs with an average of approximately eV per pair in a-Se. An applied , typically around 10 V/μm, separates and drifts these charge carriers to oppositely charged electrodes, producing a measurable proportional to the intensity. Unlike indirect detectors, this direct process avoids spreading, preserving spatial fidelity and enabling high-resolution imaging. Readout is achieved using pixelated (TFT) arrays integrated beneath the conversion layer, where each pixel collects and stores charge until readout via row and column addressing. These arrays often feature fill factors exceeding 90%, minimizing inactive areas and maximizing . (DQE) can reach up to 0.7 at low energies (e.g., 20–30 keV), reflecting efficient charge collection and low noise.00470-9/fulltext) These detectors are primarily applied in , where a-Se-based systems received FDA approval in the early 2000s, revolutionizing breast imaging with reduced dose and improved contrast. They are also used in for intraoral imaging, offering portability and . Key advantages include spatial resolutions greater than 10 line pairs per millimeter (lp/mm) and low inter-pixel crosstalk, supporting detailed depiction of fine structures like microcalcifications. Challenges include charge trapping in amorphous materials, which reduces collection efficiency and signal uniformity, particularly under prolonged bias. effects, arising from buildup, can distort the internal field and degrade long-term performance. Dark current management is critical, as thermal generation in these semiconductors must be suppressed to maintain low noise, often through blocking layers or optimized biasing.

Indirect Conversion Detectors

Indirect conversion detectors operate by first absorbing s in a material, which promptly converts the energy into visible photons, and then detecting those photons with an underlying array to generate electrical signals. This two-step process contrasts with direct conversion methods by introducing an intermediate optical stage, which can introduce blurring but allows for higher due to the use of high materials in the .00470-9/fulltext) The is typically emitted in the , with wavelengths ranging from 400 to 700 , matching well with the sensitivity of silicon-based photodiodes. Common scintillator materials include thallium-doped cesium iodide (CsI(Tl)), which is deposited as structured needle-like columns (approximately 5–10 µm in diameter) to minimize lateral light spreading and preserve , and oxysulfide (Gd₂O₂S, often doped with , known as GOS) in powder form for more cost-effective but less structured applications. In operation, s are absorbed primarily via photoelectric interactions in the layer (typically 200–550 µm thick), producing thousands of visible light photons per absorbed , which are then coupled via direct deposition or fiber optic tapers to an array of (a-Si) photodiodes. The resulting is stored in capacitors and read out using a (TFT) matrix, enabling digital image acquisition with pixel sizes of 100–200 µm. (DQE) for these systems reaches 0.6–0.8 at low spatial frequencies under typical diagnostic energies (e.g., 70 kVp), reflecting efficient capture and signal transfer. The readout process involves sequential scanning of TFT rows, converting the charge to voltage and digitizing it at 12–16 bits per , yielding a of approximately 1000:1 suitable for general radiography applications. However, light scattering within the degrades the modulation transfer function (), particularly in unstructured materials like GOS, where the spread is often modeled by a Gaussian () with a of 0.3–0.5 mm, limiting to about 5–7 line pairs per millimeter (lp/mm). These detectors were first introduced commercially in 1995, with early examples including GE's amorphous silicon-based systems for flat-panel . Compared to direct conversion detectors using amorphous selenium (a-Se), indirect systems offer higher X-ray absorption due to the higher effective atomic numbers in scintillators like (Z ≈ 55 for , 53 for I), enabling thicker layers without excessive voltage requirements and thus better overall for higher-energy X-rays.00470-9/fulltext) Limitations include the resolution loss from light diffusion, which is mitigated but not eliminated by columnar structures, making them particularly advantageous for applications prioritizing dose efficiency and over the highest spatial frequencies.

Detectors for Dose Measurement

Gas-Filled Detectors

Gas-filled detectors operate on the principle of ionizing a gas medium with X-rays, collecting the resulting charge under an applied to measure dose. These devices are particularly suited for in and calibration due to their simplicity and ability to provide accurate measurements of or air kerma. X-rays interact with the gas atoms primarily through , producing photoelectrons that ionize the gas molecules and create electron-ion pairs. The primary types of gas-filled detectors include ionization chambers, proportional counters, and Geiger-Müller counters, distinguished by their operating voltage and degree of charge amplification. Ionization chambers function at low voltages (typically 50-300 V), where the collected charge equals the number of primary ion pairs without significant recombination or multiplication, ensuring recombination-free operation. Proportional counters operate at higher voltages (around 1000-3000 V), enabling gas amplification through Townsend avalanche, with gains of approximately 10^3 to 10^4, while maintaining proportionality to the initial energy deposited. Geiger-Müller counters apply even higher voltages (about 900 V), resulting in self-quenching avalanches with gains near 10^6, but they saturate and provide no energy information. Common fill gases include dry air for ionization chambers, argon-methane mixtures (P-10 gas, 90% Ar and 10% CH4) for proportional counters, and neon or xenon with quenching agents (e.g., alcohol or halogens) for Geiger-Müller counters to prevent continuous discharge. In operation, incident X-rays ionize the gas, and the drifts the electrons and ions to electrodes, generating a measurable or . The D (in ) is calculated as D = \frac{Q}{m} \cdot \frac{W}{e}, where Q is the collected charge (in C), m is the mass of the gas (in kg), W is the average energy required to produce an ion pair (approximately 33.97 /ion pair in air), and e is the (1.602 × 10^{-19} C); equivalently, using W/e \approx 33.97 J/C. These detectors are often configured in cylindrical or parallel-plate geometries to optimize field uniformity and collection efficiency. An early for X-ray measurements was developed by in 1913 for his work on X-ray diffraction. Performance characteristics include an energy resolution of about 10-20% full width at half maximum (FWHM) for proportional counters detecting X-rays, limited by statistical fluctuations in ion pair production. Ionization chambers excel in air kerma measurements (SI unit: gray, Gy), serving as primary standards for calibrating X-ray beams up to several hundred keV. Applications encompass radiation protection monitoring, such as survey meters for dose rates from background to 50 rem/hr, and calibration of X-ray sources in medical and industrial settings. Limitations include low detection for high-energy X-rays due to the low of the gas medium, as well as sensitivity to and variations, which affect gas and thus charge collection. Thin windows (e.g., ) are required to minimize X-ray attenuation, but overall, these detectors are less efficient than solid-state alternatives for high-flux or energetic beams.

Semiconductor Dosimeters

Semiconductor dosimeters are solid-state devices that employ , primarily , to quantify doses by detecting charge generated from ionizing interactions within the material. These detectors leverage the high density of silicon to produce substantial charge signals relative to gas-based systems, enabling compact designs for accurate dose measurement without spatial resolution focus. Unlike thermoluminescent dosimeters, they provide readout, facilitating immediate dose assessment in clinical and scenarios. Key types include diodes, often derived from silicon solar cell structures, standard p-n diodes, and metal-oxide-semiconductor field-effect transistors (MOSFETs). diodes feature a p-type and n-type interface forming a , while MOSFETs utilize a layer where traps charges to alter device characteristics. Silicon-based examples predominate due to their and well-characterized response to X-rays. Operation relies on X-ray photons generating electron-hole pairs in the semiconductor's through photoelectric absorption and ; the built-in sweeps these carriers to electrodes, yielding a current or voltage shift proportional to the . In diodes, this process exhibits a of approximately 40 nC/ under typical conditions. For PN solar cells, the enables bias-free operation, where radiation directly induces a measurable without external voltage. The charge collection efficiency depends on carrier drift in the depletion field, ensuring proportional response to dose. These dosimeters offer real-time readout via integrated electrometers, contrasting with TLDs by eliminating the need for annealing or delayed processing, though they require periodic recalibration for long-term use. They demonstrate over the 50-200 keV range relevant to diagnostic and therapeutic X-rays, with response variations typically below 5% across this spectrum. As alternatives to TLDs, they provide superior but are limited by radiation hardness, with sensitivity degradation rates of 0.1-3.4% per kGy depending on beam and diode type. Applications include personal dosimeters for occupational exposure monitoring and in vivo radiotherapy verification, such as measuring or entrance doses during treatments. Their advantages encompass compact form factors (often <1 cm³), robustness without gas handling or high-voltage supplies, and ease of integration into arrays for multi-point measurements. P-type silicon diodes, in particular, exhibit enhanced radiation tolerance, maintaining stability up to several hundred in clinical beams.

Radiochromic Film Dosimeters

Radiochromic film dosimeters are self-developing polymer-based films that undergo a permanent color change upon exposure to , such as X-rays, enabling high-resolution mapping of two-dimensional distributions. These films provide a tissue-equivalent response due to their composition approximating attenuation properties and do not require processing, distinguishing them as versatile tools for . Introduced in the 1980s, they have become essential for verifying complex dose patterns in clinical settings. The active layer of radiochromic films is typically a thin (around 17–50 μm) polymer matrix, such as or poly(vinyl butyral), embedded with radiation-sensitive monomers or dyes including diacetylenes for polydiacetylene-based formulations or leuco dyes like microcrystalline leucomalachite green. Commercial variants, such as Gafchromic films from Ashland Advanced Materials, incorporate these components without silver halides, avoiding the light sensitivity and development needs of traditional radiographic films. This composition ensures stability under normal storage conditions and compatibility with various radiation types. The dosimetric response arises from radiation-induced polymerization of diacetylene monomers or activation (oxidation) of leuco dyes, forming conjugated polymer chains or colored chromophores that exhibit strong visible light absorption, often with a peak around 550 nm for Gafchromic EBT models. This color intensification correlates directly with absorbed dose, quantified via optical density (OD), which obeys Beer's law: \text{OD} = \epsilon \, d \, c where \epsilon is the molar absorptivity of the chromophore, d is the optical path length through the film, and c is the chromophore concentration. X-ray energy deposition, mainly through Compton scattering above 100 keV, triggers these irreversible chemical reactions, with full color stabilization occurring over 24–72 hours post-exposure. Performance characteristics include sub-millimeter , often better than 0.1 mm, allowing precise capture of dose gradients in small fields. Dose sensitivity spans 0.1–1000 across models like EBT3 (optimized for 0.1–10 ) and HD-V2 (up to 500 ), with energy independence for beams above 100 keV ensuring consistent response in megavoltage and kilovoltage regimes. Readout is achieved via flatbed scanners, providing digital dose maps with uncertainties typically under 2–3% after . In applications, radiochromic films excel in intensity-modulated radiotherapy (IMRT) plan verification, where they confirm multi-leaf collimator-defined dose patterns, and in for assessing off-axis and penumbra doses in intracranial treatments. Their near-tissue equivalence (effective ~7.4) and immediate post-scan analysis facilitate rapid without real-time constraints. Limitations include gradual post-irradiation , up to 5–10% in the first day for some models, requiring delayed scanning for accuracy, and to relative above 60%, which can alter response by 5–15% if films are exposed during or storage. against known uniform fields is essential to mitigate batch-to-batch variations and scanner artifacts, ensuring to primary standards.

Advanced and Emerging Detectors

Photon-Counting Detectors

Photon-counting detectors represent an advanced class of X-ray imaging systems that directly tally individual and discriminate their energies, enabling imaging capabilities beyond traditional energy-integrating detectors. These devices typically employ hybrid pixel architectures, combining a sensor layer—such as (CdTe) or (CZT)—bump-bonded to a (CMOS) readout chip. Exemplary implementations include the Medipix and Timepix families developed by collaborations, where the sensor converts X-ray into electron-hole pairs, and the CMOS chip processes the resulting charge signals. Each pixel features an adjustable energy threshold, allowing for per-pixel discrimination and binning of photon energies into multiple channels, typically 2–8 bins, to facilitate material decomposition in imaging. In operation, an incident absorbed in the generates a charge proportional to its ; if the amplitude exceeds the pixel's (often set around 20–25 keV for CdTe to reject low-energy noise), it is counted as a , with the binned accordingly. This single-photon avoids of noise, as only photon-induced signals contribute to the output, yielding noise-free images even at low doses. resolution typically achieves 5–10% (FWHM) at 60 keV, limited by factors like charge sharing and escape in the . Unlike direct conversion detectors that merely integrate charge, photon-counting systems provide timestamped events, enabling high and rejection of pile-up through advanced readout schemes. Performance metrics highlight their suitability for high-resolution spectral imaging, with pixel pitches ranging from 50–110 μm in research prototypes like Medipix3 (55 μm) to 225–500 μm in clinical designs, supporting count rates exceeding 10^6 photons/s/mm² and up to 3.5 × 10^8 in optimized systems. This enables applications such as multi-material decomposition for distinguishing tissues like , , and contrast agents, alongside dose reductions of 30–50% compared to energy-integrating detectors due to improved signal-to-noise efficiency and spectral optimization. Advantages include enhanced without electronic noise artifacts and elimination of signal pile-up in low-flux regimes, though challenges persist at high fluxes from charge-sharing effects and limited count rates in dense photon fields. Developments accelerated in the with FDA clearance of early systems like the Sectra MicroDose for in 2011, followed by full-body integration; the Siemens NAEOTOM Alpha received FDA approval in 2021 as the first commercial photon-counting scanner for clinical use, demonstrating spectral capabilities in routine exams. By 2025, adoption has expanded from to emerging applications, driven by prototypes achieving high- handling and cost reductions, though high expenses and flux limitations remain barriers to widespread deployment.

Novel Material-Based Detectors

Novel material-based X-ray detectors leverage innovative compounds such as metal halide perovskites, exemplified by methylammonium lead iodide (MAPbI3), which enable high photoconductive gain due to efficient multiplication under electric fields. These materials facilitate direct conversion with a low energy required for electron-hole pair creation, approximately 5 eV per pair, surpassing traditional semiconductors in sensitivity for low-dose applications. , such as 6,13-bis(triisopropylsilylethynyl)pentacene (TIPS-pentacene) blended with , offer inherent flexibility, allowing fabrication on bendable substrates while maintaining high X-ray sensitivity up to 1.3 × 10^4 μC/Gy·cm². Recent developments in the 2020s include single-crystal detectors, such as those based on MAPbI3 or cesium lead (CsPbBr3), achieving sensitivities around 10^5 μC/·cm² in prototypes suitable for high-resolution . -enhanced scintillators, like CsPbBr3/reduced oxide nanocomposites, demonstrate faster response times and improved optical yield, enabling sub-millisecond decay for dynamic detection. These prototypes highlight the potential for scalable solution-processing, with uniform films produced via rheological engineering to minimize defects. Operationally, these detectors function in avalanche or photoconductive modes, where applied biases amplify signals through carrier multiplication, yielding ultralow detection limits below 1 nGy/s. Stability enhancements, such as encapsulation with cross-linked polymers or operation in photovoltaic mode, mitigate degradation from humidity and ion migration, retaining over 90% performance after prolonged exposure. Integration onto flexible substrates supports wearable formats, preserving sensitivity under bending radii as low as 5 mm. Applications span low-dose portable imaging, where perovskite devices enable high-quality radiographs at doses under 300 nGy, and space radiation monitoring, leveraging their radiation hardness for detection. Post-2020 advances include EU-funded initiatives like the PEROXIS and PERFORM projects, which have accelerated prototyping of photon-counting arrays for medical diagnostics. By 2025, commercialization efforts target hybrid -organic panels for cost-effective, flexible systems in clinical and settings, with ongoing work on lead-free alternatives such as bismuth-based perovskites achieving sensitivities around 10^4 μC/·cm² to address concerns. Key challenges involve lead toxicity in perovskites, prompting shifts to lead-free alternatives like bismuth-based variants, though these often trade off . Long-term remains limited by environmental factors, with unencapsulated devices degrading after hours of operation. Compared to (CdTe), perovskites offer superior flexibility and lower production costs but lower (), requiring thicker layers for comparable absorption, while CdTe's brittleness restricts conformable designs.

References

  1. [1]
    X-ray Detectors - Electrical Current Detections - Imagine the Universe!
    The most common type of X-ray detector uses an electric current to measure incoming X-rays. In this type of detector, an X-ray interacts with a material ...Missing: definition principles
  2. [2]
    X-ray Imaging - Medical Imaging Systems - NCBI Bookshelf - NIH
    In this chapter, the physical principles of X-rays are introduced. We start with a general definition of X-rays compared to other well known rays, e. g., ...
  3. [3]
    Medical X-ray Imaging - FDA
    Feb 21, 2023 · Computed tomography (CT), fluoroscopy, and radiography ("conventional X-ray" including mammography) all use ionizing radiation to generate images of the body.Radiography · Pediatric X-ray Imaging · Medical Imaging
  4. [4]
    [PDF] Neutron and X-ray Detectors - DOE Office of Science
    NEUTRON AND X-RAY DETECTOR WORKSHOP 23. In this section, general properties and typical applications of X-ray detectors are described. The state of the art ...
  5. [5]
    [PDF] Photon cross sections, attenuation coefficients, and energy ...
    Key words: Attenuation coefficient; Compton scattering; cross section ... gamma rays; pair production; photoelectric absorption; photons; x-rays. 1 ...
  6. [6]
    [PDF] 03 - Interaction of Photons with Matter.
    May 7, 2011 · Compton scattering occurs at all energies in all materials, but is the dominant interaction in the medium energy range. • Pair production only ...
  7. [7]
    Compton Effect Example - Physics
    Δλ = λ/ - λ = (h/mec) (1 - cosθ). The combination of factors h/mec = 2.43 x 10-12 m, where me is the mass of the electron, is known as the Compton wavelength.Missing: m_e | Show results with:m_e
  8. [8]
    None
    ### Summary of Photon Interactions with Matter
  9. [9]
    Interaction of X-Rays with Matter | MicroCT Facility (MCT)
    This dependence is variable depending on the X-ray energy. The attenuated X-ray intensity (I) through a material can be measured by the Beer-Lambert Law: I = I0 ...
  10. [10]
    Flat-panel detectors: how much better are they? - PubMed Central
    X-ray signal detection: indirect and direct conversion. Although the TFT array and associated electronics are common to FPD systems, significant differences in ...Image Intensifier Technology · Flat-Panel Detectors For... · Image Intensifier/tv Versus...
  11. [11]
    Digital radiography - Book chapter - IOPscience
    ... photostimulable phosphor material, known as photostimulable storage phosphor (PSP). When x-rays are received, a latent image is formed which is subsequently ...
  12. [12]
    Measurement of mobility and lifetime of electrons and holes in ... - NIH
    The measured values for μτe/h (mobility-lifetime product) are in agreement with earlier published data. Keywords: Solid state detectors; Charge transport and ...
  13. [13]
    Silicon photomultiplier‐based scintillation detectors for photon ... - NIH
    Jun 25, 2021 · Here, E is the deposited X‐ray energy in the scintillator [keV] and Y is the light yield of the scintillator [photons/keV]. Y is often a ...
  14. [14]
    A comparative analysis of OTF, NPS, and DQE in energy integrating ...
    Conclusions: This article develops analytical models of OTF, NPS, and DQE for energy integrating and photon counting digital x-ray detectors. While many ...
  15. [15]
    Direct-conversion flat-panel imager with avalanche gain
    Avalanche multiplication is a physical process in which impact ionization of charge produces charge amplification. It has been used in radiographic gas imagers ...
  16. [16]
    Recent Development in X-Ray Imaging Technology - PubMed Central
    For instance, α-Se-based direct conversion flat-panel X-ray detectors exhibit better imaging spatial resolution than that of indirect conversion flat-panel ...
  17. [17]
  18. [18]
    [PDF] REVIEW - AAPM
    While thicker screens capture more X-rays, they also create more light scatter and blur the image. Therefore, it is impossible to offer a screen-film system ...
  19. [19]
    Spatial resolution | Radiology Reference Article - Radiopaedia.org
    Feb 20, 2024 · Spatial resolution is expressed in line pairs per mm (lp mm). The ... film-screen radiography: 8 to 12 line pairs per mm 1. digital ...
  20. [20]
    Digital X-Rays: Past, Present, and Future - Maven Imaging
    Nov 8, 2024 · Today, digital x-ray technology is widely used in medical imaging and has largely replaced traditional film x-rays. Digital x-ray systems ...Missing: timeline | Show results with:timeline
  21. [21]
    Computed radiography utilizing scanning laser stimulated ...
    Sep 1 1983. Computed radiography utilizing scanning laser stimulated luminescence. Authors: M Sonoda, M Takano, J Miyahara, and H KatoAuthors Info & ...Missing: 1983 | Show results with:1983
  22. [22]
    Storage Phosphors for Medical Imaging - PMC - PubMed Central - NIH
    BaFCl:Eu2+ and BaFBr:Eu2+ had been of the first rare-earth X-ray phosphors for screen/film radiography [5]. The BaFX matrix is a layered material with a ...
  23. [23]
    [PDF] aapm report no. 93
    The photostimulable phosphor (PSP) stores absorbed x-ray energy in crystal structure “traps,” and is sometimes referred to as a “storage” phosphor. This ...
  24. [24]
    [PDF] Fluoroscopic Technology From 1895 To 2019 Drivers
    The X-ray image intensifier (II) introduced into practice in the 1950s was a ... Mason Sones performing diagnostic cardiac angiography using a 1956 11-inch image ...Missing: peak | Show results with:peak
  25. [25]
    X-Ray Image Intensifier - an overview | ScienceDirect Topics
    In the U.S., the video frame rate is 30 frames per second. Thus, X-ray pulse rates of 30, 15, 7.5 frames per secondare feasible. The temporal relationships ...
  26. [26]
    [PDF] The “Legacy” Fluoroscopy - AAPM
    Flux Gain: Electrons are accelerated (~30kV). Typical flux gain values; 50~60. Total Brightness Gain of Image. Intensified Fluorocopy System. For a 9” mode ...
  27. [27]
    Artifacts/Miscellaneous | Radiology | SUNY Upstate
    The first artifact, pincushion distortion, is caused by the spherical input phosphor structure and photocathode electron image mapped onto the planar output ...
  28. [28]
    Digital Radiography (Direct Vs Indirect Flat Panels)
    The major advantage of direct flat panel detectors is that the electrons do not spread nearly as much as visible photons. This results in a higher resolution ...Missing: screen- | Show results with:screen-
  29. [29]
    High spatial resolution direct conversion amorphous selenium X-ray ...
    Amorphous selenium (a-Se) as a wide-bandgap thermally evaporated photoconductor exhibits ultra-low thermal generation rates for dark carriers and has been ...
  30. [30]
    TFT readout array: a magnified view of one pixel. - ResearchGate
    Mercuric iodide (HgI<sub>2</sub>) polycrystalline films have great potential as direct-conversion X-ray detectors for digital X-ray imaging.
  31. [31]
    Photon counting performance of amorphous selenium and its ... - NIH
    Photon counting detectors (PCD) have the potential to improve x-ray imaging; however, they are still hindered by high costs and performance limitations.
  32. [32]
    Enhanced Detection Efficiency of Direct Conversion X-ray Detector ...
    Nov 28, 2013 · Amorphous selenium has an acceptable X-ray absorption coefficient for low X-ray energy, good charge transport properties and low dark current.
  33. [33]
    Digital radiology using active matrix readout of amorphous selenium ...
    Jun 4, 1998 · A flat-panel x-ray imaging detector using a layer of amorphous selenium (-Se) for direct conversion of x rays (to charge) and an active ...
  34. [34]
    (PDF) Direct conversion X-ray sensors: sensitivity, DQE and MTF
    Aug 7, 2025 · direct flat-panel detectors depends strongly on the fill-factor. Hence, for accurate modelling it is important to determine. the fill-factor ...
  35. [35]
    FDA approves Selenia system for full-field digital mammography
    Its Lorad Selenia is the first digital mammography system based on an amorphous selenium flat-panel detector, which directly converts x-rays into electrical ...Missing: dental | Show results with:dental
  36. [36]
    Flexible x-ray imaging detector based on direct conversion in ...
    Jun 10, 2014 · In this paper, the authors propose a mechanically flexible direct conversion x-ray detector as a potential solution for portable and ...
  37. [37]
    [PDF] Design and Performance Characteristics of Digital Radiographic ...
    35 x 43 cm -- 2.5 lp/mm (200 µm). 24 x 30 cm -- 3.3 lp/mm (150 µm). 18 x 24 cm -- 5.0 lp/mm (100 µm). Screen/film resolution: 7-10 lp/mm (80 µm - 25 µm).
  38. [38]
    Impact of charge carrier trapping on amorphous selenium direct ...
    Oct 4, 2017 · The incomplete charge collection due to carrier trapping degrades the signal strength.Missing: challenges polarization
  39. [39]
    Investigation of the signal behavior at diagnostic energies of ... - NIH
    However, high levels of charge trapping, lag, and polarization, as well as pixel-to-pixel variations in x-ray sensitivity are of concern. While the results of ...
  40. [40]
    (PDF) Digital radiography with large-area flat-panel detectors
    Aug 6, 2025 · Detectors with indirect conversion are built with unstructured or structured scintillators, the latter resulting in less lateral diffusion of ...
  41. [41]
  42. [42]
    [PDF] Performance of Digital Radiographic Detectors: Quantification and ...
    Methods for measuring the MTF, the NPS, and the DQE are then described. The chapter ends with an outline of detector performance factors that may be considered ...
  43. [43]
    MTF compensation for digital radiography system with indirect ...
    Aug 9, 2025 · We studied the noise characteristics of two flat panel detectors with structured columnar scintillator (CsI) and granular scintillator (Gd2O2S).
  44. [44]
    None
    ### Summary of Gas Detectors
  45. [45]
    Introduction to Radiation Detectors - Mirion Technologies
    The different types of gas-filled detectors are: ionization chambers, proportional counters, and Geiger-Mueller (G-M) tubes. The major differentiating ...
  46. [46]
    Nuclear Medicine Instrumentation - StatPearls - NCBI Bookshelf
    Nov 14, 2023 · Gas detectors are categorized into 3 main types: ionization chambers, proportional counters, and GM counters. Ionization chambers are commonly ...
  47. [47]
    The reflection of X-rays by crystals - Journals
    The revolving table in the centre carries the crystal. The ionisation chamber is tubular, 15 cm. long and 5 cm. in diameter. It can be rotated about the axis of ...
  48. [48]
    [PDF] Physical aspects of irradiation - NIST Technical Series Publications
    1 W is often expressed in ev, where 1 ev=1.602X10->2 erg. W is fully discussed in section IAS. For air, W has a value of 33.7 ev per ion pair for electrons ...
  49. [49]
    [PDF] Calibration of Radiation Detectors in Terms of Air-Kerma Using ...
    In addition, the temperature and the pressure of the air surrounding the detector must be measured for the case of ionization chambers that are open to the ...
  50. [50]
    Gamma and X-Ray Detection - Mirion Technologies
    Typical resolutions are about 16 to 20% full-width at half maximum (FWHM). Operating voltages depend upon the fill gas as well as the geometry. For X rays, ...
  51. [51]
    [PDF] GASEOUS DETECTORS FOR SCIENCE (AND DISCUSSION)
    Oct 5, 2012 · Energy Resolution for X-rays. No. of electrons created is given by: N = E ⁄ w where w is the average energy to create an electron/ion pair ...
  52. [52]
    [PDF] AAPM Report No. 87
    The key structure in the silicon diodes used for in vivo dosimetry is the pn junction. N-type silicon is doped with impurities of a pentavalent element (e.g.,.
  53. [53]
    Steady-State Response of Silicon Radiation Detectors of the ... - NIH
    Since then, several authors investigated the possibility of using silicon solar cells for x- and gamma-ray dosimetry [2,3,4]. Silicon solar cells are ...
  54. [54]
    P-channel MOSFET as ionizing radiation detector - ScienceDirect.com
    The principle of operation of MOSFETs as dosimeters is based on their measurement of the total absorbed radiation dose through a shift in the threshold voltage ...
  55. [55]
    [PDF] tectors of the Diffused pn Junction Type to X Rays. I: Photovoltaic ...
    Since then, several authors investigated the possibility of using silicon solar cells for x- and gamma-ray dosimetry [2,3,4]. Silicon solar cells are ...
  56. [56]
    Energy and field size dependence of a silicon diode designed for ...
    Mar 8, 2017 · Purpose. To investigate the energy dependence/spectral sensitivity of silicon diodes designed for small-field dosimetry and obtain response ...
  57. [57]
    Performance Characterization of Dosimeters Based on Radiation ...
    Jan 30, 2022 · The irradiation effect on the dose responses of the diodes was investigated within ranges of 0–50 kGy and 0–275 kGy, following similar ...Missing: MOSFET | Show results with:MOSFET<|control11|><|separator|>
  58. [58]
    Performance characteristics of mobile MOSFET dosimeter for ...
    The high sensitivity MOSFET dosimeters were calibrated to measure absorbed dose during image guidance in radiotherapy, mainly with kV CBCT. The performance ...
  59. [59]
  60. [60]
    Dosimetric evaluation of unlaminated radiochromic films exposed to ...
    Oct 27, 2025 · INTRODUCTION. Radiochromic GafChromic films, introduced in the 1980s, provide accurate, high‐resolution radiation dose measurements. These films ...
  61. [61]
    [PDF] Radiochromic film for medical radiation dosimetry
    However, most radiochromic film dosimeters utilise materials which turn a blue colour when exposed to radiation. The image formation in radiochromic products.
  62. [62]
    [PDF] United States Patent - OSTI.GOV
    “A radiochromic film based on leucomalachite green for high-dose dosimetry applications” Rad Meas. 62 (2014). Soliman, et al. “Leuco crystal violet poly(vinyl ...<|separator|>
  63. [63]
    Radiochromic Film Dosimetry and its Applications in Radiotherapy
    The feature of radiochromic film that makes it useful as a dosimeter is that the amount of color change or darkening is proportional to the absorbed dose. As ...
  64. [64]
    Reference radiochromic film dosimetry: Review of technical aspects
    Aug 9, 2025 · Radiation dose deposited within a sensitive layer of the radiochromic film initiates polymerization of the active component, the degree of which ...
  65. [65]
    Energy and dose dependence of GafChromic EBT3-V3 film across a ...
    Dec 9, 2019 · Within the range 6 MV-2 mm Al HVL and dose higher than 500cGy, GafChromic EBT3-V3 films are energy independent. In this dose range, films can be ...
  66. [66]
    Radiochromic Film | Role and Applications in Radiation Dosimetry
    Oct 30, 2017 · It focuses on practical uses of radiochromic film in radiation dosimetry for diagnostic x-rays, brachytherapy, radiosurgery, external beam ...<|control11|><|separator|>
  67. [67]
    Precise film dosimetry for stereotactic radiosurgery and stereotactic ...
    Oct 4, 2016 · The purpose of this study is to evaluate the dosimetric uncertainty associated with Gafchromic™ (EBT3) films and establish a practical and ...
  68. [68]
    [PDF] Radiochromic Film - AAPM
    ... (ISP). Since 1980s: – Since 1980s: Radiochromic films based on polydiacetylene have been introduced for medical applications pp. Page 6. Radiochromic Films.Missing: history | Show results with:history
  69. [69]
    Investigation of EBT3 radiochromic film's response to humidity
    Therefore, it is recommended to build a new calibration curve for radiochromic films for a specific situation involving dose measurements in liquid water.Missing: fading | Show results with:fading
  70. [70]
    Temperature, humidity and time. Combined effects on radiochromic ...
    ... radiochromic film dosimeters have been investigated in the relative humidity (RH) range 11-94% and temperature range 20-60°C for irradiation by 60 Co ...
  71. [71]
    [PDF] Radiochromic film dosimetry system: from calibration to in vivo ...
    Jan 26, 2011 · Humidity in the room should be kept constant. The sensitivity to light is also ignored as films are kept in opaque envelope except when ...
  72. [72]
    Medipix3 | medipix.web.cern.ch
    Medipix3 is a CMOS pixel detector readout chip designed to be connected to a segmented semiconductor sensor.
  73. [73]
    Photon-counting x-ray detectors for CT - IOPscience
    One of the main advantages of the PCD is the improved spatial resolution compared to conventional detectors. PCDs that have been designed for full-body clinical ...
  74. [74]
    Medipix4, a large 4-side buttable pixel readout chip with high ...
    The Medipix4 chip is the latest member in the Medipix/Timepix family of hybrid pixel detector chips aimed at high-rate spectroscopic X-ray imaging using high-Z ...
  75. [75]
    Spectroscopic X-ray imaging with photon counting pixel detectors
    Single particle counting hybrid pixel detectors simultaneously provide low noise, high granularity and high readout speed and make it possible to build ...
  76. [76]
    Investigations on the Performance of a 5 mm CdTe Timepix3 ... - NIH
    Dec 13, 2024 · Timepix3 [7] is a hybrid pixel–photon-counting detector (Figure 1) developed by the Medipix collaboration (Geneva, Switzerland). It features ...
  77. [77]
    First Photon-counting CT System Cleared by the FDA | DAIC
    Oct 4, 2021 · The Siemens Naeotom Alpha is the first commercialized photon-counting CT scanner. It gained FDA clearance Sept. 30, 2021.
  78. [78]
    Siemens Healthineers Receives FDA Clearance for Naeotom Alpha ...
    Mar 10, 2025 · Siemens Healthineers has received Food and Drug Administration clearance for its Naeotom Alpha class of photon-counting computed tomography scanners.
  79. [79]
    Halide lead perovskites for ionizing radiation detection - Nature
    Mar 6, 2019 · Advance in the development of perovskite X-ray imaging detectors. The first perovskite radiation detector made by MAPbI3 single crystal showed ...
  80. [80]
    Highly sensitive and flexible organic X-ray detectors - ICMAB
    May 1, 2020 · The fabricated detectors reach a record sensitivity (1.3·104 μC/Gy·cm2), the highest value reported for organic-based direct X-ray detectors, ...
  81. [81]
    Stable perovskite single-crystal X-ray imaging detectors ... - Nature
    May 8, 2023 · Here we show that both single-photon-counting and long-term stable performance of perovskite X-ray detectors are attained in the photovoltaic mode of operation ...
  82. [82]
    Enhanced X-ray photon response in solution-synthesized CsPbBr 3 ...
    Dec 1, 2018 · Thus, a significant enhancement in optical and X-ray photon response is demonstrated in CsPbBr3/rGO nanocomposites as compared with pure CsPbBr3 ...
  83. [83]
    Rheological engineering of perovskite suspension toward high ...
    Oct 27, 2023 · We report a synergistic strategy of rheological engineering the perovskite suspensions to achieve X-ray flat panel detectors with pixel-level high uniformity.
  84. [84]
    Halide perovskites: A dark horse for direct X‐ray imaging
    Nov 17, 2020 · In practice, a photoconductive gain is widely observed in perovskite X-ray detectors when shallow defects are present in the perovskite ...
  85. [85]
    Flexible perovskite scintillators and detectors for X-ray detection
    Dec 22, 2022 · Perovskite films could be fabricated on a flexible substrate for special X-ray detections. With no grain boundaries existed, perovskite single ...
  86. [86]
    Revolutionizing X-ray Imaging: A Leap toward Ultra-Low-Dose ...
    Nov 13, 2024 · We further demonstrate the promise of our perovskite X-ray detectors for low-bias portable applications with high-quality X-ray imaging and ...
  87. [87]
    Perovskite Photon Counting X-ray Detectors for Medical Imaging ...
    Apr 29, 2025 · The PERFORM project aims to develop perovskite photon-counting detectors for X-ray imaging, using advanced PCDs to reduce radiation and improve ...
  88. [88]
    Home | The EU-funded PEROXIS project
    The EU-funded PEROXIS project aims to develop a groundbreaking highly sensitive and high spatial resolution direct X-ray detection technology.
  89. [89]
    Lead‐Free Halide Perovskites for Direct X‐Ray Detectors - PMC - NIH
    However, the instability and the toxicity of lead‐based perovskites have greatly hindered its practical applications. Alternatively, lead‐free perovskites with ...
  90. [90]
    Perovskite materials in X-ray detection and imaging - RSC Publishing
    Feb 22, 2024 · This review will serve as a roadmap for directing the difficulties associated with perovskite materials in X-ray detection and imaging.
  91. [91]
    Perspective of perovskite-based X-ray hybrid pixel array detectors
    Jul 29, 2024 · This paper presents potential perovskite-based detector materials and compares their performance with the state-of-the-art CdTe-based detectors.