1 Introduction

When an energetic particle passes through matter, it induces ionization or excitation. If the particle is charged, its electromagnetic field interacts directly with the orbital electrons of the atoms in the matter. Photon can undergo the Photoelectric effect, the Compton Scattering, or Pair Production, transferring part or full of its energy to the orbital electrons of the atom in the matter, resulting in ionization or excitation. In the case of uncharged neutral particles, such as neutrons, charged particles are produced by nuclear reactions, which then cause ionization or excitation. Elements that utilize the ionizing effect, luminescence phenomenon, and physical or chemical changes caused by nuclear radiation in gases, liquids, or solids for nuclear radiation detection are known as nuclear detectors. The nuclear detectors use an appropriate detection medium as the matter that interacts with the particles, and the ionization or excitation generated by the particles in the detection medium is transformed into various forms of signals. Nuclear electronics reads and processes the signal from the nuclear detectors and extracts the information. This information can then be used to directly or indirectly determine parameters of nuclear radiation, such as the type, energy, intensity, arrival time, or lifetime.

More than 100 types of radiation detectors can give electrical signals. Gas detectors were introduced as early as 1908 [1]. However, it was not until 1931 that the problem of fast counting was solved after the advent of pulse counters [2]. In 1947, scintillation counters appeared, significantly improving the efficiency of detecting particles due to their much greater density than gases [3]. The most notable is the NaI (Tl) scintillator, which has a high energy resolution for \(\gamma\) rays [4]. In the early 60 s, the successful development of semiconductor detectors led to a new development of energy spectrum measurement technology. Modern nuclear detectors used in high-energy physics, nuclear physics, and other scientific and technological fields have been rapidly developed over the last few decades. This paper will discuss the principle, short history, development, and recent advances on some representing semiconductor detectors, gaseous detectors, scintillation detectors, Cherenkov detectors, transition radiation detectors, and readout techniques.

2 Semiconductor detector

When the radiation ray or particles incident into the sensitive region of the semiconductor detector, the radiation ray or particles continuously loses energy and generates electron–hole pairs, and is separated from the two ends of the electrode drift under the action of the external electric field to collect, thus forming an output pulse signal. According to the structure classification, the structure of semiconductor detector has crystal conductivity type, p-n junction type, p-i-n junction type, Schottky junction type and so on.

The number of electron holes created is related to the energy of absorbed particles or rays and the energy of electron-hole pair formation, shown in the following formula:

$$\begin{aligned} N=\frac{E}{\varepsilon _{i}} \end{aligned}$$
(1)

where E is the absorbed energy, \(\varepsilon _{i}\) is the electron–hole creation energy. For silicon, the electron hole pair creation energy is 3.6 eV, while silicon carbide is 8.9 eV. This section will discuss the silicon detectors and the wide bandgap semiconductor detectors.

2.1 Silicon detector

Silicon detectors are very well suited to detecting and measuring ionizing radiation and light caused by interacting with charged particles and photons (soft X-rays). A great feature of silicon detectors is that they can operate with low noise at room temperature due to the band gap of 1.12 eV. In addition, silicon has a high density of 2.329 g/cm\(^3\), a small average ionization energy of 3.62 eV, good mechanical stability, and is easy to produce on a large scale. These advantages make silicon the perfect option for many physics experiments and equipment.

1960 Bromley showed a proposal for a 3D array of silicon diodes that would function as a solid-state “cloud chamber” [5]. The Charge-Coupled Device (CCD), introduced in 1969, with good spatial resolution and low noise in photon detection, received the Nobel Prize in 2009. Since the first half of the 1980s, Silicon Strip Detectors (SSD) have become dominant in particle tracking [6]. In the 1980s and 1990s, an evolution started from strip to pixel detectors, mainly consisting of monolithic and hybrid types. The Low Gain Avalanche Diode (LGAD) recently provided fast detection of Minimum Ionizing Particles (MIPs) with a timing resolution of several picoseconds. Recent decades have witnessed substantial growth in silicon detectors, expanding from a few thousand channels to billion channels with dramatically increased granularity facilitated by rapid advancements in semiconductor technology [7].

2.1.1 Charge-coupled device

CCD comprises a series of coupled Metal-Oxide-Semiconductor (MOS) capacitors as its fundamental structure. Electrons are produced when photons hit MOS capacitors because of the photoelectric effect, and will be collected as signal charge packets by potential wells of each MOS capacitor, which are formed by applying positive voltages to gate electrodes. Subsequently, the packets stored in each MOS capacitor are transferred from one pixel to its neighbor through sequential adjusting gate voltages of each MOS capacitor. The last capacitor dumps its charge into a charge amplifier, which converts the charge into a voltage, and then readout. MOS capacitors share readout circuit, so the transferring process will be repeated continuously until all the signal charge packets are read out.

CCD-based detectors can be implemented in several different architectures. The most common are full-frame, frame-transfer, and interline. In a full-frame device, all of the imaging area is active. While in frame-transfer CCDs, half of the silicon area is covered by an opaque mask, typically aluminum. It is necessary to quickly transfer input signals from the image area to the opaque or storage area to minimize smear in this design. Unlike frame-transfer, the interline architecture involves only a pixel shift from the imaging area to the storage area, enabling shutter times of less than \({1}\,{\upmu \text{s}}\) and virtually eliminating smear.

The advantages of CCD are apparent. It has excellent spatial resolution because its pixel size can be very small. Pixels as small as \({1.56}\,{\upmu \text{m}} \times {1.56}\,{\upmu \text{m}}\) have appeared as early as 2006. Also, the CCD has low noise because its readout process does not produce noise, making it well-suited for small signal detection. In addition, shared readout provides good pixel-to-pixel consistency. What’s more, it’s worth noting that modern CCD technology allows for manufacturing large wafer-level devices, simplifying the mechanical design of detectors, which is crucial for experiments requiring a large sensitive area [8].

The primary utilization of CCD is in astronomical observation. For instance, the Skipper-CCD, designed for the Oscura experiment [9] and employed in low-energy neutrino studies and dark matter searches, is crafted from high-resistance (HR) n-type silicon wafers. The sensor has a thickness of 675 \(\upmu\)m, an effective area of 1.9 cm \(\times\) 1.6 cm, and charge can be read through amplifiers at the four corners. Meeting the Oscura’s stringent requirements, including a readout time of less than 2 h and a radioactive background rate of 0.01 event/kg/day/keV, the sensor exhibits a dark current of approximately 10\(^{-6}\) e\(^-\)/pix/day. Test results conducted at Fermilab (as depicted in Fig. 1) reveal that 71% of the sensors exhibit sub-electron noise, with a surface dark current of about 0.031 e\(^-\)/pix/day [9].

Fig. 1
figure 1

(Color online) Left: Skipper-CCD in Cu module. Center and right: Setups at FNAL for testing packaged sensors. Reproduced from [9]

The “Mozi” Wide Field Survey Telescope (WFST) is a new generation survey telescope that is being built in China. It is equipped with a mosaic CCD camera with 0.73 gigapixels on the primary focal plane for high quality image capture over a 6.5-square-degree field of view [10]. Comprising nine E2V CCD290-99 chips, each with a resolution of 9 K \(\times\) 9 K and a pixel size of 10 \(\upmu\)m, the camera supports full-frame or split full-frame readout modes. The 3D model of the CCD detector system is illustrated in Fig. 2. The readout can occur through 8 or 16 output channels, with a typical readout noise of 4 e\(^-\)@0.5MHz. To meet the scientific objective of rapid sky surveying, the maximum readout speed reaches 3 MHz, and the readout noise is maintained below 15 e\(^-\)@1MHz [11]. Nowadays, CCD also finds applications in fields such as spectral analysis and heavy ion Computed Tomography (CT) [12].

Fig. 2
figure 2

(Color online) 3D model of the CCD detector system on “Mozi”. Reproduced from [11]

2.1.2 Silicon strip detector

The SSD is the first detector device using the lithographic capabilities of microelectronics, mainly used as a vertex and tracking detector. As shown in Fig. 3, the SSD is commonly fabricated through selective etching of microstrips on the surface of N-type silicon wafers. The highly doped P+ microstrip forms a PN junction on the surface and extends its depletion layer across the entire N-type substrate with applied bias voltage. After covering aluminum on the top, a microstrip that acts as a charge-collecting electrode is completed. Multiple strips form a single-sided SSD, which provides one-dimensional position detection. When particle hits, electron–hole pairs are generated in the depletion region because ionization caused by deposited energy. Holes, driven by an external electric field, are collected by the strips, converted into electrical signals, and processed by the readout circuit. The double-sided SSD featuring heavily doped P+ junction strips that collect the holes and the N+ ohmic strips that collect electrons can realize two-dimensional position detection. Since SSDs have low noise, good linearity, a high count rate, and excellent position and energy resolution, they are widely used in high-energy physics, nuclear physics, and space detection facilities [13, 14].

Fig. 3
figure 3

(Color online) Working principle diagram of AC coupled single-sided SSD. Reproduced from [15]

In high-energy physics, the radiation-hardened Silicon Microstrip Vertex Detector (SVX’) was incorporated into the CDF experiment during the Tevatron \({{\overline{p}}}p\) collider run 1B in 1993. As the first silicon vertex detector in a hadron collider environment, SVX’ modules (also referred to as barrels) consist of four layers of SSDs segmented into twelve \({30}^\circ\) wedges. Three wafers are bonded to form a 25.5 cm long “ladder” with strips aligned for \(\gamma \phi\) coordinate measurements. The AC-coupled readout is facilitated by the radiation-resistant SVXH3 chip [16], utilizing a \({1.2}\,{\upmu \text{m}}\) CMOS process with radiation tolerance exceeding 1 Mrad. The chip provides a gain of 21 mV/fC and an equivalent input noise of 1300 e\(^-\) at a strip input capacitance of approximately 30 pF [17], supporting data sparse and serial readout modes [18]. When a single strip is hit, the positional resolution can reach \({13}\,{\upmu \text{m}}\) [16]. The SSD has also been widely used in high-energy experiments, such as the Semiconductor Tracker (SCT) in ATLAS for two-dimensional track measurement [19], the Silicon Strip Tracker (SST) in CMS for readout of the \(\gamma \phi\) coordinate and measuring three-dimensional information for the separation of particle tracks [20],and the NA11 spectrometer at CERN SPS for flight paths reconstruction [21], and so on.

SSD plays a vital role in nuclear physics experiments. The Institute of Modern Physics (IMP) has constructed a gas-filled recoil separator called Spectrometer for Heavy Atom and Nuclear Structure (SHANS) to study the heavy and superheavy nuclei properties. A silicon semiconductor detector box (Si-box) is installed at the focal plane position of the separator, as illustrated in Fig. 4. The Si-box incorporates three Position Sensitive Silicon Detectors (PSSD) at the back as implantation detectors, with a thickness of \({300}\,{\upmu \text{m}}\) and a collective effective area of 150 mm \(\times\) 50 mm, offering 3 mm horizontal position resolution. The PSSD demonstrates a typical energy resolution of 50 keV FWHM for 5–10 MeV \(\alpha\) particles. Leveraging this PSSD detector, it can be used to measure \(\alpha\) particles of 1–20 MeV, implantation and fission fragments with energies of 5–200 MeV, and detect the esca** radioactive decay events [22]. Similar applications of SSD in nuclear physics can also be found in MUST, TIARA, and HIRFL-RIBLL. The MUST (“MUr \(\grave{a}\) Strips”) detector consists of 8 SSD–Si (Li) telescopes used to identify recoiling light charged particles through time of flight, energy loss and energy measurements and to determine precisely their scattering angle through X and Y position measurements [23]. The Transfer and Inelastic All-angle Reaction Array (TIARA) consists of 8 resistive charge division detectors forming an octagonal barrel around the target and a set of double-sided silicon-strip annular detectors positioned at each end of the barrel, designed to study direct reactions induced by radioactive beams in inverse kinematics [24]. The detection system consists of five \(\Delta E - E\) telescopes, consisting of one double sided silicon strip detector (DSSD) as the \(\Delta E\) detector, and a square silicon detector as the E detector, designed to measure the angular distributions of both elastic scattering and breakup simultaneously, on the Radioactive lon Beam Line in Lanzhou at Heavy lon Research Facility in Lanzhou(HIRFL-RIBLL) [25].

Fig. 4
figure 4

(Color online) Schematic view of the focal plane detection array. Reproduced from [26]

The SSD has been a “standard” component for space detection [27]. The Dark Matter Particle Explorer (DAMPE), launched in 2015, incorporates silicon-tungsten tracker-converter (STK) to collect the electron-positron pairs converted from the incoming photons by the tungsten converters. As shown in Fig. 5, the STK contains of 6 tracking planes each consisting of 2 layers of SSD arranged orthogonally. Each single layer consists of 16 ladders, and each ladder comprises 4 modules bonded end-to-end. The SSDs are glued on the flex part of the Tracker Front-end Hybrid (TFH) board to form a ladder. Each layer of SSD has an active area of 0.55 m\(^2\), with a total radiation length of 0.976% \(X_0\), spatial resolution \(<{80}\,{\upmu \text{m}}\) within \({60}^\circ\) incidence, and a total weight of 154.8 kg.

Fig. 5
figure 5

(Color online) Left: Exploded view of the STK; Right: The STK single ladder, made by four SSDs. Reproduced from [28]

The readout is performed on one every other strip (corresponding to 384 channels per ladder) in the SSD. The signal sha** and amplification is performed by six VA140 ASIC chips mounted on the TFH [28]. The VA140 is a 64-channel low-noise charge-sensitive amplifier ASIC from the Norwegian IDEAS company. Each channel of the VA140 comprises a charge-sensitive preamplifier, a filter sha** circuit, and a sample and hold circuit. The input charge dynamic range spans from \({-200}\,{\text{fC}}\) to 200 fC, with a sha** time of only \({6.5}\,{\upmu \text{s}}\). This ASIC employs serial readout through a multiplexing circuit and the power consumption is 0.29 mW/channel.

Based on the successful development and operation of DAMPE, the Purple Mountain Observatory of the CAS, together with several institutions, proposed the Very Large Area Gamma-ray Space Telescope (VLAST).As shown in Fig. 6, the VLAST plans to use SSD to build Silicon Tracker and low Energy gamma-ray Detector (STED). The main function is to realize the conversion of high-energy gamma photon-electron pairs, and to achieve high-angular resolution observation of gamma photons through the track measurement of electron pairs. The main technical indicators of STED are [29]:

  • Number of detector layers: 8 super layers; (each super layer includes a CsI detection layer and two large silicon micro-strip detection layers);

  • Active detection area of each layer: \(\ge\) 2.8 m \(\times\) 2.8 m;

  • Detection energy range: 1 MeV –100 MeV;

  • Spatial resolution \(< {0.1}^\circ\) (@50 GeV).

Fig. 6
figure 6

(Color online) The schematic plot of the payload of VLAST. Reproduced from [29]

The VA140 from IDEAS company is also a potential candidate for the readout of the SSD on VLAST. In addition, the IMP is now develo** a readout chip called SiReadout. The first version of the chip is shown in Fig. 7. The preliminary single-channel test indicates that it has a gain of 2.01 mV/fC, power consumption of approximately \({220}\,{\upmu \text{W}}\), a peaking time of \({3.12}\,{\upmu \text{s}}\), linearity error below 1%, and an Equivalent Noise Charge (ENC) of 992 e\(^-\) at zero F plus 13.5 e\(^-\) per pF.

Fig. 7
figure 7

(Color online) Microphoto of the first version of SiReadout

Also, the ALPHA Magnetic Spectro-meter (AMS) experiment similarly utilizes SSD to build a magnetic spectrometer to measure momenta, charges and mass of the particles [30], the Fermi Large Area Telescope (Fermi-LAT) builds the LAT tracker to reconstruct the incoming photon direction, etc [31].

2.1.3 Silicon pixel detector

Silicon pixel detector is currently one of the main detectors and is expected to play a significant role in the future due to their exceptional spatial resolution, reduced material budget, low noise, and rapid readout rates. It mainly consists of a densely packed pixel array and peripheral readout circuits, with each pixel containing a charge-sensitive unit for collecting electrons and outputting electrical signals and its front-end signal processing circuit. The distinct signals generated by particles interacting with different pixels offer precise two-dimensional positional information. Silicon pixel detectors can be classified architecturally into monolithic and hybrid types.

Monolithic active pixel sensor The monolithic type integrates the charge-sensitive unit and front-end circuit on a single substrate, minimizing complexities in interface designs. This integration ensures high granularity, a low material budget, and a relatively small equivalent input capacitance. Monolithic Active Pixel Sensor (MAPS) is currently one of the most promising and well-developed technologies among monolithic types.

Fig. 8
figure 8

(Color online) Left: The array structure of MAPS; Right: The principle of MAPS

MAPS shares the similar detection principle as SSD. As depicted in Fig. 8, a lightly doped P-type epitaxial (EPI) layer grows on a substrate heavily doped with P-type impurities, the Nwell/P-EPI diode is the charge-sensitive unit and the front-end circuit is also on the same substrate connecting to it directly. MAPS typically features pixel sizes ranging from 10–\({30}\,{\upmu \text{m}}\), resulting in a position resolution of 3–\({7}\,{\upmu \text{m}}\). The readout speed can achieve hundreds of kilohertz, and noise is maintained at approximately 10e\(^-\) level. MAPS demonstrates a good balance in resolution, material budget, radiation hardness, readout speed, and power consumption.

The EPI process in MAPS significantly impacts charge collection. The standard EPI process has a thickness of approximately \({10}\,{\upmu \text{m}}\) which causes a small depletion layer, and an electrical resistivity of around 10 \(\Omega\) cm, making thermal diffusion the dominant way for charge collection. Therefore, it needs a long time to collect electrons and the collection efficiency is very low. In contrast to that is an HR EPI process with a thickness of up to \({40}\,{\upmu \text{m}}\) and an electrical resistivity that potentially can reach \(\sim 1000\, \Omega\) cm. The depletion layer is thicker, and charges can also be collected by drifting. The process of collecting is faster and more efficient. Moreover, it reduces the probability of electron–hole pair recombination, enhancing radiation hardness and improving the Signal-to-Noise Ratio (SNR). In the standard twin-well process (Fig. 9 left), PMOS cannot be used in pixel design because the Nwell where PMOS resides competes with the Nwell of the diode, leading to a reduction in Charge Collection Efficiency (CCE). Hence, the deep Pwell is needed to isolate the two. In addition, applying a reverse bias voltage to the substrate could expand the depletion layer region which means increasing CCE and improving radiation hardness, but it may also cause the peripheral readout circuit on the same substrate not to work properly, so deep Nwell is needed for protection. Simulations can give the preliminary study of charge collection in certain process for specific particles in the early phase of design [32,33,34]. In summary, the prevailing approach for MAPS involves the utilization of a complete quadra-well process (Fig. 9 right) with HR EPI.

Fig. 9
figure 9

(Color online) Left: twin-well process; Right: quadra-well process

In experimental physics, MAPS finds primary applications as vertex and track detectors in collider experiments. The Heavy Flavor Tracker (HFT) detector in the STAR experiment is the world’s first detector to use MAPS to enhance vertex resolution and extend measurement capabilities in the heavy flavor domain [35]. The HFT detector incorporates MAPS in its innermost two layers (at radii of 2.8 and 8 cm from the beamline) among four concentric cylinders near the STAR interaction point, and these layers feature 400 Ultimate (MIMOSA-28) chips, covering a total silicon area of 0.16 m\(^2\), as illustrated in Fig. 10. Every layer costs a global material budget of only 0.5% \(X_0\) because the chip on it operates normally with air cooling at room temperature [36].

Fig. 10
figure 10

(Color online) Left: Schematic view of the HFT inside the TPC inner field cage, reproduced from [36]; Right: MIMOSA28 on its PCB. Reproduced from [37]

The Ultimate chip, designed by the IPHC, fabricated by \({0.35}\,{\upmu \text{m}}\) HR OPTO process (400 \(\Omega\) cm), features a pixel array of 928 \(\times\) 960 pixels with a pixel pitch of \({20.7}\,{\upmu \text{m}}\), providing a sensitive area of approximately 3.8 cm\(^2\). Its pixel includes a sensing diode for charge collection as well as an amplification circuit and a Correlated Double Sampling (CDS) circuit for signal extraction and noise removal. Each pixel column is terminated with a high precision discriminator and is read out in a rolling shutter mode at 5 MHz. The discriminator’s outputs are processed through an integrated zero suppression logic and the sparsified data are sent out to the acquisition via two 160 Mbps LVDS outputs. The Ultimate demonstrated a MIPs detection efficiency near to 100%, a fake hit rate of less than 10\(^{-4}\), and power consumption maintained at 150 mW/cm\(^2\) [37].

The ALICE Inner Tracking System 2 (ITS2) stands as the world’s largest detection system employing MAPS. As illustrated in Fig. 11 left, the detector layout features concentric 7 layers with radii ranging from 23 mm to 400 mm. These layers consist of over 24120 ALICE Pixel Detector (ALPIDE) chips (Fig. 11 right) [38], forming a 10 m\(^2\) active detection area with a total of 12.5 \(\times 10^9\) pixels.

Fig. 11
figure 11

(Color online) Left: Layout of the ITS2 detector, reproduced from [39]; Right: The ALPIDE chip. Reproduced from [40]

ALPIDE is fabricated using the TowerJazz 180 nm quadra-well HR CMOS process (\(>1\, \text{k}\Omega\) cm), and features 512 \(\times\) 1024 pixels with the pitch of \({28}\,{\upmu \text{m}}\), providing a sensitive area of 30 mm \(\times\) 13.8 mm [41]. Its pixel includes an octagon collection diode which achieves nearly 100% CCE and a distinctive front-end circuit consisting of a continuously active discriminating amplifier and a multiple-event memory, which yields an ENC of less than 10 e\(^-\). The in-pixel multiple-event memory is read out asynchronously through a priority encoder circuit in each double column. This is fast and power efficient as the expected occupancy is low and only hit pixels are read out in a hit-driven mode. Data are collected at the periphery and shipped off the detector by means of a high-speed serial link [42]. Extensive beam tests confirm that ALPIDE achieves a detection efficiency exceeding 99%, a false hit probability below the required 10\(^{-6}\) level, and a spatial resolution better than the desired \({5}\,{\upmu \text{m}}\). Even after exposure to a TID of 2.7 Mrad and a non-ionizing energy loss (NIEL) flux of 1.7\(\times\)10\(^{13}\) (1 MeV neq/cm\(^2\)), surpassing the anticipated detector lifetime, it sustains performance and provides a readout rate of 100 kHz with a power density of less than 40 mW/cm\(^2\) [43].

MAPS also plays a crucial role in high-resolution beam telescopes for tracking charged particles, exemplified by the Desy II beam telescope. The first generation adopted the EUDET-type framework, employing MIMOSA-26, which provided outstanding spatial resolution and a lower material budget. Achieving a distribution resolution of \({1.83}\,{\upmu \text{m}}\) for a 6 GeV electron or positron beam on the Device Under Test (DUT), the rolling shutter readout operated at a frequency of approximately 10k frames per second. As MIMOSA-26 production ceased and updating became unfeasible, a new beam telescope was constructed based on ALPIDE technology (Fig. 12). This telescope features a sensitive area of 414 mm\(^2\), a trigger rate of up to 100 kHz, and significantly reduced noise levels compared to the previous telescope. This enhancement allows for more efficient beam utilization and a cleaner tracking environment. The spatial resolution remains in the micron range (approximately \({5}\,{\upmu \text{m}}\) at 6 GeV), with a low fake hit rate [44].

Fig. 12
figure 12

(Color online) Design of the ALPIDE beam telescope system. Reproduced from [44]

MAPS is increasingly pivotal in proton and heavy-ion imaging, particularly in the burgeoning field of ion beam radiotherapy. Unlike traditional photon radiation, ion beams exhibit a Bragg peak, a concentrated dose deposition, whose depth can be finely adjusted by altering the beam energy. This feature enhances protection for healthy tissues [45]. Proton and heavy-ion imaging research contributes to refining treatment plans for proton/heavy-ion therapy [46]. Current methods of making particle therapy treatment plans rely on converting X-ray CT HU value into Relative Stop** Power (RSP) of proton/heavy ions, which led to inevitable range uncertainties. However, the Bergen proton CT (PCT) system (Fig. 13), incorporating a Digital Tracking Calorimeter (DTC), can directly measure the RSP to address this problem.

Fig. 13
figure 13

(Color online) Principle of the Bergen proton CT. Reproduced from [47]

The DTC comprises 43 layers, each equipped with 108 ALPIDE chips, which feature the first two layers as tracking layers for acquiring the position information of incoming particles and the subsequent 41 layers as a digital calorimeter. As a particle traverses the detector, it deposits energy in the sensitive layer, generating a 3D digital hit map along its path. The final prototype has a measuring area of 27 cm \(\times\) 16.6 cm, which is adequate for imaging the human head [47].

Another compact full digital system that utilizes the MAPS with a similar structure to Bergen PCT but for heavy-ion imaging is the Hi’CT proposed by IMP (Fig. 14). The simulation results achieve less than 1% deviation in Relative Stop** Power (RSP) and better image quality of CATPHAN phantoms. The real-time imaging capability of Hi’CT may reduce target deviations caused by organ or respiratory movement, achieving optimized treatment effectiveness and improved quality of life for patients [48].

Fig. 14
figure 14

(Color online) Conceptual design of the Hi’CT. Reproduced from [48]

MAPS technology has advanced with the adoption of stitching technique in wafer-level chip manufacturing. The stitching technique is the way to continue the scaling of technology in the semiconductor industry because of the reticle size limitation [49]. It is achieved by dividing the reticle into sections that could individually be exposed and aligned with neighboring sections [50]. As Fig. 15 left shows, the design reticle gets subdivided in sub-frames that correspond to sub-frames of the photomasks. During the photolithographic patterning of wafers, these are selectively exposed to adjacent locations according to a pre-established pattern. This requires very accurate translations and alignment of the wafers between each exposure. The peripheral structures along the outer edges of the core array and at the four corners are designed in dedicated sub-frames of the reticle. The red lines in the figure indicate the dicing lanes. With a judicious design of the geometries in the reticle sub-frames, large chips with diagonals approaching the wafer diameter become feasible, provided their core area can be constructed as an array of sub-units regularly repeating in space. Interconnections across the sub-units are made with the joining of wiring geometries at the abutment boundaries [51].

Fig. 15
figure 15

(Color online) Left: Principles of stitching, reproduced from [51]; Right: Layout of the ITS3 Inner Barrel. Reproduced from [52]

The upgrade of ALICE’s ITS3 (Fig. 15 right) is prepared to utilize stitching technology to manufacture a kind of single MAPS on a 12-inch silicon wafer, achieving dimensions of 27 cm \(\times\) 9 cm. ITS3 is expected to significantly reduce the material budget from 0.35%\(X_0\) to 0.05%\(X_0\) because the new MAPS will be capable of self-bending without the need for complex support and cooling systems. Simultaneously, ITS3 can place the innermost layer at a radial distance of only 18 mm from the interaction point. These features will enhance the resolution of impact parameters to twice that of all momenta and significantly improve tracking efficiency at low transverse momentum [52]. In addition, large scientific facilities such as the Electron-Ion Collider in China (EicC), the Electron-Ion Collider (EIC), and the Nuclotron-based Ion Collider Facility (NICA) are also preparing to adopt wafer-scale MAPS implemented with stitching technology.

As mentioned before, reverse bias can enlarge the depletion layer region, but normal CMOS processes typically have a maximum bias substrate voltage of 5 V [53]. High voltage (HV) CMOS technology is starting to be used in MAPS to improve sensor properties. Unlike the quadra-well process mentioned earlier, in the HV CMOS process, the deep Nwell is used to protect the entire circuit except for the charge-sensitive unit, which is another kind of diode composed of EPI and the deep Nwell itself as shown in Fig. 16. Such a design can even achieve a 100% fill-factor [54]. A typical example is ALTASpix3,originally designed for the upgrade of the ATLAS tracking experiment and a option for the silicon tracking detector of the Circular Electron Positron Collider (CEPC). Its deep Nwell connects to the positive power supply rail at 1.8 V and the maximum voltage of the substrate is around 66 V, corresponding to a depleted region of approximately \({30}\,{\upmu \text{m}}\) [55].

Fig. 16
figure 16

(Color online) The principle of HV MAPS. Reproduced from [55]

The Silicon-On-Insulator (SOI) CMOS technology has also been used to significantly enhance radiation hardness for MAPS design. SOI wafers are composed of a thin Silicon film (device layer) on top of an insulating layer (buried oxide) that is manufactured on a standard Silicon substrate. In traditional SOI technologies, the substrate serves solely as a mechanical support for the silicon film. In detector applications, as shown in Fig. 17, the substrate has high resistivity and is utilized for creating the sensor (a matrix of fully depleted diodes) monolithically coupled to the readout electronics integrated in the silicon film above the insulating layer [56]. Laboratory laser testing of SOI CMOS pixel detectors has demonstrated excellent TID tolerance up to 1 Gray [57].

Fig. 17
figure 17

(Color online) The principle of HV MAPS. Reproduced from [56]

One consideration to address limitations in pixel size is constructing circuit chips in three-dimensional (3D) structures [58]. One method is 3D stacking, illustrated in Fig. 18, employing Through-Si Via (TSV) for vertical communication.

Fig. 18
figure 18

(Color online) Schematic cross-section of the two tiers (sketch not to scale). Reproduced from [59]

Another method, known as 3D detectors, decouples electrode distance and substrate thickness. It allows to place the electrodes perpendicular to the silicon wafer penetrating through the substrate. Therefore, in contrast to standard planar detectors, the depletion region laterally grows between the electrodes [60], as depicted in Fig. 19.

Fig. 19
figure 19

(Color online) Left: Schematic cross-sections of planar sensor; Right: 3D sensor. emphasizing the decoupling of active thickness (\(\Delta\)) and collection distance (L) in 3D sensors. Reproduced from [61]

Nuclear physics experiments also require MAPS to have the capability for time and energy measurements beyond position only. Time measurement is needed to address the high counting rate requirements, such as the China Hyper-Nuclear Spectrometer (CHNS) and precise monitoring of heavy ion beams, which may require counting rates as high as 10\(^6\) to 10\(^8\) counts per second, corresponding to a time precision of 10 ns to \({1}\,{\upmu \text{s}}\). Energy measurement contributes to position resolution and particle identification. For example, the B\(\rho\)-DE-TOF (magnetic rigidity-ionization energy loss-time of flight) method, which combines momentum generated by trajectory curvature and the Bethe-Bloch formula, is mainly used in the identification of particles in medium to high energy radioactive beams and the identification of reaction products on secondary targets. According to the relationship between ionization energy loss \(\Delta E\) and particle charge number Z, and velocity v: \(\Delta E \propto Z^2/v^2\), using \(\Delta E\) combined with TOF measurement can realize particle identification [62]. Meanwhile, for real-time monitoring of heavy ion beams at facilities such as the Heavy Ion Research Facility in Lanzhou (HIAF), the spatial resolution can be improved by calculating the center of mass of large clusters using energy information [63]. It can also provide calibration for time measurements.

MAPS development in China started in the 2010s and has grown rapidly. The representing ones for MIPs detection are the MIC series [64] from Central China Normal University (CCNU), the Taichupix [65] and Jadepix [66] from the Institute of High Energy Physics (IHEP) of the CAS, the SuPix from Shandong University [71]