Mohammad
Rostami
*a,
Mehdi
Janbazi
*b,
Ali
Biganeh
b and
Farnaz
Ghavami
a
aFaculty of Physics, Kharazmi University, Tehran, Iran. E-mail: mrostami@khu.ac.ir
bPhysics and Accelerators School, Nuclear Sciences and Technology Research Institute, Tehran, Iran. E-mail: mehdijanbazi@yahoo.com
First published on 17th November 2025
ZnO is an essential material used in various devices, but its performance can be significantly enhanced by introducing or removing defects, such as oxygen and zinc vacancies. In this article, we explore the relationship between different types of defects in pure and Al/Ga-doped ZnO and their corresponding optical and electronic properties, which are vital for X-ray detection. The nanoparticles were initially synthesized using the sol–gel auto-combustion method, with calcination temperatures of 400 °C, 500 °C, and 600 °C. The samples were then analyzed using multiple techniques, including XRD, FESEM, FTIR, photoluminescence (PL), electron paramagnetic resonance (EPR), and positron annihilation lifetime spectroscopy (PALS). Subsequently, DFT+U calculations were conducted to examine the electrical, optical, and theoretical EPR properties, as well as to identify defects that occur at different calcination temperatures. Analyzing the connection between defects and PL spectra in X-ray scintillation detectors, along with exploring the relationship between electron and hole concentrations in X-ray semiconductor detectors, provides valuable insights into the fundamental properties of ZnO. These insights pave the way for utilizing ZnO in developing next-generation scintillation and semiconductor X-ray detection technologies.
An efficient X-ray detector based on ZnO requires intense luminescence under ultraviolet excitation and the ability to sustain an ultra-low leakage current, typically within a few tens of nanoamperes, under reverse-bias conditions.19,20 Furthermore, in the absence of incident radiation, ZnO should exhibit a low dark current density (typically on the order of a few microamperes per square centimeter) to ensure high signal-to-noise performance in low-background environments.21 Native defects, including vacancies and interstitials, mainly influence the performance of detectors based on ZnO. Such intrinsic defects can, however, result in increased leakage current as well as the generation of visible and infrared light, which, in turn, decreases the response current of a detector. The type, concentration, and distribution of these defects can be associated with annealing temperature and the reaction of the doping elements.22,23 In this context, extrinsic impurities can significantly modify the defects and electronic properties of ZnO.24 Therefore, understanding how dopants such as Al and Ga interact with the defects in ZnO can significantly enhance the performance of ZnO-based radiation detectors.
This study explores how impurities and thermal treatment influence defect formation in pure and Al- or Ga-doped ZnO (3% and 5%), using a combination of PL, EPR, and PALS measurements along with DFT+U calculations. It also analyzes how these factors affect the performance of scintillation and semiconductor detectors, particularly in relation to defects in the materials. Nanoparticles with the chemical formula Zn1−xAlxO and Zn1−xGaxO (x = 0.00, 0.03, and 0.05) are synthesized at calcination temperatures of 400, 500, and 600 °C. Additional discussion on the rationale behind the chosen calcination temperatures and doping concentrations can be found in Section S1 of the SI. XRD, FESEM, FTIR, PL, EPR, and PALS are employed to elucidate the relationships between defect states, excitonic effects, and electronic transitions.
DFT+U calculations were also performed to investigate the electrical and optical properties, with a focus on identifying defects. The origins of typical emissions are discussed, providing new insights into the properties and transition mechanisms of defect luminescence. Moreover, calculations involving the g-tensor and hyperfine coupling constants (HFCCs) are performed to obtain theoretical EPR spectra,25 further facilitating the analysis of defect-related electronic structures. By analyzing the relationships between defects and PL spectra in scintillation detectors and electron–hole concentrations in semiconductor detectors, this study provides insights into the properties of pure and doped ZnO and its potential applications in next-generation scintillation and semiconductor X-ray detectors. This article will be divided into three parts, beginning with an overview of the experimental and computational methods employed. The Results section examines the material's structure, spectroscopic findings (PL, EPR, and PALS), defect stability, electronic properties, and impact of defects on charge carriers, as well as their potential applications in X-ray detectors. The article ends with a conclusion that ties together the main insights and their significance.
Various values for U p(O) and U d(Zn), ranging from 0 to 12 eV, were investigated to ensure precise calculations of the band gap energy, as illustrated in Fig. S1(c). It was found that a cutoff energy of 40 Ry, k-points with dimensions 2 × 3 × 3, a charge density of 320 Ry, and Hubbard potential corrections of U 3d(Zn) = 10 eV, U 2p(O) = 11 eV, U 2p(Al) = 5 eV, and U 3d(Ga) = 3 eV provide optimal convergence.29 The dominant defects were identified based on a comprehensive evaluation of the formation energies, complemented by EPR spectra obtained using the GIPAW method.28 The computational EPR spectra were simulated using the MATLAB-based EasySpin package,30 specifically the pepper subroutine. The simulations incorporated parameters including the g-factor, HFCCs of aluminum and gallium atoms, a microwave frequency of 9.776 GHz, a magnetic field sweep range of 320–390 mT, and a linewidth of 0.9 mT.
![]() | (1) |
nλ = 2dhkl sin θ | (2) |
![]() | ||
| Fig. 1 (a) The XRD diagrams of the pure and doped samples, (b) the FTIR spectrum of pure ZnO, and the FESEM images of (c) pure ZnO, (d) ZnO:Al (3%), and (e) ZnO:Ga (3%) samples. | ||
The crystallite size (Dc) can be evaluated using Scherrer's relation:
![]() | (3) |
FTIR spectroscopy was used to study the chemical bonding and defects (see Fig. 1b and Fig. S2d (SI)). The hydroxyl groups on the nanoparticles are responsible for the peak at approximately 3400 cm−1. The C–H stretching vibrations of alkane functional groups result in peaks between 2830 and 3000 cm−1. The peaks at around 1630 and 1384 cm−1 are due to the asymmetric and symmetric stretching vibrations of the zinc carboxylate functional group, respectively.32 Usually, Zn–O stretching vibrational modes fall between 400 and 500 cm−1. The vacancies in the bulk of the ZnO material are responsible for the peak at 1100 cm−1.33 The extra peaks in the FTIR spectrum of the sample synthesized at 500 °C can be due to the change in ZnO lattice vibrational modes because of vacancies. These vacancies can create new vibrational modes that are not present in a perfect ZnO lattice.34 The shift of the peak positions from 466 to 433 and 489 cm−1 can also be attributed to changes in the Zn–O bond environment due to vacancies. In particular, the peak at 534 cm−1 for the sample calcined at 500 °C is due to vacancies.34,35 However, it would be preferable to use EPR, PALS, and DFT+U studies to further validate the existence of vacancies.
The FESEM micrographs are also presented in Fig. 1(c)–(e). The average size of pure ZnO is approximately 40 nm and decreases with increasing dopant concentrations. The abundance percentages of oxygen, zinc, aluminum, and gallium atoms were obtained using EDS analysis. Fig. S3 and S4 display the EDS elemental maps obtained from ZnO:Al(3%) and ZnO:Ga (3%), respectively. The atomic abundance percentages for all samples are presented in Fig. S5 (SI). It is clear that the dopant addition process is well executed. As will be discussed in the following sections, pure zinc oxide primarily exhibits ZnOVO, ZnOVZn, and
defects. Regarding the defect formation energy (see Section 3.3), the formation energy of the ZnOVO defect increases with rising temperature, while the formation energy of the ZnOVZn defect decreases. As a result, the abundance percentages of oxygen atoms rise, whereas those of zinc atoms fall. It can be noted that for zinc oxide doped with 3% aluminum, the predominant defects are ZnOVO:Al and ZnOVZn:Al. Regarding defect formation energy, as the temperature rises, the formation energy of the ZnOVO:Al defect remains constant, while that of the ZnOVZn:Al defect decreases (see Section 3.3). Consequently, the abundance percentages of oxygen increase, while those of zinc decrease. The same pattern is observed for other doped samples as well.
![]() | ||
| Fig. 2 The deconvolution of the PL spectra for (a) pure, (b) ZnO:Al (3%), (c) ZnO:Al (5%), (d) ZnO:Ga (3%), and (e) ZnO:Ga (5%) samples calcined at 500 °C. | ||
PALS of the pure and 3% doped samples calcined at 500 °C was performed using a digital PALS spectrometer at room temperature. The source was sandwiched between two identical samples. The spectrometer was equipped with a 500 MHz CAEN waveform digitizer. The digitizer samples directly from the anode signals of the two plastic detectors. The timing resolution was measured to be 160 ps employing the coincidence gamma lines of 60Co (1173 and 1332 keV). The channel width was adjusted to 28 ps. The details of the PALS spectrometer are published in our previous study.39 The statistics of 1 million positron annihilation events were compiled over 7 days for each sample. The PALS spectra of the samples were analyzed using the LT-10 code.40Fig. 3 shows the PALS spectra of the pure sample. The results of the data analysis are presented in Table 1. Three parameters were considered for data analysis. τi (ns) is the lifetime of the ith component, and I(i) is the corresponding intensity. The shortest lifetime and its related intensity are attributed to the para-positronium self-annihilation. The intermediate component is due to the positron annihilation with free electrons. The longest-lived component and its associated intensity are related to positron annihilation in free-volume sites.
| Sample | τ 1 (ns) | I 1 (%) | τ 2 (ns) | I 2 (%) | τ 3 (ns) | I 3 (%) | κ (s−1) |
|---|---|---|---|---|---|---|---|
| ZnO | 0.259 ± 0.002 | 65.59 ± 0.13 | 0.714 ± 0.019 | 31.98 ± 0.21 | 1.841 ± 0.031 | 2.43 ± 0.10 | (7.87 ± 0.02) × 108 |
| ZnO:Al (3%) | 0.368 ± 0.001 | 62.41 ± 0.18 | 0.669 ± 0.008 | 29.34 ± 0.17 | 1.732 ± 0.020 | 8.25 ± 0.24 | (3.59 ± 0.06) × 108 |
| ZnO:Ga (3%) | 0.262 ± 0.001 | 63.79 ± 0.07 | 0.743 ± 0.013 | 33.95 ± 0.08 | 1.974 ± 0.026 | 2.26 ± 0.05 | (8.39 ± 0.10) × 108 |
The third-lifetime component (τ3) and its intensity (I3) obtained from PALS provide insights into the presence and nature of large open-volume defects, such as voids or vacancy clusters, where ortho-positronium (o-Ps) formation can occur. In this study, the ZnO:Al (3%) exhibited the highest I3 value (8.25%), indicating a significantly higher concentration of such defects compared to the pure ZnO and ZnO:Ga (3%). This suggests that Al doping induces extensive structural disorder, likely through the formation of vacancy clusters and charge-compensating defects associated with the substitution of Zn by ZnO:Al (3%). Although the τ3 value of ZnO:Al (3%) is slightly shorter (1.732 ns), it still falls within the expected range for o-Ps annihilation in large but possibly interconnected or partially confined voids. In contrast, the ZnO:Ga (3%) shows the longest τ3 (1.974 ns), indicating the presence of larger and more isolated free volumes; however, its I3 parameter is low (2.26%), suggesting a much lower density of such sites. The pure ZnO exhibits intermediate behavior, with τ3 = 1.84 ns and I3 = 2.43%, indicating a relatively compact microstructure with fewer and smaller open-volume defects. These observations highlight the distinct impact of the dopant type on the defect structure and porosity of ZnO at the atomic scale.
To provide a more quantitative interpretation of the PALS results, the standard two-state positron trapping model was applied. The positron trapping rate κ was calculated using the following relationship:
![]() | (4) |
EPR spectroscopy has uncovered significant insights into the defects in the crystal structure of the materials. This technique, suitable for samples with unpaired electrons, was applied to pure and 3% and 5% Al- and Ga-doped zinc oxide powders synthesized at 500 °C. The EPR spectra of the pure and 5% Al- and Ga-doped samples showed that they are diamagnetic, and the experimental spectra only contained noise signals. Fig. 4 shows the EPR spectrum of pure ZnO, and almost the same spectrum is observed for 5% Al/Ga-doped samples. The EPR spectra of 3% doped zinc oxide are shown in Fig. 4, which exhibit resonance signals at magnetic fields of 356 and 357 mT for aluminum and gallium, respectively. The EPR spectrum of zinc oxide doped with 3% gallium has a narrower width compared to that of zinc oxide doped with 3% aluminum. A wider EPR spectrum suggests a greater presence of defects with varying resonance field positions for the 3% Al-doped sample. For more details, refer to Section 3.4 on theoretical EPR spectroscopy.
![]() | ||
| Fig. 4 The experimental EPR spectra of ZnO, ZnO:Al (3%), and ZnO:Ga (3%) samples calcined at 500 °C. | ||
. To examine the effect of impurities and intrinsic defects on the electronic properties, the structure of pure and doped samples with intrinsic defects was carefully optimized. The a and c lattice parameters of pure zinc oxide were calculated to be 3.27 and 5.22 Å, respectively, with an insignificant deviation from the experimental value obtained earlier, indicating the accuracy and validity of the model. Besides, the spin-polarized band structure of pure zinc oxide and the total and partial densities of states of the oxygen and zinc atoms are presented in Fig. 5. The pure zinc oxide semiconductor has a direct band gap of 3.39 eV, which aligns well with the experimental results (3.37 eV).41 The partial density of states reveals that, for all structures, the conduction band minimum of zinc oxide is formed by the overlap of Zn-4s, Zn-3d, and O-2p orbitals. Furthermore, the 2p orbitals of most oxygen atoms essentially form the valence band maximum (VBM).
We calculated the formation energy of the defects to understand the stability of different defects. The focus is on native defects when they are neutral, and we used the following equation to determine the formation energies:
![]() | (5) |
and
. These limits ensure that ZnO remains thermodynamically stable with respect to metallic Zn and molecular O2, in complete agreement with both experimental growth conditions and previous theoretical studies.42–44 As a result, this range offers a relevant and meaningful description of realistic synthesis environments for both pure and doped ZnO. Under the O-rich synthesis conditions,
and
and under the Zn-rich synthesis conditions,
and
. Besides, μAl = EAl(bulk) and
, where EAl(bulk) and EGa(bulk) are the total energies of bulk Al and Ga, respectively.
Synthesis conditions at various calcination temperatures can influence the formation energy and the occurrence of different defects. Low oxygen flow rates at low temperatures are considered O-poor conditions, while high oxygen flow rates at high temperatures are deemed O-rich conditions. Given that nanoparticles were prepared within the equilibrium temperature range (400, 500, and 600 °C), the formation energy for all defects was calculated under equilibrium conditions. To assess the effect of the supercell size on defect formation energies, ZnO:Al, ZnO:Ga, ZnOVO, ZnOVO:Al, and ZnOVO:Ga configurations were modeled in 2 × 2 × 2, 2 × 2 × 3, 2 × 3 × 3, and 3 × 3 × 3 supercells (32, 48, 72, and 108 atoms, respectively) and fully optimized. The corresponding formation energies were calculated using the GGA functional under equilibrium conditions. Additionally, the formation energies were calculated using GGA+U approximations under equilibrium conditions for all states and defects of a 2 × 3 × 3 supercell to investigate the effect of the Hubbard potential (see Fig. S8).
As shown in Fig. S8, the size of the superlattice has little effect on the formation energy of the ZnO:Al and ZnO:Ga states, as well as the ZnOVO, ZnOVO:Al, and ZnOVO:Ga defects. The formation energies of the 2 × 3 × 3 and 3 × 3 × 3 superlattices are nearly identical, with the maximum difference of about 0.38 eV occurring for the ZnO:Al and ZnO:Ga states. The minimal impact of the superlattice size on formation energy results from the neutral charge of the superlattices, which prevents Coulomb interactions between defect charges and their periodic images. In contrast, the Hubbard potential significantly lowers the defects' formation energy relative to the superlattice size. Nonetheless, the trend of decreasing formation energy across all states and defects is consistent, with defect energies being similar in both GGA and GGA+U calculations.
The intrinsic defect formation energies for pure zinc oxide and zinc oxide doped with 3% and 5% aluminum and gallium were calculated under equilibrium conditions, and the results are shown in Fig. 6 and Fig. S9 (SI). To examine the effect of interactions between the dominant defects (oxygen and zinc vacancies), nearby Zn–O vacancy pairs (VOZn) and distant Zn–O vacancy pairs
were also simulated in pure and Al- and Ga-doped samples. Their formation energies are also reported in Fig. 6. As shown in Fig. S9, when aluminum replaces oxygen, the formation energies for defects are significantly higher than when it replaces zinc. Therefore, the probability of aluminum substituting for zinc is higher than that of substituting for oxygen. Under equilibrium temperature conditions, the formation probabilities of ZnOVO and ZnOVZn defects in pure zinc oxide; ZnOOi:Al, ZnOVO:Al, and ZnOVZn:Al defects in 3% Al-doped ZnO; ZnO:Ga, ZnOVO:Ga, and ZnOVZn:Ga defects in 3% Ga-doped ZnO; ZnOOi:Al, ZnOVO:Al, and ZnOVZn:Al defects in 5% Al-doped ZnO; and ZnOVO:Ga and ZnOVZn:Ga defects in 5% Ga-doped zinc oxide are higher than those of other defects. The highest error corresponds to the ZnOVZn:Al (5%) defect, while the lowest error is associated with the ZnOVO:Ga (3%) defect.
![]() | ||
| Fig. 6 Intrinsic defect formation energies for pure and 3% and 5% Al/Ga doped ZnO, with error bars under equilibrium conditions. | ||
![]() | (6) |
The HFCCs for Al and Ga atoms when they replace zinc atoms are negligible, so for each defect, excluding ZnOVOZn:Al and ZnOVOZn:Ga defects, there is a resonant magnetic field whose position is determined only by the value of the g factor. However, when aluminum replaces oxygen and ZnOVOZn:Al and ZnOVOZn:Ga defects, they exhibit a large HFCC, resulting in six resonant magnetic fields (2ml + 1 in number) that are equally spaced, with their separation determined by the hyperfine coupling constant of the aluminum atom. By comparing the position of the resonant field in the theoretical EPR with the experimental spectrum, it is possible to determine whether defects are present or present in negligible concentrations. A comparison of the simulated EPR spectrum of aluminum replacing oxygen with the experimental EPR spectrum reveals that, similar to the defect formation energy results, aluminum cannot replace oxygen. Comparison of the simulated resonance field position with experimental results shows that the concentration of ZnOZni:Al and ZnOVOZn:Al defects for Al-doped materials and the concentration of ZnOZni:Ga, ZnOOi:Ga, ZnOVO:Ga, and ZnOVOZn:Ga defects for Ga-doped samples are insignificant and practically zero. Besides, the concentration of ZnOOi:Al and ZnOVO:Al defects is very low and causes a broadening of the EPR spectrum for the Al-doped samples. A high intensity of the EPR spectrum was observed at the resonance fields of 356 and 357 mT for Al- and Ga-doped zinc oxide, respectively. These peaks indicate that 3% Al-doped ZnO contains one, two, or all three types of states: ZnO:Al, ZnOVZn:Al, and
. At the same time, the 3% Ga-doped sample exhibits ZnO:Ga, ZnOVZn:Ga, and
defects.
From the previous two sections, we found that ZnOVZn, ZnOVO, and
defects are present in pure zinc oxide. For 3% Al-doped ZnO, the ZnOVZn:Al is the dominant defect; however, ZnOOi:Al, ZnOVO:Al, and
defects are less prevalent. ZnOVZn:Ga is the most prevailing defect, and
is present at a lower concentration in 3% Ga-doped ZnO. Therefore, zinc oxide doped with 3% aluminum exhibits the highest defects compared to pure zinc oxide and the 3% Ga-doped sample, which is consistent with PALS results. For the 5% Ga-doped ZnO, ZnOVZn:Ga and
are the most dominant defects, and the same condition is also observed for the 5% Al-doped ZnO, although the ZnOOi:Al defect is present in 5% Al-doped ZnO at a lower concentration.
Fig. 8 presents the spin-dependent band structures and electron density distributions of pure ZnO for the dominant defects, while Fig. S10–S13 (SI) display the results for doped ZnO, along with the spin density distribution. The corresponding figures for the less dominant defects are presented in Fig. S14–S18 (SI). Since the aluminum and gallium density of states do not contribute to the VBM and CBM, they are not shown in the corresponding figures. As can be seen, the electron distribution is centered around most of the oxygen atoms. For zinc oxide doped with 3% Al and Ga, we observe an asymmetry between the spin-up and spin-down states. The asymmetry in the total density of states arises from the 2p orbitals of the oxygen atoms, and as can be seen, the spin density is located around these oxygen atoms.
The presence of defects induced new energy states in the band gap. It changes the optical band splitting, which causes the emission of photons with different energies (under laser light excitation) and also changes the concentration of electrons and holes in the zinc oxide semiconductor. In Fig. 9, the energy of the emitted photons in the presence of defects is compared with the energy of the experimentally emitted photons at synthesis temperatures of 400 to 600 °C. In the experimental PL spectra for all modes, the emitted photon energy is in the ranges of 3.14–3.22 eV (brown triangle symbols), 2.74–2.94 eV (red diamond symbols), and 2.52–2.66 eV (blue plus symbols), which correspond to emissions near the band edge, shallow donor, and deep central states, respectively. The purple cross symbols also correspond to the very deep central states, which are absent in the experimental PL spectra.
![]() | ||
| Fig. 9 Comparison of the energy of photons emitted from defects in pure and doped zinc oxide with the energy of experimentally emitted photons at the calcination temperatures of 400 to 600 °C. | ||
Comparison of the photon energies emitted from the experimental PL spectra with the photon energies calculated from the band structures shows that the emitted photon energy corresponding to the near band edge (brown triangle) is obtained from pure ZnO. Besides, the two emitted photon energies corresponding to the shallow donor states (red rhombus) and the deep central states (blue sum) are obtained from the ZnOVZn or
, and three ZnOVO, ZnOVZn and
defects, respectively. For ZnO doped with 3% Al, the energies of the emitted photons corresponding to the shallow donor state, the deep central state, and the one near the band edge are obtained from ZnOVZn:Al, ZnOVO:Al defects and mainly from the ZnO:Al state (with a negligible amount from ZnOOi:Al and
defects), respectively. For zinc oxide doped with 5% aluminum, the energy of the emitted photons corresponding to the shallow donor level, the deep central level and the one near the band edge comes from ZnOVZn:Al defects,
defects and mainly the ZnO:Al state (with a small amount of ZnOOi:Al defect), respectively. For 3% Ga-doped zinc oxide, the energies of the emitted photons corresponding to the shallow donor states, the deep central state, and the one near the band edge are obtained from ZnOVZn:Ga defects,
defects, and ZnO:Ga state, respectively. Finally, for 5% gallium-doped zinc oxide, the energy of the emitted photons corresponding to the shallow donor level and the deep central level originates from the
defects, and the photon energy corresponding to NBE originates from the ZnO:Ga state.
The carrier concentrations and free volumes at room temperature for pure and doped zinc oxide with dominant defects are reported in Table S1 of the SI. Section S2.2 of the SI provides the theoretical framework and the procedure used to calculate the charge carrier density. It is found that zinc vacancy defects have minimal impact, while oxygen vacancy defects lead to an increase in electron concentration. According to Table S1, zinc oxide doped with 3% aluminum has higher electron and hole concentrations than pure zinc oxide and zinc oxide doped with 3% gallium, which is in good agreement with the PALS results. Table S1 shows that adding aluminum and gallium does not significantly affect the free volumes, while the free volumes for zinc vacancies are greater than those for oxygen vacancies. The presence of oxygen and zinc vacancies simultaneously and at a distance from each other increases the free space relative to the zinc vacancy. Therefore, the free volumes for zinc oxide doped with 3% gallium are increased compared to those for pure and 3% Al-doped ZnO and are in good agreement with the PALS results. Increasing the aluminum and gallium doping concentrations results in a significant increase in both the electron concentration and the free volume.
The comprehensive identification of intrinsic defects through PL, EPR, PALS, and DFT+U analyses allows us to link microscopic defect structures with macroscopic detector performance. In scintillation detectors, strong luminescence near the band edge and weak luminescence in defects, especially those related to deep-level energy emission, are required to achieve fast-response X-ray detection.46,47 However, defects that induce shallow trap-to-NBE transitions may also suppress visible emissions and enhance near-band-edge (NBE) emission, which improves fast response and timing resolution. Conversely, defects that induce deep trap-to-NBE transitions act as radiative recombination centers, increasing visible luminescence but introducing slow decay components that reduce temporal resolution in scintillation detectors. For pure and doped zinc oxide, the PL spectrum exhibited three photon emission peaks in the energy ranges of 3.14–3.22 eV (NBE), 2.74–2.94 eV (shallow donor states), and 2.52–2.66 eV (deep central states). The lower the intensity of shallow donor state and deep central state peaks relative to the NBE, the faster the detector response. For scintillator applications, Fig. 2 shows that the shallow trap-to-NBE and deep trap-to-NBE intensity ratios are 0.19 and 0.08, respectively, which are considerably lower than those of the other samples. This indicates lower defect concentrations and thus a more perfect crystal lattice, consistent with the higher defect formation energies of ZnOVZn:Ga (3%) (1.16 eV) and
(6.64 eV) in the 3% Ga-doped sample (Fig. 6). Therefore, a scintillation detector made of zinc oxide doped with 3% gallium can have a faster response than the other samples.
Detecting X-rays with semiconductor devices involves optimizing two key aspects: achieving rapid device responses and detecting weak signals. To do this, researchers often rely on optical conductors that cover a wide area and can keep unwanted currents, especially dark current, to an absolute minimum. ZnO-based detectors should exhibit low dark current, which enhances the signal-to-noise ratio and improves energy resolution. However, to keep the dark current low, the material needs to have very few free carriers before it is exposed to X-rays. Additionally, one should also consider the issue of trap states, which are imperfections or defects within the material that can trap electrons or holes. Sometimes, they release visible light, and sometimes they do not; however, what they do create is interference. They boost noise and reduce sensitivity. For semiconductor-type X-ray detectors, reducing the oxygen vacancy concentration minimizes trap-assisted leakage currents, while controlled doping with gallium effectively decreases defect density and stabilizes the electron concentration, thereby enhancing sensitivity and reducing noise. Therefore, optimizing synthesis temperature and dopant concentration to minimize the oxygen vacancy while maintaining structural integrity is the key pathway for achieving both high-speed scintillation and high-sensitivity semiconductor X-ray detection performance. According to DFT calculations and PALS, the 3% Ga-doped sample exhibits the lowest carrier concentration (6.22 × 109 cm−3) compared with 1.82 × 1013 cm−3 for pure ZnO, 5.14 × 1012 cm−3 for 3% Al, 8.39 × 1019 cm−3 for 5% Al, and 4.77 × 1016 cm−3 for 5% Ga, suggesting reduced leakage current and noise. It therefore exhibits reduced leakage current and noise, making 3% Ga-doped zinc oxide an ideal material for highly sensitive semiconductor X-ray detectors. The large amount of free volume causes the electrons and holes to remain there for a long time after the X-ray radiation is produced, allowing them to recombine later and thus slowing down the detector response. Although the free volume for 3% Ga-doped ZnO is larger than that for 3% Al-doped ZnO and pure ZnO, the number of significant free volumes created by
defects is small. Furthermore, this is less significant compared to the sensitivity of the semiconductor detectors.
(i) We found that at three calcination temperatures, ZnOVO, ZnOVZn, and
defects are present in pure zinc oxide. In samples doped with 3% Al, the ZnOVZn:Al defect can be identified at higher concentrations, while ZnOVO:Al, ZnOOi:Al, and
defects occur at very low concentrations. ZnOVZn:Ga is the most prevailing defect, and
is present at a lower concentration in 3% Ga-doped ZnO. For the 5% Ga-doped ZnO, the most dominant defect is
. A similar pattern is observed in the 5% Al-doped ZnO, although the ZnOOi:Al and ZnOVZn:Al defects appear at a lower concentration.
(ii) 3% Al-doped ZnO exhibits higher electron and hole concentrations compared to both pure zinc oxide and 3% Ga-doped ZnO.
(iii) Increasing the aluminum and gallium doping concentrations results in a significant increase in both the electron concentration and the free volume.
(iv) The 3% Ga-doped ZnO sample exhibits the lowest shallow and deep trap-to-NBE intensity ratios (0.19 and 0.08), consistent with higher defect formation energies and a reduced defect density, indicating superior crystal quality and a faster response for scintillator applications.
(v) It also exhibits the lowest carrier concentration (6.22 × 109 cm−3) among all studied samples, which minimizes leakage current and noise, making 3% Ga-doped ZnO a highly promising candidate for sensitive semiconductor X-ray detectors.
Supplementary information (SI) is available. See DOI: https://doi.org/10.1039/d5cp03107a.
| This journal is © the Owner Societies 2026 |