Next Article in Journal
Avalanche Multiplication in Two-Dimensional Layered Materials: Principles and Applications
Previous Article in Journal
Electrocatalytic Innovations at Atomic Scale: From Single-Atom to Periodic Ensembles for Sustainable Energy Conversion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Selection and Experimental Verification of Wide-Bandgap Semiconductor for Betavoltaic Battery

by
Jiachen Zhang
1,*,
Kunlun Lv
1,
Yuan Yin
1,
Yuqian Gao
1,
Ye Tian
1,
Yuncheng Han
2 and
Jun Tang
1
1
School of Semiconductors and Physics, North University of China, Taiyuan 030051, China
2
Hefei Institutes of Physical Science, Chinese Academy of Sciences, Hefei 230031, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(9), 635; https://doi.org/10.3390/nano15090635
Submission received: 19 February 2025 / Revised: 17 March 2025 / Accepted: 21 March 2025 / Published: 22 April 2025
(This article belongs to the Section Theory and Simulation of Nanostructures)

Abstract

:
Wide-bandgap semiconductor betavoltaic batteries have a promising prospect in Micro-Electro-Mechanical Systems for high power density and long working life, but their material selection is still controversial. Specifically, the silicon carbide (SiC) betavoltaic battery was reported to have higher efficiency, although its bandgap is lower than that of gallium nitride (GaN) or diamond, which is inconsistent with general assumptions. In this work, the effects of different semiconductor characteristics on the battery energy conversion process are systematically analyzed to explain this phenomenon, including beta particle energy deposition, electron–hole pair (EHP) creation energy and EHPs collection efficiency. Device efficiencies of the betavoltaic battery using SiC, GaN, diamond, gallium oxide (Ga2O3), aluminum nitride (AlN) and boron nitride (BN) are compared to determine the optimum semiconductor. Results show that SiC for the betavoltaic battery has higher efficiency than GaN, Ga2O3 and AlN because of higher EHPs collection efficiency, less energy loss and fewer material defects, which is the optimal selection currently. SiC betavoltaic batteries were prepared, with the device efficiency having reached 14.88% under an electron beam, and the device efficiency recorded as 7.31% under an isotope source, which are consistent with the predicted results. This work provides a theoretical and experimental foundation for the material selection of betavoltaic batteries.

1. Introduction

With the rapid development of micro-electro-mechanical systems (MEMS), micro-energy generators are especially in demand in extreme environmental conditions, such as deep space, the deep sea and the polar region, where there are extreme ambient temperatures, weak sunlight intensity and high maintenance costs [1]. Conventional micro-electrochemical batteries require frequent charging and degrade easily in extreme environments, which limits their application in MEMS [2]. The betavoltaic battery is an energy generator converting the decay energy of isotope sources into electricity, which has significant potential as an alternative energy generator due to its long operational life, high energy density, easy miniaturization and strong resistance to extreme environments.
A betavoltaic battery mainly consists of an isotope source and a semiconductor energy converter [3]. The working principle of the betavoltaic battery is shown in Figure 1a, which consists of the following steps: beta particles decay from the isotope source, with part of them emitting from the source surface, passing through the electrode and injecting into the semiconductor material; the electron–hole pairs (EHPs) are created and then separated by the built-in electric field of the semiconductor junction, such as the Schottky or PN junction; and, finally, electrons and holes are collected by electrodes to generate currents. During the transportation processes of beta particles, self-absorption of the isotope source, blocking of the electrode metal and backscattering of the semiconductor material lead to energy loss, which makes no contribution to the battery output performance. In processes of EHPs separation and collection in semiconductors, the EHPs recombination and electrode resistance will further reduce the output performance. In addition, during the generation of EHPs, the EHP creation energy of the semiconductor will also affect the output performance of the battery. The higher EHP creation energy usually leads to a smaller radiation generation current because of fewer EHPs created.
The total energy conversion efficiency (η) refers to the proportion of battery output power to isotope source power, which can be divided into isotope source efficiency (ηs) and semiconductor device efficiency (ηd). The ηs is inversely proportional to the thickness of the isotope source. A thinner isotope source tends to have a higher source efficiency and results in a lower battery power density at the same time. To determine the optimum thickness of the isotope source, the ηs and power density of the betavoltaic battery should be balanced to obtain a high output power [4]. 63Ni is the most common isotope source for betavoltaic batteries due to its long half-life of 100 years and soft beta particle energy below the irradiation damage threshold of the semiconductor lattice, which is about 200 keV [5]. The corresponding ηs is about 35% when the thickness of 63Ni is 500 nm with single-side emission. An electron beam from a scanning electron microscope is often used to simulate isotope sources because of its controllable and accurate location, direction and energy.
Figure 1. Models of the device. (a) Typical principles of a betavoltaic battery with PN junction. (b) Larry Olsen’s prediction and the device efficiency of a betavoltaic battery experimental device [6,7,8,9,10,11,12,13,14,15]. (c) The ideal simulation model of the betavoltaic battery; the colors in the figure are used to highlight the layering. (d) The prediction model of the betavoltaic battery experimental device. (e) Simulation model reflecting actual experimental conditions.
Figure 1. Models of the device. (a) Typical principles of a betavoltaic battery with PN junction. (b) Larry Olsen’s prediction and the device efficiency of a betavoltaic battery experimental device [6,7,8,9,10,11,12,13,14,15]. (c) The ideal simulation model of the betavoltaic battery; the colors in the figure are used to highlight the layering. (d) The prediction model of the betavoltaic battery experimental device. (e) Simulation model reflecting actual experimental conditions.
Nanomaterials 15 00635 g001
The ηd is closely related to the characteristics of the used semiconductor material. Larry Olsen [3] proposed that ηd is positively correlated with the semiconductor bandgap, i.e., betavoltaic batteries with wider bandgap semiconductors have higher efficiency, as shown in Figure 1b. Typical wide-bandgap semiconductors used for betavoltaic batteries include silicon carbide (SiC), gallium nitride (GaN) and diamond, with bandgaps as 3.27 eV, 3.44 eV and 5.47 eV, respectively. However, the selection of optimal wide-bandgap semiconductors for high-efficiency betavoltaic batteries is still controversial. Chen Zhao et al. [6] considered SiC as the optimum energy converter for betavoltaic batteries because it has a low recombination current. Szymon Grzanka et al. [7] chose GaN as the semiconductor material for a betavoltaic battery due to its high material density, which increases the absorption of beta particles. C. Delfaure et al. [8] selected diamond as the energy converter for the reasons of the low electron backscattering and the high diffusion length of EHPs. GaN and diamond have been widely studied in betavoltaic batteries in recent years due to their large bandgap and strong irradiation resistance, with their maximum experimental achieved ηd as 6.6% [9] and 9.4% [8], respectively, as shown in Figure 1b, while the experimental ηd of SiC betavoltaic batteries were already more than 12% [6,10,11], being significantly higher than those of GaN and diamond betavoltaic batteries, which is inconsistent with Larry Olsen’s prediction. We need further systematic analysis to obtain a better efficiency performance of betavoltaic batteries.
Recently developed ultra-wide-bandgap semiconductors (Eg > 4 eV), such as gallium oxide (Ga2O3), aluminum nitride (AlN) and boron nitride (BN), have promising applications in betavoltaic batteries in the future due to their large bandgap, strong irradiation resistance and low intrinsic carrier concentration. In addition, combined with innovations in thermoelectric and piezoelectric materials, these semiconductors show great potential for energy recovery, self-powered sensors and more [16,17]. Although the material fabrication technologies of ultra-wide-bandgap semiconductors are not mature enough to prepare stable devices [18], the theoretical study of betavoltaic batteries with ultra-wide-bandgap semiconductors is worth anticipating.
The purpose of this work is to establish the optimal selection principles for the wide-bandgap semiconductors applied in betavoltaic batteries through systematical analysis and experimental verification. By analyzing principal factors affecting the EHPs generation and transport processes, including beta particle energy deposition, EHP creation energy, and EHPs collection efficiency, the ηd of betavoltaic batteries using different semiconductors (SiC, GaN and diamond) were compared by empirical formula calculation and Monte Carlo simulation. Furthermore, the theoretical ηd of betavoltaic batteries with ultra-wide-bandgap semiconductors were predicted. The optimal wide-bandgap semiconductor betavoltaic battery device is designed and fabricated experimentally, and its output current–voltage (I–V) characteristics were measured and compared with those of betavoltaic batteries from other semiconductor experiments.

2. Methods

2.1. Theoretical Model

2.1.1. Efficiency Calculation

The total efficiency η is one of the principal factors for evaluating the performance of the betavoltaic battery, which can be described as
η = P m a x P i n = I s c · V o c · F F A · E a v e · q
where Pmax is the maximum output power of the battery; Pin is the power of the isotope source; Isc is the short-circuit current; Voc is the open-circuit voltage; FF is the filling factor; A is the activity of the isotope source; Eave is the average energy of isotope source charged particles; and q is the electron charge.
Voc, Isc and FF can be obtained by analyzing the I–V characteristics of betavoltaic battery experimental devices and simulating beta particles’ energy deposition combined with empirical formulas. Isc can be expressed as in [19].
I s c = 0 H q · G x · C E x d x
where H is the thickness of the semiconductor; x is the distance between the energy deposition location of the incident particle and the semiconductor surface; G(x) is the production rate of EHPs; and CE(x) is the collection rate of EHPs.
G(x) is closely related to the energy deposition of beta particles and can be expressed as
G x = A · E x ε
where ε is the EHP creation energy, which can represent the energy to generate one single EHP; and E(x) is the energy deposition function.
CE(x) is related to the position of EHPs generation and the semiconductor junction region and can be expressed as in [20].
C E x = 1 t a n h x L p , x > 0       1 , x 0
where Lp is the diffusion length of minority carriers, which is closely related to material defects; and x is the distance between the EHPs generation position and the junction region, which is related to the width of junction region W. When x ≤ 0, EHPs are generated in the junction region and can be collected completely (i.e., CE(x) is 100%).
The width of the PN junction region is affected by the intrinsic carrier concentration ni and the doping concentration of the semiconductor and can be expressed as follows, as in [21].
W = V b i 2 ε r ε 0 q N A + N D N A N D = 2 k T ε 0 ε r q 2 · ln N A N D n i 2 · N A + N D N A N D
where Vbi is the built-in potential of the PN junction; ε0 is the vacuum permittivity; εr is the relative permittivity of semiconductor; NA is the P-type semiconductor doping concentration; ND is the N-type semiconductor doping concentration; k is the Boltzmann constant; and T is the working temperature.
Considering that the recombination current affects the I–V characteristics, the Voc can be expressed as in [10].
V o c = n k T q l n I s c I 0 + 1
where n is the ideality factor, which reflects the proportional relationship between the diffusion current and the recombination current in PN junction, generally ranging from 1 to 2 [22]; and I0 is the proportional coefficient, which is related to the diffusion coefficient, minority lifetime and diffusion length of minority carriers [10].
Without considering the influence of series and parallel resistances, FF can be derived from the Shockley equation and expressed as:
F F 1 1 l n I S C I 0 1 l n l n I S C I 0 l n I S C I 0

2.1.2. Upper-Limit Efficiency

Larry Olsen assumed that CE(x) in the semiconductor is 100%, and I0 only considers radiation recombination; the upper-limit device efficiency can be deduced by Formulas (1), (2) and (6), as in [3].
η d = V o c · F F ε
where ε was described by Shockley and Klein based on experimental results, consisting of bandgap losses Eg, thermalization losses Ef, and optical phonon interactions γEγ. According to the simple two-band (STB) approximation, where Ef = 0.9 Eg, ε is approximately as in [3,20,23].
ε = E g + 2 E f + γ E γ = 2.8 E g + γ E γ 2.8 E g + 0.5   e V
The relationship between the EHP creation energy and bandgap in Equation (9) is also known as the Shockley–Klein relationship.

2.1.3. Energy Deposition Simulation

The performance of the betavoltaic battery with different semiconductors can be predicted based on Formulas (1)–(7) by combination with Monte Carlo simulations.
The ideal theoretical model consists of an isotope source and a semiconductor. The semiconductor includes P-type and I-type regions, and the doping concentrations were 1 × 1019 cm−3 and 1 × 1015 cm−3, respectively. The junction region widths of each semiconductor were 1.80 μm (SiC), 1.94 μm (GaN) and 1.82 μm (diamond) at this doping concentration, according to Equation (5). The semiconductor was segmented into 450 layers, with a minimum thickness of 0.1 μm to ensure the simulation accuracy, as shown in Figure 1c. The isotope source was set as one side-emitting surface 63Ni source, and the source efficiency of the mono-energy electron beam was assumed as 100%.
The prediction model of the experimental device consists of electrodes, a passivation layer, a 0.2 μm P-type semiconductor region and a 5 μm I-type semiconductor region, as shown in Figure 1d. The isotope source was a volume source, with consideration of the self-absorption effect, and the electron beam was a unidirectional-emitting single-energy surface source with a beam intensity of 90 pA (equivalent to 15 mCi). The recombination effects associated, because of defects introduced during fabrication, were not considered, and the ideal factor of the experimental device was set to 1.
The controversy mentioned in Section 1 can be theoretically analyzed by using the theoretical model mentioned above to determine the current optimal selection of wide-bandgap semiconductor for the betavoltaic battery. The output performance of a betavoltaic battery with the selected semiconductor will be further verified by the following experiments.

2.2. Experiment

2.2.1. Fabrication

As demonstrated in Figure 1d, the SiC-based betavoltaic battery device conceptualized in this study incorporates semiconductors with PIN structures, surface passivation and electrodes. The N-type semiconductor substrate has a resistivity of 0.017 Ω·cm with a thickness of 365.5 μm. The P-type and I-type structures of the semiconductors were fabricated by chemical vapor deposition, corresponding to lightly doped N-type layers and heavily doped P-type layers, respectively. The N-type layer exhibits a doping concentration of 1 × 1015 cm−3 and a thickness of 5 μm, while the P-type layer demonstrates a doping concentration of 1 × 1019 cm−3 and a thickness of 0.2 μm. The effective area of the prepared semiconductor transducer device is 4 mm × 4 mm. The surface passivation was performed by dry oxygen oxidation for 3 h at 1100 °C and annealing in N2 with constant temperature for 1 h. The passivation layer covered all surfaces of the PIN structure except the electrode to eliminate the surface state and reduce leakage current, with a thickness of about 24 nm. The electrodes were deposited by magnetron sputtering with a multilayer structure and subjected to rapid thermal annealing (RTA) at 1050 °C for 3 min in N2 atmosphere to form Ohmic contact. In the present study, Ni/Ti/Al/Ni multilayers were selected as electrodes on the P-type semiconductor surface, patterned in a grid-like pattern with 9% of the total active area in order to minimize the loss of penetration energy of beta particles, with a thickness of less than 200 nm. Similarly, Ni/Ti/Al multilayers were selected as electrodes on the N-type semiconductor surface, and the electrodes on the N-type substrate were completely covered in order to ensure low electrode resistance.

2.2.2. Measurement

I–V characteristics of the devices were maintained both under the condition of isotope source and of electron beam. A Keithley 4200A-SCS semiconductor parameter analyzer (Tektronix, Beaverton, OR, USA)obtained I–V characteristics of devices under an isotope source, with the devices placed in a Faraday shielding chamber to prevent the effects of natural light and electromagnetic interference. The 63Ni isotope source has an active area of 1 × 3 cm2 with total activity as 15 mCi, and the physical diagram of the devices is shown in Figure 2. The electron beam simulation measurement system was built by the China Academy of Engineering Physics and consisted of a scanning electron microscope (KYKY-EM6200, KYKY Technology Co., Ltd., Beijing, China) to generate electron beams with different kinetic energy, a Faraday Cup to measure the electron beam current generated by the electron gun and a Keithley 6487 picoammeter (Tektronix, Beaverton, OR, USA) to measure the I–V characteristics of the devices [24]. The acceleration voltage of the electron beam is 20 kV with a fixed beam intensity of 90 pA, which simulates the energy of emitted beta particles of the 63Ni isotope source.

3. Results and Discussion

3.1. Selection of Typical Wide-Bandgap Semiconductor

3.1.1. Effect of the Electron–Hole Pair Creation Energy

Assuming that the collection rate of EHPs is 100% and the energy of beta particles is all deposited in the semiconductor, the upper-limit device efficiency of the betavoltaic battery is mainly affected by the EHP creation energy. In Olsen’s prediction, the EHP creation energy followed the Shockley–Klein relationship with the bandgap, and the upper-limit device efficiency is shown as the black line in Figure 3a, according to Formulas (2) and (6)–(9). It can be seen that the upper-limit device efficiency increases positively with the semiconductor bandgap, reaching to a maximum of about 30%.
However, the validity of the Shockley–Klein relationship between the ε and the bandgap in Equation (9) has been widely questioned [25,26,27,28,29]. Because it is based on the early semiconductor’s experimental results, where the semiconductor materials had many defects and low carrier collection efficiencies, the ε predicted by the Shockley–Klein relationship produced high apparent values [30]. In addition, the complex bandgap structures of wide-bandgap semiconductors cause inter-band tunneling and make the STB approximation not as accurate as classical semiconductors with a narrow bandgap, which violates the hypothesis described as Equation (9), and the ε will be lower than the Shockley–Klein relationship’s prediction [22]. According to the experimental results of the ε, many studies have modified the Shockley–Klein relationship and re-fitted various relationships, as shown in Figure 3b [25,26,27,28,29]. However, these relationships still cannot describe the relationship between the ε experimental value and the bandgap well, which will lead to an underestimation of the theoretical efficiency of the betavoltaic battery [25].
The red dots in Figure 3b show the currently available experimental ε values of SiC, GaN and diamond [22,23,25,26,27,28,29,30,31,32,33,34,35,36]. Many studies have been conducted on the EHP generation energy of SiC, because SiC is a good candidate for high-temperature alpha detectors due to its wide bandgap and low noise. Meanwhile, the EHP creation energy of GaN has been less studied and is more advantageous in X-ray detectors due to its large atomic number. The experimental ε of GaN is closer to the Shockley–Klein relationship than that of SiC and diamond, as shown in Figure 3b, which means its ratio to the bandgap is highest, and the upper-limit efficiency of a GaN betavoltaic battery will be the lowest according to Equation (8). The modified upper-limit device efficiency is shown as the red line in Figure 3a based on experimental values ε′, where the error lines around the red dots represent the results for different ε′. As can be seen from Figure 3a, the upper-limit device efficiency is no longer positively correlated with the bandgap and is higher than Larry Olsen’s prediction, since the ε′ and the bandgap do not satisfy the Shockley–Klein relationship.
Figure 3. Effect of the electron-hole pair creation energy. (a) The upper-limit device efficiencies of a betavoltaic battery with different semiconductors based on the Shockley-Klein relationship of ε (Olsen’s prediction) and experimental ε′ values (this work). (b) Experimental and theoretical values of EHP creation energy for different semiconductors [20,23,25,26,27,28,30,32,33,36].
Figure 3. Effect of the electron-hole pair creation energy. (a) The upper-limit device efficiencies of a betavoltaic battery with different semiconductors based on the Shockley-Klein relationship of ε (Olsen’s prediction) and experimental ε′ values (this work). (b) Experimental and theoretical values of EHP creation energy for different semiconductors [20,23,25,26,27,28,30,32,33,36].
Nanomaterials 15 00635 g003

3.1.2. Effect of the Energy Deposition of Beta Particles

The energy of beta particles is not all deposited in the semiconductor, as in Olsen’s prediction, but a fraction of electrons is returned from the semiconductor surface by elastic scattering, called backscattering, and its amount is closely related to the nucleus size of the semiconductor. The interaction process and transport path of beta particles and different semiconductors are simulated through the Geant4 program, as shown in Figure 4a. In this simulation, the beta particle sources are set as a point source to clearly display the physical process. It can be seen that GaN has the greatest backscattering caused by the highest density and largest atomic number compared with those of SiC and diamond.
The distribution of the energy deposition of beta particles in semiconductors is also important for the battery’s efficiency, which affects the EHPs generation process, as shown in Equations (2) and (3). The energy deposition distribution of beta particles in each semiconductor was simulated by a continuous-spectrum one-side-emitting surface 63Ni source, as shown in Figure 4b. Due to the large atomic number, the energy of beta particles deposited in GaN is mainly closer to the incident surface than that of SiC and diamond, resulting in more energy deposited in the P-type layer. In the PIN structure device, the junction region mainly falls in the N-type layer with a low doping concentration, while it is thin in the P-type layer with ahigh doping concentration. The width of the depletion region in the P-type layer is only 18.0 nm (SiC), 19.4 nm (GaN) and 18.2 nm (diamond) at the doping concentration in Section 2.1.3, and the energy deposited in a P-type layer is not conducive to EHPs collection. Figure 4b shows the proportion of the energy deposition of beta particles in SiC, GaN and diamond at different depths to the total energy of the 63Ni source. The total deposition energy of beta particles in GaN is 8% lower than that of SiC and 13% lower than that of diamond due to the previously mentioned backscattering, and the energy deposited in the first 5 μm of the junction region is also lower than that of SiC and diamond. These phenomena indicate that there is more energy loss in GaN, which has a negative impact on the battery’s energy conversion efficiency.

3.1.3. Effect of the Electron–Hole Pairs Collection Efficiency

The EHP collection process has a significant impact on the efficiency of the battery [37,38]. The CE(x) outside the semiconductor junction region is actually less than 100%, and it is determined by the junction width W and the minority carrier diffusion length Lp, as in Equation (4). As mentioned in Section 2.1.3, there is no obvious difference in the junction widths for SiC, GaN and diamond under the same doping concentration conditions, which means that Lp significantly affects CE(x). The Lp is closely related to the semiconductor band structure, impurity content and surface state, and it is affected by direct recombination, surface recombination, interface recombination and material defects [22]. GaN is a direct bandgap semiconductor, and its Lp is shorter than that of SiC, diamond and other indirect bandgap semiconductors [39]. Moreover, different from the homogeneous epitaxy of SiC and diamond, GaN is generally prepared by heteroepitaxial growth due to technical limitations. The lattice mismatch between the GaN epitaxy material and the substrate will introduce extra defects [40], which greatly reduces the Lp of GaN. The Lp of different semiconductors are shown in Figure 5a. The Lp of GaN is 1.9 μm, which is the smallest among the listed semiconductor materials. According to Equation (4), the minimum Lp of GaN will result in its CE(x) outside the junction region being lower than that of other semiconductors. The CE(x) of different semiconductors is shown in Figure 5a, which shows the CE(x) of GaN significantly decreases outside the junction region. The CE(x) of GaN has been reduced to 35% at 5 μm from the junction edge, and almost no charge carriers are collected beyond 12 μm. In contrast, the CE(x) of SiC is significantly higher than that of GaN, with 62% at 5 μm and 14% at 12 μm, and the CE(x) of diamond is higher, with 72% at 5 μm and 26% still at 12 μm. The higher collection efficiency of SiC and diamond than GaN ensures that the EHPs generated outside the junction region can be collected more, which will be beneficial to obtain higher energy conversion efficiency.

3.1.4. The Ideal Efficiency

The radiation-generated current distribution of SiC, GaN and diamond betavoltaic batteries can be obtained by considering the energy deposition distribution, average ionization energy and collection efficiency, as shown in Figure 5b. It can be seen that the short-circuit current of GaN is larger than that of diamond and SiC in the first 1 μm and then lower after 1 μm. This is due to the high atomic number and density of GaN, which makes the energy deposition distribution of beta particles in GaN shallow, leading to the radiation-generated current of its battery being mostly concentrated near the surface. The ideal efficiency of a betavoltaic battery using SiC, GaN or diamond is shown in Table 1, without considering the effects of isotope source self-absorption, device electrodes and passivation. The large bandgap of semiconductors will make it difficult for electrons to transition from the valence band to the conduction band, resulting in a low intrinsic carrier concentration. According to Equation (6), the lower the intrinsic carrier concentration of the semiconductor, the higher the PN junction’s built-in electric field, with the same condition of doping concentrations. This means that the betavoltaic battery using semiconductors with a larger bandgap has a higher Voc, and the Voc of the GaN betavoltaic battery tends to be actually higher than that of the SiC betavoltaic battery. However, due to the smaller collection efficiency and higher energy loss, the GaN betavoltaic battery has about 20% lower radiation-generated current than the SiC betavoltaic battery, resulting in its ideal efficiency being the lowest among the three kinds of materials. Generally speaking, a betavoltaic battery using a wide-bandgap semiconductor with an indirect bandgap, small atomic number and low density will achieve high efficiency.
Diamond has the highest theoretical efficiency as a betavoltaic battery energy converter, but the development of its experimental devices with a PIN junction is still a challenging task [41]. N-type diamond materials have unacceptable plenty defects due to the immature doping technology, and it is difficult to form a stable and reliable PN junction. The performance of diamond experimental devices is still far below expectations, with the Voc of PIN-type devices only 0.2–1 V [42,43,44]. In general, SiC will be more suitable for betavoltaic batteries when considering both theoretical efficiency and material preparation maturity.

3.2. Prediction of Ultra-Wide-Bandgap Semiconductor

The device efficiency ηd is related to the EHP creation energy ε, energy deposition of beta particles E(x) and collection efficiency of EHPs CE(x), which can be summarized as the following equation:
η d = f ε , E x , C E x
Compared with Larry Olsen’s prediction, Equation (10) corrects ε as ε′ based on experimental results and further emphasizes the effect of E(x) and CE(x). The characteristics of representative ultra-wide-bandgap semiconductors are shown in Table 2. The ηd changing with different semiconductors under different conditions is shown in Figure 5c. Correcting ε to ε′ based on the black line, the result is shown as the blue line. The yellow and red lines were obtained by further considering of E(x) and E(x) together with CE(x), respectively. It can be seen that EHP creation energy ε playing a decisive role in the trend, i.e., experimental ε′ is not positively proportional to the Eg, leads to device efficiency ηd not being simply proportional to Eg. The introduction of the energy deposition E(x) effect leads to a decrease in ηd, which is more pronounced for a betavoltaic battery using semiconductors with higher atomic numbers and material density, such as GaN and Ga2O3. The introduction of CE(x) further reduces the device efficiency, which is especially pronounced for AlN. The short Lp of AlN results in a low CE(x) outside the junction, which makes its ηd even lower than that of SiC. The BN betavoltaic battery has the overall highest ideal efficiency, which means BN will be the optimal candidate among ultra-wide-bandgap semiconductors for betavoltaic batteries, although it is currently challenge to prepare stable homogeneous PN junctions [18].

3.3. Efficiency Prediction of the SiC Experimental Device

SiC betavoltaic batteries with a PIN junction were designed and fabricated to compare their performance with other semiconductor betavoltaic batteries, and their parameters were determined according to the energy deposition distribution in Figure 4b. The output performance of the SiC device under an isotope source or electron beam was predicted by the following method, described in Section 2.1.2, as shown in Table 3. The prediction efficiency under a 63Ni source is 17.54%, which is obviously lower than the ideal efficiency predicted in Table 1. This is due to the significantly reduced energy deposition in the SiC semiconductor compared to the ideal condition (as mentioned in Figure 4b), as shown in Figure 5d. The energy deposition of beta particles in different parts for the SiC device under a 63Ni source is shown in Figure 5d. Energy deposition in the 63Ni source and source substrate is up to 60%, which significantly reduces the energy deposition in the SiC semiconductor and makes the predicted efficiency much lower than the ideal value obtained in Table 1.

3.4. Measurement of the SiC Experimental Device

3.4.1. Measurement of Electron Beam

SiC betavoltaic battery devices were fabricated, and the I–V characteristics under an electron beam were analyzed. The obtained Isc, Voc and FF factors are in the range of 141.38 nA to 195.28 nA, 1.23 V to 1.54 V and 0.86 to 0.93, respectively. The best achieved device efficiency is 14.88%, according to Equation (1), with the corresponding I–V characteristics as shown in Figure 6a. This obtained SiC device efficiency is significantly higher than the available GaN device efficiency at 6.6% [9] and diamond device efficiency at 9.4% [8]. It is verified experimentally that SiC is the optimum wide-bandgap semiconductor for the betavoltaic battery.
The measured ideal factor of the best SiC device can be calculated as 1.36, according to the dark current analysis of I–V characteristics. A large ideal factor means a high proportion of recombination current in the junction region, which will lead to the decrease in barrier height and battery efficiency. The ideal factor was assumed to be 1 in the prediction situation, which results in the predicted Voc and η being higher than the measured results, and can be modified based on the measured ideal factor. The modified Voc and η predictions were 1.66 V and 16.32%, respectively, which were close to the measured results, as shown in Table 3. The defects in material growth and device surface treatment will increase the ideal factor, as it relates to the recombination current. Surface oxidation and the roughness of the electrode caused by the annealing process may be the reason why the ideal factor is higher than 1. More detailed surface treatment process exploration will be carried out in the future to further improve the battery performance.

3.4.2. Measurement of 63Ni Source

The I–V characteristics of SiC experimental devices with a 63Ni source were also analyzed. The obtained Isc, Voc and FF factors are in the range of 2.85 nA to 2.91 nA, 1.22 V to 1.41 V and 0.81 to 0.89, respectively. The best achieved device efficiency is 7.31%, which is lower than that under an electron beam, as shown in Table 3, with the corresponding I–V characteristics as shown in Figure 6b. This is due to the self-absorption effect of the isotope source, which results in lower energy deposition in a semiconductor than an electron beam.
The measured short-circuit current Isc of the SiC betavoltaic battery devices prepared in this work is slightly lower than the theoretically predicted value of 2.99 nA, with an overall difference of less than 3%. The measured value of the fill factor FF is similar to the theoretically predicted value of 0.91, whereas the measured value of the open-circuit voltage Voc has a higher difference from the theoretically predicted value of 1.89 V, resulting in a certain discrepancy in the energy conversion efficiency. According to Equation (6), the open-circuit voltage Voc increases with the decrease in the reverse saturation current I0. Substituting the measured reverse saturation current I0 value of 4.26 × 10−26 A into Equation (6), the corrected open-circuit voltage is obtained as 1.42 V, which is close to the measured result, and the theoretical device conversion efficiency is 12.83%, as shown in Table 3.
The output performance of the SiC betavoltaic batteries prepared in this paper is compared with the experimental results of other SiC betavoltaic batteries, and the results are shown in Table 4. The results show that the SiC betavoltaic batteries in this paper have a high fill factor and energy conversion efficiency.

4. Conclusions

In order to screen the optimum selection of wide-bandgap semiconductors, the device efficiencies of betavoltaic batteries with different semiconductors were analyzed by considering principle factors in the generation and collection processes of EHPs. For typical wide-bandgap semiconductors, the prediction efficiency of betavoltaic batteries with GaN was lower than that of SiC and diamond in this work. And the main reasons are summarized as follows: (1) The experimentally determined EHP creation energy deviated from the simple empirical assumption due to the complex band structure of wide-bandgap semiconductor materials. (2) The EHPs collection efficiency of GaN is relatively lower because of the short lifetime of its minority carrier as a direct-gap semiconductor, and shorter diffusion length resulting from more apparent defects in its heteroepitaxial growth process. (3) GaN has a higher atomic number and density than SiC and diamond, leading to higher backscattering and more energy loss of beta particles. The ideal efficiencies of betavoltaic batteries with ultra-wide-bandgap semiconductors were further predicted. Similarly, the ideal efficiencies of Ga2O3 and AlN betavoltaic batteries are even relatively lower than that of the SiC betavoltaic battery, while the BN betavoltaic battery has the potential to achieve the best performance. Considering both the theoretical efficiency prediction and the preparation maturity, SiC is the currently optimum choice for betavoltaic batteries, and its superiority was verified through experiments. The best obtained device efficiencies of SiC devices under an electron beam and a 63Ni source are 14.88% and 7.31%, respectively, in this work, and the energy conversion efficiency reaches 2.34~2.56%, an increase of about one time, which is obviously higher than that of betavoltaic batteries with other semiconductor materials reported in the literature. The device efficiency of SiC betavoltaic batteries can be further improved by optimizing the surface process to reduce the device ideal factor. The performances of betavoltaic batteries are worthy, as expected, along with the rapid development of preparation techniques related to ultra-wide-bandgap semiconductors. This work provides theoretical support for the development of betavoltaic batteries with ultra-wide-bandgap semiconductors.

Author Contributions

J.Z.: Data curation, Investigation, Software, Visualization, Writing—review and editing. K.L.: Conceptualization, Investigation, Methodology, Writing—original draft. Y.Y.: Formal analysis, Methodology. Y.G.: Visualization. Y.T.: Formal analysis, Methodology. Y.H.: Writing—review and editing. J.T.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the HFIPS Director’s Fund (Grant No. YZJJ202311-TS) and the Joint Funds of the National Natural Science Foundation of China (No. U2436201). And this research was supported by the Open Project of Shanxi Provincial Key Laboratory of Advanced Semiconductor Optoelectronic Devices and System Integration (2024SZKF03).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Zambelli, C.; Olivo, P.; Gaddi, R.; Schepens, C.; Smith, C. Characterization of a Mems-Based Embedded Non Volatile Memory Array for Extreme Environments. In Proceedings of the 3rd IEEE IMW, Monterey, CA, USA, 22–25 May 2011. [Google Scholar] [CrossRef]
  2. Nathan, M.; Peled, E.; Haronian, D. Micro Electrochemical Energy Storage Cells. United States Patent RE42273, 4 May 2011. Available online: https://www.freepatentsonline.com/6197450.html (accessed on 10 May 2022).
  3. Olsen, L.C. Review of Betavoltaic Energy Conversion. In Proceedings of the XII Space Photovoltaic Research and Technology Conference (SPRAT XII), Cleveland, OH, USA, 20–22 October 1992; Available online: https://ntrs.nasa.gov/citations/19940006935 (accessed on 15 May 2022).
  4. Theirrattanakul, S.; Prelas, M. A methodology for efficiency optimization of betavoltaic cell design using an isotropic planar source having an energy dependent beta particle distribution. Appl. Radiat. Isot. 2017, 127, 41–46. [Google Scholar] [CrossRef]
  5. Prelas, M.; Boraas, M.; Aguilar, F.D.; Seelig, J.-D.; Tchouaso, M.T.; Wisniewski, D. Nuclear Batteries and Radioisotopes; Springer: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
  6. Zhao, C.; Liu, A.; Bai, S.; Zhao, Y. Understanding Efficiency Differences of Betavoltaic Batteries Measured by Electron Gun Mimicked Source and Radioactive β Source. Appl. Phys. Lett. 2020, 117, 193901. [Google Scholar] [CrossRef]
  7. Grzanka, S.; Marona, L.; Winiewski, P.; Targowski, G.; Zaręba, B.; Wincel, K.; Klimasz, M.; Jacek, T.L.; Laskowski, K.P.; Kacperski, J.; et al. Highly Stable GaN-Based Betavoltaic Structures Grown on Different Dislocation Density Substrates. Solid-State Electron. 2020, 167, 107784. [Google Scholar] [CrossRef]
  8. Delfaure, C.; Pomorski, M.; Sanoit, J.D.; Bergonzo, P.; Saada, S. Single Crystal CVD Diamond Membranes for Betavoltaic Cells. Appl. Phys. Lett. 2016, 108, 246–247. [Google Scholar] [CrossRef]
  9. Heuser, T.; Braun, M.; McIntyre, P.; Senesky, D.G. Electron Beam Irradiation of Gallium Nitride-On-Silicon Betavoltaics Fabricated with a Triple Mesa Etch. J. Appl. Phys. 2021, 130, 174503. [Google Scholar] [CrossRef]
  10. Zhao, C.; Lei, L.; Liao, F.; Yuan, D.; Zhao, Y. Efficiency Prediction of Planar Betavoltaic Batteries Basing on Precise Modeling of Semiconductor Units. Appl. Phys. Lett. 2020, 117, 263901. [Google Scholar] [CrossRef]
  11. Thomas, C.; Portnoff, S.; Spencer, M.G. High Efficiency 4H-SiC Betavoltaic Power Sources Using Tritium Radioisotopes. Appl. Phys. Lett. 2016, 108, 8037–8058. [Google Scholar] [CrossRef]
  12. Bao, R.; Brand, P.J.; Chrisey, D.B. Betavoltaic Performance of Radiation-Hardened High-Efficiency Si Space Solar Cells. IEEE Trans. Electron Dev. 2012, 59, 1286–1294. [Google Scholar] [CrossRef]
  13. Khan, M.R.; Smith, J.R.; Tompkins, R.P.; Kelley, S.; Litz, M.; Russo, J.; Leathersich, J.; Shahedipour-Sandvik, F.; Jones, K.A.; Iliadis, A. Design and Characterization of GaN P-I-N Diodes for Betavoltaic Devices. Solid-State Electron. 2017, 136, 24–29. [Google Scholar] [CrossRef]
  14. Grushko, V.; Beliuskina, O.; Mamalis, A.; Lysakovskiy, V.; Mitskevich, E.; Kiriev, A.; Petrosyan, E.; Chaplynskyi, R.; Bezshyyko, O.; Lysenko, O. Energy conversion efficiency in betavoltaic cells based on the diamond Schottky diode with a thin drift layer. Appl. Radiat. Isot. 2020, 157, 109017. [Google Scholar] [CrossRef]
  15. Bormashov, V.S.; Troschiev, S.Y.; Tarelkin, S.A.; Volkov, A.P.; Teteruk, D.V.; Golovanov, A.V.; Kuznetsov, M.S.; Kornilov, N.V.; Terentiev, S.A.; Blank, V.D. High Power Density Nuclear Battery Prototype Based on Diamond Schottky Diodes. Diam. Relat.Mater. 2018, 84, 41–47. [Google Scholar] [CrossRef]
  16. Ishibe, T.; Tomeda, A.; Komatsubara, Y.; Kitaura, R.; Uenuma, M.; Uraoka, Y.; Yamashita, Y.; Nakamura, Y. Carrier and phonon transport control by domain engineering for high-performance transparent thin film thermoelectric generator. J. Appl. Phys. Lett. 2021, 118, 151601. [Google Scholar] [CrossRef]
  17. Li, F.; Zhang, S.; Xu, Z.; Wei, X.; Shrout, T.R. Critical Property in Relaxor-PbTiO (3) Single Crystals—Shear Piezoelectric Response. J. Adv. Funct. Mater. 2011, 21, 2118–2128. [Google Scholar] [CrossRef] [PubMed]
  18. Liao, M.; Shen, B.; Wang, Z. Ultra-Wide Bandgap Semiconductor Materials; Elsevier: Amsterdam, The Netherlands, 2019. [Google Scholar] [CrossRef]
  19. Tang, X.; Ding, D.; Liu, Y.; Chen, D. Optimization design and analysis of Si-63Ni betavoltaic battery. Sci. China Technol. Sci. 2012, 55, 990–996. [Google Scholar] [CrossRef]
  20. Chandrashekhar, M.; Thomas, C.I.; Spencer, M.G. Measurement of the Mean Electron-Hole Pair Ionization Energy in 4H-SiC. Appl. Phys. Lett. 2006, 89, 042113. [Google Scholar] [CrossRef]
  21. Honsberg, C.; Doolittle, W.A.; Allen, M.; Wang, C. GaN Betavoltaic Energy Converters. In Proceedings of the Thirty-First IEEE Photovoltaic Specialists Conference, Lake Buena Vista, FL, USA, 3–7 January 2005. [Google Scholar] [CrossRef]
  22. Anderson, B. Fundamentals of Semiconductor Devices; McGraw-Hill Higher Education: New York, NY, USA, 2006. [Google Scholar]
  23. Ye, Z. Determination of Average Energy Consumption by Electron-Hole Generation in Silicon by α Particles Atomic Energy. Sci. Technol. 1977, 11, 273. (In Chinese) [Google Scholar]
  24. Song, Z.; Zhao, C.; Liao, F.; Zhao, Y. Perovskite-Betavoltaic Cells: A Novel Application of Organic-Inorganic Hybrid Halide Perovskites. ACS Appl. Mater. Interfaces 2019, 11, 32969–32977. [Google Scholar] [CrossRef]
  25. Lioliou, G.; Barnett, A.M. Electron-hole pair creation and conversion efficiency in radioisotope microbatteries. Appl. Radiat. Isot. 2022, 180, 110042. [Google Scholar] [CrossRef]
  26. Butera, S.; Lioliou, G.; Krysa, A.B.; Barnett, A.M. Measurement of the Electron–Hole Pair Creation Energy in Al0.52in0.48P Using X-Ray Radiation. Nucl. Instrum. Methods Phys. Res. 2018, 879, 64–68. [Google Scholar] [CrossRef]
  27. Kobayashi, T. Average Energy to Form Electron-Hole Pairs in Gap Diodes with Alpha Particles. Appl. Phys. Lett. 1972, 21, 150–152. [Google Scholar] [CrossRef]
  28. Bloom, S.; Alig, R.C. Electron-Hole-Pair Creation Energies in Semiconductors. Phys. Rev. Lett. 1975, 35, 1522–1525. [Google Scholar] [CrossRef]
  29. Chaudhuri, S.K.; Zavalla, K.J.; Mandal, K.C. Experimental Determination of Electron-Hole Pair Creation Energy in 4H-SiC Epitaxial Layer: An Absolute Calibration Approach. Appl. Phys. Lett. 2013, 102, 031109. [Google Scholar] [CrossRef]
  30. Bertuccio, G.; Maiocchi, D. Electron-Hole Pair Generation Energy in Gallium Arsenide by X. and Γ Photons. J. Appl. Phys. 2002, 92, 1248–1255. [Google Scholar] [CrossRef]
  31. Lo Giudice, A.; Fizzotti, F.; Manfredotti, C.; Vittone, E.; Nava, F. Average Energy Dissipated by Mega-Electron-Volt Hydrogen and Helium Ions Per Electron-Hole Pair Generation in 4H-SiC. Appl. Phys. Lett. 2005, 87, 222105. [Google Scholar] [CrossRef]
  32. Yong, J.; Fan, X.; Rong, R.; Wu, J.; Bai, S.; Li, L. Properties Measurement of Silicon Carbide Detector. Nucl. Electron. Detect. Technol. 2012, 32, 1372–1375+1427. [Google Scholar]
  33. Wrbanek, J.; Wrbanek, S.; Fralick, G. Micro-Fabricated Solid-State Radiation Detectors for Active Personal Dosimetry. 2007. Available online: https://ntrs.nasa.gov/citations/20070014753 (accessed on 20 May 2022).
  34. Nava, F.; Bertuccio, G.; Cavallini, A.; Vittone, E. Silicon Carbide and its Use as a Radiation Detector Material. Meas. Sci. Technol. 2008, 19, 102001. [Google Scholar] [CrossRef]
  35. Krasnov, A.A.; Legotin, S.A. Advances in the Development of Betavoltaic Power Sources (a Review). Instrum. Exp. Tech. 2020, 63, 437–452. [Google Scholar] [CrossRef]
  36. Yakimov, E.B.; Polyakov, A.Y.; Shchemerov, I.V.; Smirnov, N.B.; Vasilev, A.A.; Vergeles, P.S.; Yakimov, E.E.; Chernykh, A.V.; Ren, F.; Pearton, S.J. Experimental Estimation of Electron-Hole Pair Creation Energy in β-Ga2O3. Appl. Phys. Lett. 2021, 118, 202106. [Google Scholar] [CrossRef]
  37. Pivrikas, A. Charge Transport Characteristics in Doped Organic Semiconductors Using Hall Effect. J. Electron. 2024, 13, 4223. [Google Scholar] [CrossRef]
  38. Morab, S.; Sundaram, M.M.; Pivrikas, A. Review on Charge Carrier Transport in Inorganic and Organic Semiconductors. Coatings 2023, 13, 1657. [Google Scholar] [CrossRef]
  39. Neamen, D.A. Semiconductor Physics and Devices; McGraw-Hill Higher Education: New York, NY, USA, 2011. [Google Scholar]
  40. Zhou, M.; Li, Z.; Fan, S.; Xu, Z.; Xiong, Z. Progress of the Growth of GaN Single Crystals. China Light. Light. 2017, 9, 8–17. [Google Scholar] [CrossRef]
  41. Wang, L.; Wang, F.; Liu, F.; Wang, J.; Lu, Z. Review of Diamond Semiconductor Devices. J. Wuhan Inst. Technol. 2020, 42, 518–525. [Google Scholar] [CrossRef]
  42. Dutta, M.; Koeck, F.; Hathwar, R.; Goodnick, S.M.; Nemanich, R.J.; Chowdhury, S. Demonstration of Diamond-Based Schottky P-I-N Diode with Blocking Voltage > 500 V. IEEE Electron Device Lett. 2016, 37, 1170–1173. [Google Scholar] [CrossRef]
  43. Liu, Z.; Zhao, D. Fabrication and Characterization of (100)-Oriented Single Crystal Diamond P-I-N Junction Ultraviolet Detector. Phys. Status Solidi. A Appl. Mater. Sci. 2020, 217, 2000207. [Google Scholar] [CrossRef]
  44. Maity, A.; Doan, T.C.; Li, J.; Lin, J.Y.; Jiang, H.X. Realization of Highly Efficient Hexagonal Boron Nitride Neutron Detectors. Appl. Phys. Lett. 2016, 109, 072101. [Google Scholar] [CrossRef]
  45. Tsao, J.Y.; Chowdhury, S.; Hollis, M.A.; Jena, D.; Johnson, N.M.; Jones, K.A.; Kaplar, R.J.; Rajan, S.; Van de Walle, C.G.; Bellotti, E.; et al. Ultrawide-Bandgap Semiconductors: Research Opportunities and Challenges. Adv. Electron. Mater. 2018, 4, 1600501. [Google Scholar] [CrossRef]
  46. Zhang, C.; Liao, F.; Liang, X.; Gong, H.; Liu, Q.; Li, L.; Qin, X.; Huang, X.; Huang, C. Electronic Transport Properties in Metal Doped Beta-Ga2O3: A First Principles Study. Phys. B 2019, 562, 124–130. [Google Scholar] [CrossRef]
  47. Lopatiuk-Tirpak, O.; Chernyak, L.; Borisov, B.A.; Kuryatkov, V.V.; Nikishin, S.A.; Gartsman, K. Electron Irradiation-Induced Increase of Minority Carrier Diffusion Length, Mobility, and Lifetime in Mg-doped AlN/AlGaN Short Period Superlattice. Appl. Phys. Lett. 2007, 91, 182103. [Google Scholar] [CrossRef]
  48. Cao, B.; Tian, Y.; Wen, H.F.; Guo, H.; Wu, X.; Li, L.; Zhang, Z.; Liu, L.; Zhu, Q.; Tang, J.; et al. Recent progress on fabrication, spectroscopy properties, and device applications in Sn-doped CdS micro-nano structures. J. Semicond. 2024, 45, 091101. [Google Scholar]
  49. Gao, P.; Li, X.-Y.; Yao, X.W. 4H-SiC Schottky Betavoltaic Micro Battery. Micronanoelectron. Technol. 2010, 47, 157–162. [Google Scholar]
  50. Chandrasekhar, M.; Thomas, C.I.; Li, H.; Spencer, M.G.; Lal, A. Demonstration of a 4H-SiC Betavoltaic Cell. Appl. Phys. Lett. 2006, 88, 33506. [Google Scholar] [CrossRef]
  51. Lu, M.; Wang, G.; Yao, C.S. Gallium Nitride for Nuclear Batteries. Adv. Mater. Res. 2012, 343–344, 56–61. [Google Scholar] [CrossRef]
  52. Cheng, Z.-J.; San, H.-S.; Chen, X.-Y.; Liu, B.; Feng, Z.-H. Demonstration of a High Open-Circuit Voltage GaN Betavoltaic Microbattery. Chin. Phys. Lett. 2011, 28, 78401. [Google Scholar] [CrossRef]
Figure 2. Physical diagram of the devices. (a) Series of physical samples of betavoltaic batteries prepared. (b) SiC-based betavoltaic battery. (c) 63Ni isotope source loading test.
Figure 2. Physical diagram of the devices. (a) Series of physical samples of betavoltaic batteries prepared. (b) SiC-based betavoltaic battery. (c) 63Ni isotope source loading test.
Nanomaterials 15 00635 g002
Figure 4. Effects of the energy deposition of beta particles. (a) The backscattering of different semiconductors under 1000 beta particles. (b) The energy deposition distribution of 63Ni decaying beta particles in each semiconductor.
Figure 4. Effects of the energy deposition of beta particles. (a) The backscattering of different semiconductors under 1000 beta particles. (b) The energy deposition distribution of 63Ni decaying beta particles in each semiconductor.
Nanomaterials 15 00635 g004
Figure 5. Predictions for different bandgap semiconductors. (a) The EHPs collection efficiency of different semiconductors. (b) The radiation-generated current distribution of different betavoltaic batteries. (c) The device efficiency of a betavoltaic battery using different semiconductors by considering different effects. (d) The energy deposition distribution of beta particles in SiC semiconductor under a 63Ni volume source. Inset: the energy deposition of β particles in different parts for the SiC betavoltaic battery device under a 63Ni source.
Figure 5. Predictions for different bandgap semiconductors. (a) The EHPs collection efficiency of different semiconductors. (b) The radiation-generated current distribution of different betavoltaic batteries. (c) The device efficiency of a betavoltaic battery using different semiconductors by considering different effects. (d) The energy deposition distribution of beta particles in SiC semiconductor under a 63Ni volume source. Inset: the energy deposition of β particles in different parts for the SiC betavoltaic battery device under a 63Ni source.
Nanomaterials 15 00635 g005
Figure 6. Measurement of the devices. (a) The I-V characteristics of SiC betavoltaic battery device under an electron beam at 20 kV acceleration voltage. (b) The I-V characteristics of the SiC betavoltaic battery device under a 63Ni isotope source.
Figure 6. Measurement of the devices. (a) The I-V characteristics of SiC betavoltaic battery device under an electron beam at 20 kV acceleration voltage. (b) The I-V characteristics of the SiC betavoltaic battery device under a 63Ni isotope source.
Nanomaterials 15 00635 g006
Table 1. The ideal device efficiency of betavoltaic battery using typical wide- and ultra-wide-bandgap semiconductors.
Table 1. The ideal device efficiency of betavoltaic battery using typical wide- and ultra-wide-bandgap semiconductors.
SemiconductorSiCGaNDiamondGa2O3AlNBN
Jsc/nA·cm−2110.1695.1872.1851.1843.6264.67
Voc/V2.242.404.393.815.315.75
ηd22.17%20.69%29.29%17.98%21.57%34.65%
Table 2. The characteristics of representative ultra-wide-bandgap semiconductors [36,44,45,46,47,48] **.
Table 2. The characteristics of representative ultra-wide-bandgap semiconductors [36,44,45,46,47,48] **.
SemiconductorGa2O3AlNBN
Bandgap/eV4.96.136.4
Density/g·cm−36.13.263.49
Electron–hole pair creation energy/eV15.614.4 *15.0 *
Relative dielectric constant108.67.1
Intrinsic carrier concentration/cm−32.81 × 10−233.82 × 10−368.83 × 10−39
Minority carrier diffusion length/μm0.3250.157.6
Carrier mobility/cm2·V−1yggs−12014500
Bandgap typedirect directindirect
* It is estimated based on the relation (2.15Eg+1.23), which is re-extracted from experimental data from existing materials. ** The SiC, Ga2O3 and BN mentioned in this paper are 4H-SiC, β-Ga2O3 and h-BN respectively.
Table 3. Prediction and measurement results of SiC betavoltaic battery devices.
Table 3. Prediction and measurement results of SiC betavoltaic battery devices.
SourceConditionIsc/nAVoc/VFFηdn
Electron beamoriginal prediction197.712.250.9423.17%1
prediction with experimental n197.711.660.9216.32%1.36
measurement195.281.540.8914.88%1.36
63Nioriginal prediction2.991.890.9117.54%1
prediction with experimental I02.991.420.8812.83%1.51
measurement2.911.410.817.31%1.51
Table 4. Output performance of different betavoltaic batteries.
Table 4. Output performance of different betavoltaic batteries.
SemiconductorVoc/VFFηReferences
SiC0.500.301.3%[49]
0.720.511.2%[50]
GaN0.140.141.60%[51]
1.620.551.13%[52]
Diamond1.020.701.25%[15]
This work (SiC)1.22–1.410.81–0.892.34–2.56%--
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, J.; Lv, K.; Yin, Y.; Gao, Y.; Tian, Y.; Han, Y.; Tang, J. Optimal Selection and Experimental Verification of Wide-Bandgap Semiconductor for Betavoltaic Battery. Nanomaterials 2025, 15, 635. https://doi.org/10.3390/nano15090635

AMA Style

Zhang J, Lv K, Yin Y, Gao Y, Tian Y, Han Y, Tang J. Optimal Selection and Experimental Verification of Wide-Bandgap Semiconductor for Betavoltaic Battery. Nanomaterials. 2025; 15(9):635. https://doi.org/10.3390/nano15090635

Chicago/Turabian Style

Zhang, Jiachen, Kunlun Lv, Yuan Yin, Yuqian Gao, Ye Tian, Yuncheng Han, and Jun Tang. 2025. "Optimal Selection and Experimental Verification of Wide-Bandgap Semiconductor for Betavoltaic Battery" Nanomaterials 15, no. 9: 635. https://doi.org/10.3390/nano15090635

APA Style

Zhang, J., Lv, K., Yin, Y., Gao, Y., Tian, Y., Han, Y., & Tang, J. (2025). Optimal Selection and Experimental Verification of Wide-Bandgap Semiconductor for Betavoltaic Battery. Nanomaterials, 15(9), 635. https://doi.org/10.3390/nano15090635

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

Article Metrics

Back to TopTop