Next Article in Journal
Damage Characteristics of Silicon Solar Cells Induced by Nanosecond Pulsed Laser
Previous Article in Journal
Sub-60 fs, 1300 nm Laser Pulses Generation from Soliton Self-Frequency Shift Pumped by Femtosecond Yb-Doped Fiber Laser
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles and Device-Level Investigation of β-AgGaO2 Ferroelectric Semiconductors for Photovoltaic Applications

1
Department of Physics, School of Science, Guangdong University of Petrochemical Technology, Maoming 525000, China
2
Shenzhen Zhenyang Precision Technology Co., Ltd., Shenzhen 518103, China
*
Authors to whom correspondence should be addressed.
Photonics 2025, 12(8), 803; https://doi.org/10.3390/photonics12080803
Submission received: 15 July 2025 / Revised: 1 August 2025 / Accepted: 5 August 2025 / Published: 11 August 2025

Abstract

Ferroelectric semiconductors, with their inherent spontaneous polarization, present a promising approach for efficient charge separation, making them attractive for photovoltaic applications. The potential of β-AgGaO2, a polar ternary oxide with an orthorhombic Pna21 structure, as a light-absorbing material is evaluated. First-principles computational analysis reveals that β-AgGaO2 possesses an indirect bandgap of 2.1 eV and exhibits pronounced absorption within the visible spectral range. Optical simulations suggest that a 300 nm thick absorber layer could theoretically achieve a power conversion efficiency (PCE) of 20%. Device-level simulations using SCAPS-1D evaluate the influence of hole and electron transport layers on solar cell performance. Among the tested hole transport materials, Cu2FeSnS4 (CFTS) achieves the highest PCE of 14%, attributed to its optimized valence band alignment and reduced recombination losses. In contrast, no significant improvements were observed with the electron transport layers tested. These findings indicate the potential of β-AgGaO2 as a ferroelectric photovoltaic absorber and emphasize the importance of band alignment and interface engineering for optimizing device performance.

1. Introduction

Ferroelectric semiconductors, due to their inherent spontaneous polarization properties, can facilitate the separation of photogenerated charge carriers without the need for an external electric field, making them significant in optoelectronic research [1]. In recent years, with the increasing demand for high-performance materials in optoelectronic devices, ferroelectric semiconductors have gradually become a prominent research direction [2]. For example, β-phase ABO2 ternary oxides derived from the zinc-blende ZnO structure, with their Pna21 polar space group, show promising ferroelectric semiconductor properties [3]. First-principles calculations indicate that β-CuGaO2 possesses both a high spontaneous polarization of 83.80 μC/cm2 and a direct, narrow bandgap, making it a promising candidate for ferroelectric photovoltaic applications [4,5]. By employing the spectroscopic limited maximum efficiency (SLME) approach, our calculation indicates that the photovoltaic efficiency of β-CuGaO2 approaches 30%, which is comparable to that of traditional direct bandgap photovoltaic materials like CdTe [6]. Furthermore, as a material possessing both ferroelectric and semiconducting properties, β-CuGaO2 shows extensive application potential in fields such as energy harvesting and wastewater environmental remediation [7].
Building on the excellent properties of β-CuGaO2, Wang et al. have employed unit co-doping strategies to identify four ferroelectric semiconductor materials, including β-CuInO2, β-AgGaO2, β-CuGaO2, and β-AuInO2, which exhibit outstanding photocatalytic performance and broad spectral absorption capabilities [8]. Furthermore, Guo et al. introduced cations such as P3+, As3+, Sb3+, and Bi3+ into β-phase ABO2 ternary oxide ferroelectric semiconductors, thereby synthesizing 44 different types of β-AIBIIIO2 photocatalysts with dual polarization characteristics, which significantly improve electron–hole pair separation efficiency [9]. Computational results indicate that the solar-to-hydrogen efficiency of β-TlBiO2 can reach 17.2% [9]. In terms of experimental synthesis, β-AgGaO2 has been successfully prepared, showing excellent visible light absorption and potential ferroelectric properties, indicating its important application value in photocatalytic hydrogen production and photovoltaic fields [10]. First-principles calculations and experimental measurements show that the bandgap of β-AgGaO2 is approximately 2.1 eV, and its band edge positions can span the redox potential of water, thereby facilitating the photocatalytic water splitting process [8,11,12,13]. In contrast, BiFeO3, a commonly used ferroelectric material, can be effectively tuned through La doping to adjust the bandgap, which typically ranges from 2.2 eV to 2.7 eV, and is widely used in photovoltaic and photocatalytic research [14,15]. Although both β-AgGaO2 and BiFeO3 exhibit similar ferroelectric properties and bandgaps, BiFeO3 often requires an additional electron transport layer to enhance photoelectric conversion efficiency and improve photovoltaic performance [16,17]. In contrast, β-AgGaO2, with its excellent visible light absorption, has emerged as a potential material for solar cell structures. To further optimize its device performance, β-AgGaO2 can benefit from the incorporation of electron or hole transport layers, which facilitate efficient charge carrier separation and transport, thereby significantly enhancing overall photovoltaic efficiency and expanding its potential applications in the field of photovoltaics. In this study, density functional theory (DFT) calculations combined with numerical simulations were utilized to investigate the optical and electronic properties of β-AgGaO2, as well as to assess its theoretical power conversion efficiency and photovoltaic device performance. The computational results demonstrate that the integration of a hole transport layer (HTL) into the β-AgGaO2-based architecture can enhance photovoltaic performance by promoting more effective charge carrier separation and transport.

2. Density Functional Theory (DFT) Computational Approach

All first-principles calculations within the framework of density functional theory (DFT) were performed using the plane-wave pseudopotential approach as implemented in the VASP package [18,19]. The exchange-correlation interactions were treated with the generalized gradient approximation (GGA) in the form proposed by Perdew, Burke, and Ernzerhof (PBE) [20]. Equilibrium structural parameters, including lattice constants and atomic positions, were obtained by minimizing the total energy with the conjugate gradient algorithm. A kinetic energy cutoff of 550 eV was employed for the plane-wave basis set expansion, and k-point sampling grids were generated with the aid of the pre-processing tool VASPKIT [21]. The Gamma scheme was chosen for k-point mesh of 6 × 5 × 6. The relaxation of electronic degrees of freedom was terminated when the total energy and band structure energy changes between two electron steps were less than 10−6 eV. The convergence criterion for ionic steps was set to 0.01 eV Å−1. The calculated lattice parameters of β-AgGaO2 are a = 5.64 Å, b = 7.18 Å, and c = 5.59 Å as shown in Figure 1. The CIF file of the crystal structure is provided in the Supporting Information.
These values are in excellent agreement with the experimentally reported lattice parameters of a = 5.56 Å, b = 7.14 Å, and c = 5.46 Å [10]. The Modified Becke–Johnson (MBJ) exchange potential, which corrects the underestimation of the bandgap in conventional DFT-PBE calculations, was used to compute the electronic structure and optical absorption spectra [22]. The parameter C in the MBJ functional was set to 1.5, and the computed bandgap showed good agreement with previous computational results and experimental data. Ab initio molecular dynamics (AIMD) simulations based on first-principles calculations were also conducted to evaluate the thermodynamic stability of AgGaO2, as shown in Figure S1 of the Supporting Information.

3. Computational Analysis of Electronic and Optical Properties

3.1. Electronic Properties

The optical and electronic characteristics of materials are fundamentally determined by their band structures. The band gap and density of states (DOS) for β-AgGaO2 are depicted in Figure 2.
The bandgap calculation employs the Modified Becke–Johnson (MBJ) exchange potential, which generally yields more accurate results. The calculated bandgap is in good agreement with experimental data and HSE06 computational results (approximately 2.18 eV) [10,11,12]. As illustrated in Figure 2, the valence band maximum is mainly constituted by O-2p orbitals, whereas the conduction band minimum is chiefly attributed to Ag-5s states. The effective masses of charge carriers (holes and electrons) were calculated using a second-order derivative method, as outlined in Equation (1).
m * = d 2 E d k 2 1
Here, E denotes the band edge energy as a function of the wave vector k. The computed effective masses for electrons and holes are 0.17 m0 and −3.45 m0, respectively, which are consistent with previously reported values of 0.1 m0 and −2.9 m0 in the literature [8]. The effective density of states in the conduction band (Nc) and at the valence band maximum for holes (Nv) were determined using the following expressions, with the temperature T set to 300 K.
N c = 2 m e * k B T 2 π 2 3 2
N v = 2 2 π m h * k B T h 2 3 2
The calculated results for Nc and Nv are 2.4 × 1018 and 1.5 × 1020 cm−3, respectively. The band edge energies of β-AgGaO2 referenced to the vacuum level were estimated using the following empirical relationships [23]:
E V B M = χ + 0.5 E g
E C B M = χ 0.5 E g
χ = χ A g 1 + χ G a 1 + χ O 2 1 1 + 1 + 2
χ represents the average electronegativity of β-AgGaO2, and Eg denotes the band gap. The absolute electronegativity values for χ A g , χ G a , and χ O are 4.44 eV, 3.2 eV, and 7.54 eV, respectively. The calculated values of the valence band maximum (EVBM) and conduction band minimum (ECBM) are 4.3 eV and 6.4 eV, respectively.
The carrier mobility of β-AgGaO2 was computed using the AMSET code, which employs the momentum relaxation time approximation (MRTA) method [24]. All required input parameters, including the elastic constants (Table 1), dielectric constants (Table 2), phonon frequencies (Table 2), piezoelectric tensor (Table 3), and deformation potentials (Table 4), were derived from our DFT calculations. The resulting carrier mobilities are summarized in Table 5.
The findings demonstrate that, under moderate impurity concentrations, the electron mobility can achieve 100 cm2/V·s, while the hole mobility is 0.3 cm2/V·s. These values substantially exceed those of the intrinsic ferroelectric material BiFeO3, whose carrier mobility typically ranges from 1.6 to 1.8 cm2/V·s, with the electron mobility of β-AgGaO2 being particularly pronounced [25]. In BiFeO3, significant enhancement in electron mobility generally necessitates doping with ions such as La, which can increase electron mobility up to 80 cm2/V·s [16,17]. The mobility of charge carriers in semiconductors is influenced by various scattering mechanisms, primarily including ionized impurity (IPM), polar optical phonon (POP), piezoelectric (PIE), and acoustic deformation potential scattering (ADP), as shown in Figure 3. Consistent with other polar semiconductors, the carrier mobility in β-AgGaO2 is predominantly governed by polar optical phonon scattering [26]. The relatively low hole mobility in β-AgGaO2 is primarily attributed to strong scattering effects from polar optical phonons and piezoelectric interactions. This limited mobility significantly influences device performance and suggests the need for further optimization through material engineering or interface modification.

3.2. Optical Properties

There exists a close interrelation between the band gap, dielectric constant, and optical properties. Within the framework of the random phase approximation (RPA), the optical properties are derived from the real (ε1) and imaginary (ε2) parts of the dielectric function. From ε1 and ε2, the refractive index (n) and extinction coefficient (k) can be determined as follows [27]:
n E = 1 2 ε 1 + ε 1 2 E + ε 2 2 E  
k E = 1 2 ε 1 + ε 1 2 E + ε 2 2 E
Consequently, the reflectance (R) and absorption coefficient (α) are calculated from n and k using the following expressions [27]:
R E = n 1 2 + K 2 n + 1 2 + K 2
α E = 4 π k h c / E
As illustrated in Figure 4a, β-AgGaO2 demonstrates strong optical absorption in the visible region, with absorption coefficients on the order of 105 cm−1, which highlights its significant potential for applications in solar energy harvesting.
The photoelectric conversion efficiency is evaluated using the following physical model [28]:
P C E = 0 λ m a x W λ 1 R λ 1 e α λ d C λ d λ 0 W λ d λ
In this model, λ denotes the wavelength, and λmax is the maximum absorption wavelength corresponding to the band gap (Eg), given by:
λ m a x = h c E g
The conversion coefficient C(λ) signifies the fraction of excitation energy that results in the creation of an electron–hole pair via the minimum band gap, described as:
C λ = λ E g h c
The calculated absorption coefficient α and reflectance R are depicted in Figure 4a and Figure 4b, respectively. The photoelectric conversion efficiencies, as shown in Figure 4c, reach 13% and 20% for film thicknesses of 100 nm and 300 nm, respectively, with the system approaching absorption saturation (21%) at 300 nm. In our device simulation, the thickness of β-AgGaO2 is chosen as 300 nm, a value that is commonly used for BiFeO3 thin films in experiments [16,29,30]. The physical model described in Equation (11) calculates the theoretical maximum photovoltaic conversion efficiency, which serves as an initial reference for evaluating the photovoltaic potential of the material. To obtain a more accurate representation of the actual device performance, we construct a detailed solar cell device model in the following sections for further calculation of the solar cell characteristics of β-AgGaO2.

4. Photovoltaic Performance Analysis of β-AgGaO2 Devices with and Without Transport Layers

The β-AgGaO2 solar cell structures were designed and analyzed utilizing SCAPS-1D, a one-dimensional simulation software for solar cells developed by the Department of Electronics and Information Systems at Ghent University, Belgium [31]. SCAPS-1D enables the simulation of homojunction, heterojunction, multi-junction, as well as Schottky junction solar cells through numerical solutions of the Poisson equation [32,33].
d 2 φ x d x 2 = q ε n x p x N D + x + N A x p t x + n t x
where φ ( x ) is the electrostatic potential, q is the elementary charge, ε is the permittivity, n(x) and p(x) are the electron and hole densities, N D + x and N A x are the ionized donor and acceptor concentrations, and n t ( x ) and   p t x represent trapped charge carrier densities. SCAPS also solves the continuity equations for electrons and holes:
n t = 1 q J n x + G R
p t = 1 q J p x + G R
where Jn and Jp are the current densities of electrons and holes, G is the generation rate, and R is the recombination rate. This allows for a detailed understanding of charge transport and recombination processes across multilayer device architectures. The input parameters used in the simulation are presented in Table 6, Table 7 and Table 8. Table 6 primarily lists the parameters of the electrodes, absorption layer, and electron transport layers (ETLs). Table 7 focuses on the parameters of the hole transport layers (HTLs), while Table 8 provides the parameters of the interfacial layers. In this study, the parameters for the hole transport layers (HTLs) and electron transport layers (ETLs) were primarily selected based on values reported in the literature (Table 6, Table 7 and Table 8), which is a widely adopted approach in solar cell device modeling. However, recent simulations of CsSnI2Br-based perovskite solar cells have gone a step further by actively tuning the properties of charge transport layers to enhance photovoltaic performance [34]. Given that both charge transport layers’ interface and intrinsic defect concentrations in β-AgGaO2 can also influence device efficiency, we have included corresponding simulation results in the Supporting Information (Tables S1–S4) to provide a more comprehensive assessment of their impact. Material properties for the electrodes, HTLs, and ETLs were obtained from the literature, whereas the parameters for β-AgGaO2 were mainly derived from our DFT calculations.
The work function assigned to the gold (Au) electrode is 5.1 eV. The photovoltaic performance was evaluated across three device configurations: (i) a structure without transport layers (ITO/β-AgGaO2/Au), (ii) a structure incorporating a hole transport layer (ITO/β-AgGaO2/HTL/Au), and (iii) a structure incorporating an electron transport layer (ITO/ETL/β-AgGaO2/Au).

4.1. Photovoltaic Performance Without Transport Layers (ITO/β-AgGaO2/Au)

The performance of the solar cell without hole transport layer (HTL) and electron transport layer (ETL) was analyzed using current–voltage (J–V) characteristics and band diagrams (Figure 5).
The power conversion efficiency (PCE) of the device is 1.80%, with a short-circuit current density (Jsc) of 6.01 mA/cm2 and an open-circuit voltage (Voc) of 0.39 V. This indicates low charge extraction efficiency and significant recombination losses, resulting in reduced overall device performance. This performance is similar to that of ferroelectric La-doped BiFeO3 (FTO/BLFO/Au) photovoltaic devices, which have a PCE of 1.5%, making it difficult to achieve high-performance photovoltaic devices [17].
The band diagram further explains this phenomenon in Figure 5b: the conduction band (Ec) alignment between AgGaO2 and ITO facilitates effective electron extraction and transport. However, the poor valence band (Ev) alignment suppresses hole extraction, leading to enhanced interface recombination and hindered charge transport. Since the electron affinity was estimated using a semi-empirical formula, some degree of uncertainty may be present. To assess its influence, we varied the electron affinity from 4.0 eV to 4.6 eV in increments of 0.1 eV and recalculated the corresponding power conversion efficiencies (PCEs). As shown in Table 9, a decrease in electron affinity tends to improve band alignment. The results suggest that even small variations in electron affinity can cause noticeable changes in PCE, indicating the model’s sensitivity to this parameter. Accordingly, the use of experimental techniques such as UPS or XPS in future studies may help obtain more accurate band alignment data and improve the reliability of simulation results. The multi-point computational analysis presented here may serve as a useful reference for subsequent investigations.
When the electron affinity is 4.0 eV, the PCE increases from 1.80% to 3.50%. Although the PCE improved, the interface losses remain significant due to the lack of HTL and ETL, particularly in hole extraction, ultimately leading to a reduction in efficiency. Therefore, the introduction of HTL and ETL plays a crucial role in optimizing band alignment, reducing recombination losses, and improving carrier transport efficiency, significantly enhancing the overall performance of the device.

4.2. The Impact of the Hole Transport Layer (ITO/β-AgGaO2/HTL/Au)

To evaluate the effect of the HTL on photovoltaic performance, materials such as CFTS, CuI, P3HT, PSS, and CuSCN were tested. J–V characteristics in Figure 6 show that the choice of HTL significantly impacts the device performance.
CFTS showed the best performance, with a Jsc of 17.05 mA/cm2 and a PCE of 8.10%, demonstrating excellent hole extraction and transport capabilities. According to previous reports, CFTS exhibits excellent thermal stability and could be compatible with low-temperature solution processing, making it a promising candidate for integration with oxide semiconductors in photovoltaic or photocatalytic applications [46]. However, due to the current lack of experimental studies on the direct interface formation between CFTS and β-AgGaO2, further investigations are needed to verify the interface quality and assess the feasibility of practical device fabrication. Although CuSCN exhibited the highest Voc (1.07 V), its Jsc was low (5.88 mA/cm2), resulting in a PCE of 5.39%, indicating that the interface energy barrier limits charge transport. P3HT and CuI showed lower PCEs (around 4.6%) due to lower Jsc and Voc values. PSS exhibited slightly better performance with a Jsc of 8.85 mA/cm2 and a PCE of 5.13%. The band diagram analysis in Figure 7 reveals that band alignment plays a critical role in performance.
The valence band of CFTS is well-matched with β-AgGaO2, facilitating easy hole transport and low interface recombination losses, and forming an effective barrier in the conduction band to suppress electron leakage, ensuring high carrier selectivity. Although CuI has a higher Voc, the high valence band creates an energy barrier, leading to delayed charge extraction and increased recombination. The valence band misalignment in P3HT and PSS also limits hole extraction and collection efficiency. Although CuSCN’s valence band theoretically favors hole extraction, its low Jsc and PCE of 5.39% may be due to poor film quality, interface defects, or insufficient hole mobility. Both the interface defect density and the intrinsic defect density of β-AgGaO2 were found to degrade the photovoltaic performance of the ITO/β-AgGaO2/HTL/Au device structure. The corresponding simulation results are provided in Tables S1 and S2 of the Supporting Information. As shown in Figure 8 and Table 9, the results indicate that as the electron affinity decreases, the band alignment improves, which benefits hole transport, reduces interface recombination losses, and ultimately leads to higher photovoltaic efficiency.
In particular, if interface defects are ignored, the PCE could be further enhanced, with CFTS as the hole transport layer potentially reaching a PCE of 14.47%. Therefore, band alignment is crucial for improving performance, but material quality, defect control, and carrier mobility cannot be overlooked. CFTS shows the best overall performance, while CuI and CuSCN, despite ideal valence band positions, are limited by large energy barriers or material issues. P3HT and PSS are constrained by valence band misalignment and interface recombination.

4.3. The Impact of the Electron Transport Layer (ITO/ETL/β-AgGaO2/Au)

Although the ETL is theoretically expected to improve performance, the devices using ETL materials such as C60, WS2, PCBM, ZnO, and TiO2 did not demonstrate better performance than those without transport layers in Figure 9.
The Jsc for ZnO and TiO2 was around 6.00 mA/cm2, and the PCE was 1.80%, comparable to the performance of devices without transport layers, indicating that interface defects, poor carrier collection, or increased recombination paths still exist. C60 and WS2 performed worse, with Jsc values of 2.11 mA/cm2 and 1.60 mA/cm2, and PCEs of 0.58% and 0.43%, respectively. PCBM performed slightly better, with a Jsc of 2.58 mA/cm2 and a PCE of 0.72%.
Band diagram analysis in Figure 10 shows that C60 and WS2 exhibit significant conduction band misalignment with β-AgGaO2, creating large energy barriers that limit electron extraction.
Moreover, their valence band positions may cause hole accumulation and increased recombination, explaining their poor performance. PCBM shows better conduction band alignment but still presents an energy barrier that limits effective electron transport. ZnO and TiO2 have good conduction band alignment but did not show significant performance improvements, possibly due to interface defects, carrier traps, or poor ETL film quality. Additionally, valence band misalignment could lead to hole accumulation and recombination, further limiting performance improvements. The interface defect density and intrinsic defect density of β-AgGaO2 further reduce the photovoltaic performance of the ITO/ETL/β-AgGaO2/Au device structure. The corresponding simulation results are provided in Tables S3 and S4 of the Supporting Information. Based on the calculations in Figure 11 and Table 9, it is evident that the introduction of ETL does not significantly improve the device’s photovoltaic efficiency under varying electron affinity conditions.
This suggests that future strategies to enhance ETL performance should focus on interface passivation, optimizing film quality, or using alternative materials with better alignment and superior transport properties.

5. Conclusions

In this work, we combine first-principles calculations with device-level simulations to systematically investigate the structural, electronic, optical, and photovoltaic properties of the ferroelectric polar ternary oxide β-AgGaO2. Our density functional theory (DFT) calculations reveal that β-AgGaO2 has an indirect band gap of 2.1 eV, strong absorption in the visible region, and favorable electron mobility. Polar optical phonon scattering is identified as the primary factor limiting carrier mobility. Nevertheless, optical modeling of thin-film photovoltaic devices predicts a theoretical power conversion efficiency (PCE) of up to 21%. Device simulations using SCAPS-1D indicate that incorporating a hole transport layer (HTL) substantially enhances device performance. Among the HTLs evaluated, CFTS was found to be the most effective, as it facilitates improved carrier extraction and suppresses interface recombination. Conversely, the addition of an electron transport layer did not improve performance, suggesting that further interface optimization or the exploration of alternative materials is necessary. Overall, our findings not only demonstrate the promising photovoltaic potential of β-AgGaO2 but also underscore the critical role of energy band alignment and interface engineering in developing high-efficiency ferroelectric solar cells.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/photonics12080803/s1.

Author Contributions

Methodology, W.-J.H.; Writing—original draft, W.-J.H.; Data curation, X.-Y.Z. and Y.-L.H.; Validation, X.-T.Z.; Programming, H.-K.X.; Software, X.-F.X.; Visualization, X.-F.X.; Conceptualization, Y.-D.C.; Investigation, X.-Y.C.; Supervision, L.-T.N. and B.D.; Funding acquisition, B.D.; Writing—review and editing, X.-Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Guangdong Provincial Key Areas Special Project for Regular Higher Education Institutions (Grant No. 2023ZDZX3014), the Shenzhen Science and Technology Program (Grant No. JSGG20220831110002004), and the Maoming Science and Technology Program (Grant No. 2023023). The APC was funded by the above-mentioned projects.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

Author Bing Dai was employed by the company Shenzhen Zhenyang Precision Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Butler, K.T.; Frost, J.M.; Walsh, A. Ferroelectric materials for solar energy conversion: Photoferroics revisited. Energy Environ. Sci. 2015, 8, 838–848. [Google Scholar] [CrossRef]
  2. Zhao, X.; Song, K.; Huang, H.; Han, W.; Yang, Y. Ferroelectric materials for solar energy scavenging and photodetectors. Adv. Opt. Mater. 2022, 10, 2101741. [Google Scholar] [CrossRef]
  3. Suzuki, I.; Omata, T. Multinary wurtzite-type oxide semiconductors: Present status and perspectives. Semicond. Sci. Technol. 2016, 32, 013007. [Google Scholar] [CrossRef]
  4. Omata, T.; Nagatani, H.; Suzuki, I.; Kita, M.; Yanagi, H.; Ohashi, N. Wurtzite CuGaO2: A new direct and narrow band gap oxide semiconductor applicable as a solar cell absorber. J. Am. Chem. Soc. 2014, 136, 3378–3381. [Google Scholar] [CrossRef]
  5. Song, S.; Kim, D.; Jang, H.M.; Yeo, B.C.; Han, S.S.; Kim, C.S.; Scott, J.F. β-CuGaO2 as a strong candidate material for efficient ferroelectric photovoltaics. Chem. Mater. 2017, 29, 7596–7603. [Google Scholar] [CrossRef]
  6. Luo, G.; Bian, Y.; Wu, R.; Lai, G.; Xu, X.; Zhang, W.; Chen, X. First principles study of the electronic structure and photovoltaic properties of β-CuGaO2 with MBJ+ U approach. J. Semicond. 2020, 41, 102102. [Google Scholar] [CrossRef]
  7. Yao, M.C.; Wu, X.J.; Xu, L.L.; Meng, F.Z.; Yang, Q.; Meng, J.; Liu, X.J. β-CuGaO2: A ferroelectric semiconductor with narrow band gap as degradation catalyst for wastewater environmental remediation. Rare Met. 2022, 41, 972–981. [Google Scholar] [CrossRef]
  8. Wang, R.; Xu, L.; Liu, Q.; Shi, Q.; Liu, X. Search for simple β-AIMIIIO2-type intrinsic ferroelectric semiconductors with simultaneous robust built-in electric field and full-spectrum absorption for superior photocatalysts. J. Mater. Chem. A 2023, 11, 5233–5244. [Google Scholar] [CrossRef]
  9. Guo, X.; Xu, L.; Dai, J.; Wang, Y.; Shi, Q.; Liu, X. Dual Polarization Strategy for Boosting Electron–Hole Separation toward Overall Water Splitting within Ferroelectric β-AIBIIIO2 (BIII = P3+, As3+, Sb3+, and Bi3+ for Lone Pairs). Inorg. Chem. 2024, 63, 10031–10041. [Google Scholar] [CrossRef] [PubMed]
  10. Nagatani, H.; Suzuki, I.; Kita, M.; Tanaka, M.; Katsuya, Y.; Sakata, O.; Omata, T. Structure of β-AgGaO2; ternary I–III–VI2 oxide semiconductor with a wurtzite-derived structure. J. Solid State Chem. 2015, 222, 66–70. [Google Scholar] [CrossRef]
  11. Guo, L.; Zhu, S.; Zhang, S.; Feng, W. Elastic, electronic, optical, and spectroscopic properties of β-AgMO2 (M = Al and Ga): First-principles calculations. Comput. Mater. Sci. 2014, 92, 92–101. [Google Scholar] [CrossRef]
  12. Suzuki, I.; Iguchi, Y.; Sato, C.; Yanagi, H.; Ohashi, N.; Omata, T. Comprehensive first-principles study of AgGaO2 and CuGaO2 polymorphs. J. Ceram. Soc. Jpn. 2019, 127, 339–347. [Google Scholar] [CrossRef]
  13. Dadsetani, M.; Nejatipour, R. Calculation of Electronic and Optical Properties of AgGaO2 Polymorphs Using Many-Body Approaches. J. Electron. Mater. 2018, 47, 1059–1070. [Google Scholar] [CrossRef]
  14. Biswas, P.P.; Pal, S.; Subramanian, V.; Murugavel, P. Large photovoltaic response in rare-earth doped BiFeO3 polycrystalline thin films near morphotropic phase boundary composition. Appl. Phys. Lett. 2019, 114, 173901. [Google Scholar] [CrossRef]
  15. Mariano, M.A.S.; Mendez-González, Y.; Silva, A.C.; Monte, A.F.G.; Lima, E.C.; Guo, R.; Bhalla, A.S.; de los Santos Guerra, J. Physical characterization of BiFeO3-based thin films with enhanced properties for photovoltaic applications. J. Am. Ceram. Soc. 2022, 105, 6965–6975. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Sun, H.; Yang, C.; Su, H.; Liu, X. Modulating photovoltaic conversion efficiency of BiFeO3-based ferroelectric films by the introduction of electron transport layers. ACS Appl. Energy Mater. 2019, 2, 5540–5546. [Google Scholar] [CrossRef]
  17. Raj, A.; Kumar, M.; Kumar, A.; Singh, K.; Sharma, S.; Singh, R.C.; Pawar, M.; Yahya, M.Z.A.; Anshul, A. Comparative analysis of ‘La’modified BiFeO3-based perovskite solar cell devices for high conversion efficiency. Ceram. Int. 2023, 49, 1317–1327. [Google Scholar] [CrossRef]
  18. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  19. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  20. Perdew, J.P. Generalized gradient approximation made simple. Phys. Rev. Lett. 1997, 77, 3868. [Google Scholar] [CrossRef]
  21. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  22. Tran, F.; Blaha, P. Accurate band gaps of semiconductors and insulators with a semilocal exchange-correlation potential. Phys. Rev. Lett. 2009, 102, 226401. [Google Scholar] [CrossRef]
  23. Xu, Y.; Schoonen, M.A.A. The absolute energy positions of conduction and valence bands of selected semiconducting minerals. Am. Mineral. 2000, 85, 543–556. [Google Scholar] [CrossRef]
  24. Ganose, A.M.; Park, J.; Faghaninia, A.; Woods-Robinson, R.; Persson, K.A.; Jain, A. Efficient calculation of carrier scattering rates from first principles. Nat. Commun. 2021, 12, 2222. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, J.C.; Yeh, C.H.; Chen, Y.T.; Liao, S.C.; Huang, R.; Liu, H.J.; Hung, C.C.; Chen, S.H.; Wu, S.L.; Lai, C.H.; et al. Conduction control at ferroic domain walls via external stimuli. Nanoscale 2014, 6, 10524–10529. [Google Scholar] [CrossRef]
  26. Irvine, L.A.D.; Walker, A.B.; Wolf, M.J. Quantifying polaronic effects on the scattering and mobility of charge carriers in lead halide perovskites. Phys. Rev. B 2021, 103, L220305. [Google Scholar] [CrossRef]
  27. Fox, M. Optical Properties of Solids; Oxford University Press: New York, NY, USA, 2002. [Google Scholar]
  28. Shi, G.; Kioupakis, E. Electronic and optical properties of nanoporous silicon for solar-cell applications. ACS Photonics 2015, 2, 208–215. [Google Scholar] [CrossRef]
  29. Fruth, V.; Popa, M.; Calderon-Moreno, J.M.; Anghel, E.M.; Berger, D.; Gartner, M.; Anastasescu, M.; Osiceanu, P.; Zaharescu, M. Chemical solution deposition and characterization of BiFeO3 thin films. J. Eur. Ceram. Soc. 2007, 27, 4417–4420. [Google Scholar] [CrossRef]
  30. Moniz, S.J.A.; Quesada-Cabrera, R.; Blackman, C.S.; Tang, J.; Southern, P.; Weaver, P.M.; Carmalt, C.J. A simple, low-cost CVD route to thin films of BiFeO3 for efficient water photo-oxidation. J. Mater. Chem. A 2014, 2, 2922–2927. [Google Scholar] [CrossRef]
  31. Burgelman, M.; Nollet, P.; Degrave, S. Modelling polycrystalline semiconductor solar cells. Thin Solid Film. 2000, 361, 527–532. [Google Scholar] [CrossRef]
  32. Karthick, S.; Velumani, S.; Bouclé, J. Experimental and SCAPS simulated formamidinium perovskite solar cells: A comparison of device performance. Sol. Energy 2020, 205, 349–357. [Google Scholar] [CrossRef]
  33. Srivastava, P.; Rai, S.; Lohia, P.; Dwivedi, D.K.; Qasem, H.; Umar, A.; Akbar, S.; Algadi, H.; Baskoutas, S. Theoretical study of perovskite solar cell for enhancement of device performance using SCAPS-1D. Phys. Scr. 2022, 97, 125004. [Google Scholar] [CrossRef]
  34. Khan, A.H.H.; Wang, C.-H. Tailoring transporters for all-inorganic tin-based perovskite solar cells with efficiency exceeding 22% using SCAPS-1D simulator. Inorg. Chem. Commun. 2025, 179, 114707. [Google Scholar] [CrossRef]
  35. Abdelaziz, S.; Zekry, A.; Shaker, A.; Abouelatta-Ebrahim, M. Investigating the performance of formamidinium tin-based perovskite solar cell by SCAPS device simulation. Opt. Mater. 2020, 101, 109738. [Google Scholar] [CrossRef]
  36. Anwar, F.; Afrin, S.; Satter, S.S.; Mahbub, R.; Ullah, S.M. Simulation and performance study of nanowire CdS/CdTe solar cell. Int. J. Renew. Energy Res. 2017, 7, 885–893. [Google Scholar]
  37. Raoui, Y.; Ez-Zahraouy, H.; Ahmad, S.; Kazim, S.; Tahiri, N.; Omar, E.B. Performance analysis of MAPbI3 based perovskite solar cells employing diverse charge selective contacts: Simulation study. Sol. Energy 2019, 193, 948–955. [Google Scholar] [CrossRef]
  38. Chabri, I.; Oubelkacem, A.; Benhouria, Y. Numerical development of lead-free Cs2TiI6-based perovskite solar cell via SCAPS-1D. E3S Web Conf. 2022, 336, 00050. [Google Scholar] [CrossRef]
  39. Noorasid, N.S.; Arith, F.; Firhat, A.Y.; Mustafa, A.N.; Shah, A.S.M. SCAPS numerical analysis of solid-state dye-sensitized solar cell utilizing copper (I) iodide as hole transport layer. Environ. Energy Nat. Resour. 2022, 26, 1–10. [Google Scholar] [CrossRef]
  40. Gan, Y.; Bi, X.; Liu, Y.; Qin, B.; Li, Q.; Jiang, Q.; Mo, P. Numerical investigation energy conversion performance of tin-based perovskite solar cells using cell capacitance simulator. Energies 2020, 13, 5907. [Google Scholar] [CrossRef]
  41. Sobayel, K.; Akhtaruzzaman, M.; Rahman, K.S.; Ferdaous, M.T.; Al-Mutairi, Z.A.; Alharbi, H.F.; Alharthi, N.H.; Karim, M.R.; Hasmady, S.; Amin, N. A comprehensive defect study of tungsten disulfide (WS2) as electron transport layer in perovskite solar cells by numerical simulation. Results Phys. 2019, 12, 1097–1103. [Google Scholar] [CrossRef]
  42. Touafek, N.; Mahamdi, R.; Dridi, C. Boosting the performance of planar inverted perovskite solar cells employing graphene oxide as HTL. Dig. J. Nanomater. Biostruct. 2021, 16, 705–712. [Google Scholar] [CrossRef]
  43. Alipour, H.; Ghadimi, A. Optimization of lead-free perovskite solar cells in normal-structure with WO3 and water-free PEDOT: PSS composite for hole transport layer by SCAPS-1D simulation. Opt. Mater. 2021, 120, 111432. [Google Scholar] [CrossRef]
  44. Khatun, M.M.; Sunny, A.; Al Ahmed, S.R. Numerical investigation on performance improvement of WS2 thin-film solar cell with copper iodide as hole transport layer. Sol. Energy 2021, 224, 956–965. [Google Scholar] [CrossRef]
  45. Khattak, Y.H.; Baig, F.; Toura, H.; Beg, S.; Soucase, B.M. CZTSe kesterite as an alternative hole transport layer for MASnI3 perovskite solar cells. J. Electron. Mater. 2019, 48, 5723–5733. [Google Scholar] [CrossRef]
  46. Vanalakar, S.A.; Patil, P.S.; Kim, J.H. Recent advances in synthesis of Cu2FeSnS4 materials for solar cell applications: A review. Sol. Energy Mater. Sol. Cells 2018, 182, 204–219. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the lattice structure of β-AgGaO2.
Figure 1. Schematic illustration of the lattice structure of β-AgGaO2.
Photonics 12 00803 g001
Figure 2. Calculated (a) band structure and (b) density of states (DOS) of β-AgGaO2.
Figure 2. Calculated (a) band structure and (b) density of states (DOS) of β-AgGaO2.
Photonics 12 00803 g002
Figure 3. Calculated (a) electron mobility and (b) hole mobility of β-AgGaO2 as a function of impurity concentration.
Figure 3. Calculated (a) electron mobility and (b) hole mobility of β-AgGaO2 as a function of impurity concentration.
Photonics 12 00803 g003
Figure 4. Calculated (a) optical absorption spectra, (b) reflectance, and (c) maximum power conversion efficiency (PCE) as a function of thickness.
Figure 4. Calculated (a) optical absorption spectra, (b) reflectance, and (c) maximum power conversion efficiency (PCE) as a function of thickness.
Photonics 12 00803 g004
Figure 5. (a) Current density–voltage (J–V) curves and (b) band alignment diagram of the ITO/β-AgGaO2/Au device.
Figure 5. (a) Current density–voltage (J–V) curves and (b) band alignment diagram of the ITO/β-AgGaO2/Au device.
Photonics 12 00803 g005
Figure 6. Current–voltage (J–V) characteristics of ITO/β-AgGaO2/HTL/Au devices incorporating different hole transport layers (HTLs): (a) CFTS; (b) CuI; (c) P3HT; (d) PSS; (e) CuSCN.
Figure 6. Current–voltage (J–V) characteristics of ITO/β-AgGaO2/HTL/Au devices incorporating different hole transport layers (HTLs): (a) CFTS; (b) CuI; (c) P3HT; (d) PSS; (e) CuSCN.
Photonics 12 00803 g006
Figure 7. Energy band alignment diagrams of ITO/β-AgGaO2/HTL/Au devices employing different hole transport layers (HTLs): (a) CFTS; (b) CuI; (c) P3HT; (d) PSS; (e) CuSCN.
Figure 7. Energy band alignment diagrams of ITO/β-AgGaO2/HTL/Au devices employing different hole transport layers (HTLs): (a) CFTS; (b) CuI; (c) P3HT; (d) PSS; (e) CuSCN.
Photonics 12 00803 g007
Figure 8. Power conversion efficiency (PCE) of ITO/β-AgGaO2/HTL/Au devices with varying electron affinities and HTLs.
Figure 8. Power conversion efficiency (PCE) of ITO/β-AgGaO2/HTL/Au devices with varying electron affinities and HTLs.
Photonics 12 00803 g008
Figure 9. Current–voltage (J–V) characteristics of ITO/ETL/β-AgGaO2/Au devices incorporating various electron transport layers (ETLs): (a) C60; (b) WS2; (c) PCBM; (d) ZnO; (e) TiO2.
Figure 9. Current–voltage (J–V) characteristics of ITO/ETL/β-AgGaO2/Au devices incorporating various electron transport layers (ETLs): (a) C60; (b) WS2; (c) PCBM; (d) ZnO; (e) TiO2.
Photonics 12 00803 g009
Figure 10. Energy band alignment diagrams of ITO/ETL/β-AgGaO2/Au devices employing different electron transport layers (ETLs): (a) C60; (b) WS2; (c) PCBM; (d) ZnO; (e) TiO2.
Figure 10. Energy band alignment diagrams of ITO/ETL/β-AgGaO2/Au devices employing different electron transport layers (ETLs): (a) C60; (b) WS2; (c) PCBM; (d) ZnO; (e) TiO2.
Photonics 12 00803 g010
Figure 11. PCE of ITO/ETL/β-AgGaO2/Au devices with varying electron affinities and ETLs.
Figure 11. PCE of ITO/ETL/β-AgGaO2/Au devices with varying electron affinities and ETLs.
Photonics 12 00803 g011
Table 1. The calculated elastic constants of β-AgGaO2 (GPa).
Table 1. The calculated elastic constants of β-AgGaO2 (GPa).
C11C12C13C22C23C33C44C55C66
103.280.287.9157.390.6131.116.915.25.1
Table 2. The calculated static (ɛ0), high frequency (ɛ) dielectric constant, and effective phonon frequency (ω).
Table 2. The calculated static (ɛ0), high frequency (ɛ) dielectric constant, and effective phonon frequency (ω).
Materialɛ0ɛω (THz)
xyzxyz
AgGaO214.0610.8113.106.046.116.3010.41
Table 3. The calculated piezoelectric tensor (C m−2).
Table 3. The calculated piezoelectric tensor (C m−2).
XXYYZZXYYZZX
AgGaO2x0.000.000.000.000.00−0.37
y0.000.000.000.00−0.750.00
z−0.52−0.861.290.000.000.00
Table 4. The calculated deformation potentials for the lower conduction bands and upper valence bands.
Table 4. The calculated deformation potentials for the lower conduction bands and upper valence bands.
Material DXXDYYDZZ
AgGaO2VBM(eV)DXX1.230.390.76
DYY0.392.590.17
DZZ0.760.170.1
CBM(eV)DXX0.050.310.00
DYY0.310.710.03
DZZ0.000.030.54
Table 5. The calculated electrons (μe) and holes (μh) mobility for AgGaO2  ( c m 2 V 1 S 1 ) .
Table 5. The calculated electrons (μe) and holes (μh) mobility for AgGaO2  ( c m 2 V 1 S 1 ) .
Material ND = 1015
/cm−3
ND = 1016
/cm−3
ND = 1017
/cm−3
ND = 1018
/cm−3
ND = 1019
/cm−3
ND = 1020
/cm−3
AgGaO2μe(n)107105.2100.390.782.044.4
μh(p)0.40.30.30.30.30.3
Table 6. Input Parameters of ITO, ETLs, and Absorption layer.
Table 6. Input Parameters of ITO, ETLs, and Absorption layer.
Material PropertyITOZnOPCBMC60WS2AgGaO2
Thickness “t” (nm)500505050100300
Bandgap “Eg” (eV)3.53.321.71.82.1
Electron affinity (eV)443.93.93.954.3
Relative permittivity “Er”993.94.213.613.1
C.B”Nc” (cm−3)2.2 × 10183.7 × 10182.5 × 10218.0 × 10191 × 10182.4 × 1018
V.B”Nd” (cm−3)1.8 × 10191.8 × 10192.5 × 10218.0 × 10192.4 × 10191.5 × 1020
Electron thermal
Velocity “Ve” (cm/s)
1 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
Hole thermal velocity “Vh” (cm/s)1 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
Electron mobility
“μn” (cm2/V·s)
201000.28.0 × 10−2100100
Hole mobility
“μh” (cm2/V·s)
10250.23.5 × 10−31000.3
Donor density
“ND” (cm−3)
1 × 10211 × 10182.93 × 10171 × 10171 × 10181 × 1019
Acceptor density
“NA” (cm−3)
000000
Defect density
“Nt” (cm−3)
1 × 10151 × 10151 × 10151 × 10151 × 10151 × 1015
References[35,36][37,38][36,39][40][35,41]
Table 7. Input Optimization Parameters for the Studied HTLs.
Table 7. Input Optimization Parameters for the Studied HTLs.
Material PropertyCuSCNP3HTPEDOT:PSSCuICFTS
Thickness “t” (nm)505050100100
Bandgap “Eg” (eV)3.61.71.63.11.3
Electron affinity (eV)1.73.53.42.13.3
Relative permittivity “Er”1033.96.59
C.B”Nc” (cm−3)2.2 × 10192 × 10212.2 × 10182.8 × 10192.2 × 1018
V.B”Nd” (cm−3)1.8 × 10182 × 10211.8 × 10191 × 10191.8 × 1019
Electron thermal
Velocity “Ve” (cm/s)
1 × 1071 × 1071 × 1071 × 1071 × 107
Hole thermal velocity “Vh” (cm/s)1 × 1071 × 1071 × 1071 × 1071 × 107
Electron mobility
“μn” (cm2/V·s)
1001.8 × 10−34.5 × 10−210021.98
Hole mobility
“μh”(cm2/V·s)
251.86 × 10−24.5 × 10−243.921.98
Donor density
“ND” (cm−3)
00000
Acceptor density
“NA” (cm−3)
1 × 10181 × 10181 × 10181 × 10181 × 1018
Defect density
“Nt” (cm−3)
1 × 10151 × 10151 × 10151 × 10151 × 1015
References[37][37][42,43][44][45]
Table 8. Specifications of Defect Layers’ Interface.
Table 8. Specifications of Defect Layers’ Interface.
InterfaceDefect
Type
Capture
Cross-Section:
Electrons
/Holes (cm 2)
Energetic
Distribution
Reference
for Defect
Energy Level
Total
Density
(cm−2)
ETL/AgGaO2neutral1.0 × 10−17singleAbove the
highest EV
1.0 × 1010
1.0 × 10−18
AgGaO2/HTLneutral1.0 × 10−18singleAbove the
highest EV
1.0 × 1010
1.0 × 10−19
Table 9. The calculated power conversion efficiency (PCE, %) of photovoltaic devices with different electron transport layers (ETLs), hole transport layers (HTLs), and no transport layer as a function of electron affinity. Values in parentheses represent PCE without considering interface recombination.
Table 9. The calculated power conversion efficiency (PCE, %) of photovoltaic devices with different electron transport layers (ETLs), hole transport layers (HTLs), and no transport layer as a function of electron affinity. Values in parentheses represent PCE without considering interface recombination.
Electron Affinity(eV)ETL/HTL-FreeTiO2WS2ZnOC60PCBMCuIPSSCuSCNP3HTCFTS
4.03.503.36
(3.38)
0.87
(0.87)
3.48
(3.48)
1.16
(1.16)
1.30
(1.30)
6.30
(8.05)
7.63
(9.08)
7.09
(8.49)
6.61
(6.98)
8.90
(14.47)
4.12.932.92
(2.92)
0.72
(0.72))
2.92
(2.93)
0.97
(0.97)
1.13
(1.13)
5.73
(7.50)
6.79
(8.31)
6.52
(7.93)
5.95
(6.68)
8.29
(13.07)
4.22.362.36
(2.36)
0.57
(0.57)
2.36
(2.36)
0.77
(0.77)
0.93
(0.93)
5.17
(6.95)
5.96
(7.54)
5.95
(7.38)
5.27
(6.31)
8.19
(11.51)
4.31.801.80
(1.80)
0.43
(0.43)
1.80
(1.80)
0.58
(0.58)
0.72
(0.72)
4.61
(6.41)
5.13
(6.78)
5.39
(7.91)
4.60
(5.96)
8.10
(9.97)
4.41.261.26
(1.26)
0.28
(0.00)
1.26
(1.26)
0.00
(0.37)
0.50
(0.50)
4.04
(5.86)
4.35
(6.03)
0.00
(0.00)
3.93
(5.44)
7.80
(8.45)
4.50.730.73
(0.73)
0.00
(0.00)
0.73
(0.73)
0.00
(0.00)
0.00
(0.28)
3.48
(5.32)
4.01
(0.00)
0.00
(0.00)
3.27
(4.89)
6.88
(6.97)
4.60.270.27
(0.27)
0.00
(0.00)
0.15
(0.15)
0.00
(0.00)
0.00
(0.00)
2.97
(0.00)
3.84
(0.00)
0.00
(0.00)
2.65
(0.00)
0.00
(0.00)
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

Hu, W.-J.; Zhang, X.-Y.; Zhu, X.-T.; Hu, Y.-L.; Xu, H.-K.; Xu, X.-F.; Che, Y.-D.; Chen, X.-Y.; Niu, L.-T.; Dai, B. First-Principles and Device-Level Investigation of β-AgGaO2 Ferroelectric Semiconductors for Photovoltaic Applications. Photonics 2025, 12, 803. https://doi.org/10.3390/photonics12080803

AMA Style

Hu W-J, Zhang X-Y, Zhu X-T, Hu Y-L, Xu H-K, Xu X-F, Che Y-D, Chen X-Y, Niu L-T, Dai B. First-Principles and Device-Level Investigation of β-AgGaO2 Ferroelectric Semiconductors for Photovoltaic Applications. Photonics. 2025; 12(8):803. https://doi.org/10.3390/photonics12080803

Chicago/Turabian Style

Hu, Wen-Jie, Xin-Yu Zhang, Xiao-Tong Zhu, Yan-Li Hu, Hua-Kai Xu, Xiang-Fu Xu, You-Da Che, Xing-Yuan Chen, Li-Ting Niu, and Bing Dai. 2025. "First-Principles and Device-Level Investigation of β-AgGaO2 Ferroelectric Semiconductors for Photovoltaic Applications" Photonics 12, no. 8: 803. https://doi.org/10.3390/photonics12080803

APA Style

Hu, W.-J., Zhang, X.-Y., Zhu, X.-T., Hu, Y.-L., Xu, H.-K., Xu, X.-F., Che, Y.-D., Chen, X.-Y., Niu, L.-T., & Dai, B. (2025). First-Principles and Device-Level Investigation of β-AgGaO2 Ferroelectric Semiconductors for Photovoltaic Applications. Photonics, 12(8), 803. https://doi.org/10.3390/photonics12080803

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