Next Article in Journal
Dual-Arm Visuo-Haptic Optical Tweezers for Bimanual Cooperative Micromanipulation of Nonspherical Objects
Next Article in Special Issue
Development and Modelling of Gallium Nitride Based Lateral Schottky Barrier Diodes with Anode Recesses for mmWave and THz Applications
Previous Article in Journal
Electrical Modeling and Characterization of Graphene-Based On-Chip Spiral Inductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Study of a Solar Cell to Achieve the Highest InGaN Power Conversion Efficiency for the Whole In-Content Range

by
Rubén Martínez-Revuelta
1,
Horacio I. Solís-Cisneros
1,
Raúl Trejo-Hernández
2,
Madaín Pérez-Patricio
1,
Martha L. Paniagua-Chávez
1,
Rubén Grajales-Coutiño
1,
Jorge L. Camas-Anzueto
1 and
Carlos A. Hernández-Gutiérrez
1,*
1
Optomechatronics Group, Tecnológico Nacional de México Campus Tuxtla Gutiérrez, Carretera Panamericana Km 1080, Tuxtla Gutiérrez 29050, Mexico
2
Programa de Nanociencias y Nanotecnología, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional, Av. Instituto Politécnico Nacional 2508, Ciudad de México 07360, Mexico
*
Author to whom correspondence should be addressed.
Micromachines 2022, 13(11), 1828; https://0-doi-org.brum.beds.ac.uk/10.3390/mi13111828
Submission received: 4 September 2022 / Revised: 18 October 2022 / Accepted: 24 October 2022 / Published: 26 October 2022
(This article belongs to the Special Issue III-V/III-N Materials and Devices)

Abstract

:
A solar cell structure with a graded bandgap absorber layer based on InGaN has been proposed to overcome early predicted efficiency. Technological issues such as carrier concentration in the p- and n-type are based on the data available in the literature. The influence of carrier concentration-dependent mobility on the absorber layer has been studied, obtaining considerable improvements in efficiency and photocurrent density. Efficiency over the tandem solar cell theoretical limit has been reached. A current density of 52.95 mA/cm2, with an efficiency of over 85%, is determined for a PiN structure with an InGaN step-graded bandgap absorption layer and 65.44% of power conversion efficiency for the same structure considering piezoelectric polarization of fully-strained layers and interfaces with electron and hole surface recombination velocities of 10−3 cm/s.

1. Introduction

Silicon is the most used material in the solar cell market due to its abundance and mature technology. For silicon, theoretical photovoltaic conversion efficiency (PCE) limit is 29.1% [1]. On the other hand, for tandem InGaN solar cells, the theoretical PCE is around 60% [2,3]. InGaN-based solar cells have the potential to increase their conversion efficiencies because this alloy is suitable to absorb most of the solar spectrum ranging from 365 nm to 1771 nm (3.40–0.7 eV) [4,5,6,7]. Furthermore, it is essential to mention that due to the high InGaN absorption coefficient being around ~105 cm−1 [3,4,7], it is possible to achieve thinner absorbing layers. Thus, nanometric films with thicknesses near 100 nm are required to absorb 90% of the incident light [7]. Despite these unique properties, InxGa1−xN solar cells have not shown any significant improvement in conversion efficiency, mainly due to the difficulty of incorporating high Indium concentrations. It has been reported that if the Indium (In) molar fraction exceeds 0.20 or the Indium-Gallium-Nitride (InGaN) layer thickness increases, phase separation may occur [6,8,9]. One of the biggest problems in obtaining high-efficiency solar cells is phase separation.
Furthermore, it is reported that phase separation can occur due to doping concentration, mean atomic radius between InN and GaN, growth temperature, pressure, and magnetic field, resulting in metallic phases [10,11,12,13]. The lack of substrates leads to dislocations, the miscibility gap between InN and GaN, migration of Indium to the surface, and adatoms [5,13]. Stress in the epilayer due to both growth and substrate mismatch can change the critical growth temperature. This change implies that the growth temperature drops below the critical level of miscibility, which can cause phase separation [14,15,16]. The increase of Indium causes a narrower bandgap and reduces the crystal quality due to the mismatched lattices [16].
It is essential to mention that a challenge to consider is the p-type film’s growth and the difficulty of controlling the conductivity as the In content increases within the alloy [17]. Moreover, there is also a significant challenge in managing background donors associated with crystal defects and impurity incorporations [18]. In this work, InGaN-based solar cells will be numerically calculated by SILVACO Atlas, considering the current state-of-the-art Indium content and carrier concentration. The analysis is focused on the graded absorbed layer with linear- and step-graded Indium content, from 3.4 eV to 0.7 eV.

2. Materials and Methods

The solar cell structure is simulated using the SILVACO Atlas as the finite element numerical calculation software. The materials parameters in their hexagonal structure are modified according to the data in Table 1. The simulation considers Shockley Read Hall and Auger recombination models and the concentration-dependent carrier mobility.
For the electronic affinity calculation dependent on the In content, the band alignment parameter (0.58 for GaN/InN) specifies the fraction of the difference between the forbidden bands in the discontinuity of both the conduction band and the valence band occurs [27].
Each layer’s carrier concentrations and thickness are crucial in a PiN junction solar cell to achieve adequate depletion region extension. Due to the high absorption coefficient available for InGaN material, a fully depleted junction is ideal for achieving high photo collection. However, the depletion region is reduced due to the residual electron background and difficulty achieving high free hole concentration. The photocurrent mechanisms combine diffusion currents in the quasineutral regions and the depletion region collection.
The effect of Indium content on free electron concentration is shown in Figure 1a, where it is important to remark on the n-type nature of InGaN. When In molar fraction increases within the InGaN-based alloy, the background electron concentration tends to increase. Thus, the presented structure considered an electron concentration for intentional and non-intentional n-type doped layers with a free electron concentration ≥ 1019 cm−3 [3,16,17,18,19,20,21,22,23,24,25,26]. The high electron concentration can be attributed to the low temperature required to grow InGaN layers with high In content, provoking native defects, such as nitrogen vacancies [28,29,30]. It is observed that after a certain In molar fraction, the concentration of electrons tends to saturate [31,32,33,34]. For this reason, the absorber-layer and the n-layer electron concentrations are set at n = 1 × 1019 cm−3 and n = 2 × 1019 cm−3, respectively. However, as carrier concentration increases, carrier mobility is reduced due to scattering by impurity ionization [35].
On the other hand, due to the n-type nature described above, the p-InGaN layers deal with high electron concentration, provoking a self-compensation effect that results in hole concentration reduction [36].
This electron abundance in p-type InGaN- layers is mainly due to the lower temperature needed to grow In-rich InGaN layers instead of the required temperature to grow Ga-rich ones. For this reason, the p-type layer has been considered GaN-based, and the graded absorber layer has been set from GaN to InN across the entire Indium range within the ternary alloy. These considerations are consistent with the state-of-the-art, as illustrated in Figure 1a,b. Moreover, different works with low In content tend to present hole concentration in the mid-1018 cm−3 layer [3,18,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51], despite the acceptor activation energy decrease while In mole fraction increases [39,52]. At the same time, highly doped p-GaN layers have been reported [53]. Here is essential to mention that, despite obtaining a high electron concentration being relatively more straightforward than p-type doping, it is necessary to control the background electron concentration due to the n-nature of the alloy.
Due to the terrestrial applications considered in this work, the appropriate standard solar spectrum is the AM1.5 which corresponds to an Air Mass equal to 1.5 with a zenith angle of around 48.19°. For the solar spectra AM1.5, the absorption range is considered from 365 nm to 1771 nm (Figure 2), from the maximum available bandgap for x = 0.0, where x is the In molar fraction (GaN at 3.4 eV) up to x = 1, which is the minimum bandgap at 0.7 eV available, corresponding to InN (see Table 1). Vegard’s law, expressed in Equation (1), has been employed to determine material parameters for each layer
EInGaN = x∙EInN + (1 − x)∙EGaN − b∙x∙(1 − x)
where EInGaN is the ternary alloy bandgap, EInN is the Indium nitride bandgap, EGaN is the gallium nitride bandgap, and b is the bowing factor. In this work, b = 1.348 is considered according to ref. [54]. Since graded bandgap has been modeled, it has been simulated that a high-to-low energy ramp in the absorption layer is the best approach to improving the collection of photogenerated carriers [55,56]. The highly strained materials are obtained by grading the absorption InGaN-based layer, reducing the dislocations and phase separation [57].
The photocurrent has been calculated using a model for the graded bandgap absorber layer [58]. Moreover, the absorption constant is calculated from the absorption coefficient [59] and linearizing Equation (2) to average the constant absorption from different curves at different In content. Figure 3 presents the slopes (A) for the InGaN absorption coefficients with different In content in the alloy.
α = A (E − Eg)1/2
The gradient of the graded bandgap ideal photocurrent most decays to zero to estimate the adequate thickness necessary to achieve high efficiency. As shown in the inset of Figure 4, each incident photon generates an electron-hole pair contributing to the photogeneration.
From Figure 4, it is remarkable mentioning that ideally, 320 mA/cm2 can be generated with only 2 µm of InGaN-based absorber layer without considering optical issues. Therefore, InGaN is one of the most promising candidates for achieving high-efficiency solar cells. Thus, 2 µm is the minimal thickness to maximize the short circuit current. The simulated structure is shown in Figure 5, where the p-GaN layer is 200 nm thick, and the thickness of the absorber layer varies from 500 nm up to 2µm to guarantee photocurrent saturation. The n-InN layer is a 150 nm thick layer with a free electron concentration of 2 × 1020 cm−3.
The In molar fraction varies from the maximum available in the InxGa1−xN bandgap at x = 0, where x is the In molar fraction. Therefore, the maximum bandgap in the alloys corresponds to GaN at 3.4 eV and the minimum bandgap at 0.7 eV, corresponding to InN (Table 1). Since temperature control is crucial for InGaN-based layers, even in epitaxial systems, it is essential to consider realistic conditions. Thus, the absorber layer is regarded as a step-graded approach in the whole In molar fraction range, with 20 steps increasing 5% of Indium in each step. A bandgap diagram under dark conditions is shown in Figure 5, where a step-graded absorber layer is exhibited.
For simulation, the Shockley–Read–Hall, Auger, and Radiative recombination has been considered. A refractive index file has been created from information in the state-of-the-art, and the absorption constant has been obtained by calculating the slope in the absorption coefficient curves from different works cited in the document. For the GaN layer, optical wavelength-dependent properties have also been included in the simulation. Strains across the layers have been calculated and regulated with a polar scale varied from 0 to 1. However, the structure’s absence of strong piezoelectric polarization has been used as an approximation of the zincblende approach. Interface recombination has been considered with electron and hole surface recombination velocities at 10−3 cm/s and defects in the order of 1010 cm−2. The thermionic and Danielsson models have been considered in the code.

3. Results and Discussion

Figure 6a shows the normalized photon absorption rate. Since the solar spectra get deeper in the structure, the absorption rate diminishes due to the reduction of the depletion region and recombination in the quasineutral regions. Moreover, non-ideal models considered are SRH carrier lifetime, Auger, and radiative recombination coefficients, which indicates that not all photons absorbed generate an electron contributing to the photocurrent. Despite these considerations, the generation rate is maintained across the absorber layer as a benefit of the gradual band gap absorbing long wavelengths as the depth of the absorber layer increases. Moreover, the carrier concentration in the absorber layer affects the internal electric field, which yields the depletion region. Figure 6b shows the effect of the free electron concentration in the absorber layer on the electrostatic field distribution.
By employing the data reported in Table 2, recombination parameters such as SRH carrier lifetime (τn and τp), auger coefficient, and optical recombination (direct or radiative) coefficient are considered. Since electron concentration in the absorber layer has been simulated at two different values, recombination rates are analyzed in Figure 7. Remarkably, n = 1 × 1017 cm−3 exhibits a reduction of SRH recombination rates in the junction between the absorber layer and the n-contact layer.
The obtained spectral response has been compared against other solar cells in the state-of-the-art in Figure 8. Table 3 presents the VOC, JSC, FF, and efficiency for three different carrier concentrations in the absorber layer for 0.5 µm, 1 µm, 1.5 µm, 2 µm, and 2.5 µm. Remarkable results are colored in red. On the other hand, these results are graphically represented in Figure 9, where the J-V curve has been added.
In order to become more realistic, a simulation has been performed, adding a scale factor for the piezoelectric polarization due to the lattice mismatch between the different layers and across the entire structure. Figure 10 presents the effect of the relaxation in the structure without considering defects.
In summary, as the difference between hole and electron concentrations in the p-n junctions is reduced, the electrostatic field near this junction becomes more intense, reducing the depletion region. Furthermore, for lower free electron concentration, the depletion region is extended, and the electron mobility is higher, improving the diffusion length and, as a result, the current density, open-circuit voltage, and consequently, FF and PCE. So, according to Figure 6a,b, it is observed that low background in the absorber layer is required to achieve high collection efficiency.
Reducing SRH recombination rates in the junction between the absorber layer and the n-contact layer improves external quantum efficiency (EQE) from 50 to over 80%. Figure 8 presents a comparison of the electron concentration effect on EQE. The spectral response of the solar cell structure from 0.100 µm to 2.000 µm, and the apparent improvement is mainly due to the broadening of the depletion region. While the absorption is improved deeper in the structure, the absorption of photons with longer wavelengths contributes to the photogenerated current.
Moreover, the effect of piezoelectric polarization on power conversion efficiency results in a 15% drop due to the strained structure. This piezoelectric polarization is induced by the large lattice mismatch between the InN, InGaN, and GaN layers. The carrier surface recombination velocities have been considered for the last simulation to determine the efficiency reduction corresponding to the semiconductor–semiconductor interfaces with defects in the order of 1010 cm−2 and carrier surface recombination velocities of 10−3 cm/s.

4. Conclusions

In this work, an Indium-graded InGaN-based solar cell is modeled considering the actual device issues, such as the difficulty of incorporating high Indium content without phase separation and optical properties of the absorber layer and electrodes. Moreover, the high residual electron background concentration and the problem of achieving high free hole concentration at Indium-rich InGaN layers have been considered. A step-graded In content profile is proposed to accomplish the whole range of In compositions focused on epitaxial growth of the steps for significant temperature and thickness control. With these considerations, the present approach results in phase separation mitigation and simultaneously maximizes the short circuit current. The doping issues in the absorber layer helped illustrate that the higher efficiency with this structure is higher than the maximum tandem solar cell predicted efficiency.
On the other hand, if the doping concentration has adequately controlled, the parameters resulted in VOC = 2.6 V, JSC = 53 mA/cm2, FF = 62.40%, and PCE = 85.91%, considering a 2.5 µm thick absorber layer with n = 1 × 1017 cm−3. Therefore, this proposal reports high efficiency for InGaN solar cells, near the theoretical limit, considering real technological issues to simplify future InGaN-based step-graded layers. However, the present work has not considered considerations such as piezoelectric polarization in the wurtzite and the speed of the interface recombination. The polarization effect reduces the efficiency from 85.91% to 70.11% and 63.05% for a fully stressed structure, for 1 × 1017 cm−3 and 1 × 1019 cm−3, respectively. The absence of strong polarization in the structure drives new studies enabling the zincblende approach as an alternative for developing InGaN-based solar cells. The addition of the electron and hole surface recombination velocities have shown a reduction of the structure, obtaining a 65.44% power conversion efficiency with a JSC = 52.45 mA/cm2 and VOC = 2.56 V, emphasizing the importance of compensating this effect to stop the efficiency lack due to defects and their impact on the interfaces.

Author Contributions

Conceptualization, R.M.-R., H.I.S.-C. and C.A.H.-G.; methodology, R.M.-R.; software, R.M.-R., M.P.-P., M.L.P.-C., R.T.-H. and H.I.S.-C.; validation, R.G.-C., J.L.C.-A. and C.A.H.-G.; formal analysis, R.M.-R., R.T.-H., H.I.S.-C. and R.G.-C.; investigation, R.M.-R. and H.I.S.-C.; resources, R.G.-C., J.L.C.-A. and C.A.H.-G.; data curation, H.I.S.-C., M.P.-P., M.L.P.-C., R.T.-H. and C.A.H.-G.; writing—original draft, R.M.-R. and H.I.S.-C.; writing—review and editing, R.G.-C., M.P.-P., J.L.C.-A. and C.A.H.-G.; visualization, R.M.-R., H.I.S.-C. and M.L.P.-C.; supervision, C.A.H.-G.; project administration, C.A.H.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research is partly funded by the project for fabrication and evaluation of solar cells titled “Prototipo experimental para fabricación de películas delgadas CZTS utilizando el método SILAR” granted by Instituto de Ciencia, Tecnología e Innovación del Estado de Chiapas as part of Programa para el Desarrollo de la Investigación Científica, Desarrollo Tecnológico e Innovación 2022.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Authors thank Instituto de Ciencia, Tecnología e Innovación del Estado de Chiapas for its support. Rubén Martínez-Revuelta, Horacio I. Solís-Cisneros, and Raúl Trejo-Hernández thank Consejo Nacional de Ciencia y Tecnología in Mexico (CONACYT) for its postgraduate scholarship program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoshikawa, K.; Kawasaki, H.; Yoshida, W.; Irie, T.; Konishi, K.; Nakano, K.; Uto, T.; Adachi, D.; Kanematsu, M.; Uzu, H.; et al. Silicon heterojunction solar cell with interdigitated back contacts for a photoconversion efficiency over 26%. Nat. Energy 2017, 2, 17032. [Google Scholar] [CrossRef]
  2. Bhuiyan, A.G.; Sugita, K.; Hashimoto, A.; Yamamoto, A. InGaN solar cells: Present state of the art and important challenges. IEEE J. Photovolt. 2012, 2, 276–293. [Google Scholar] [CrossRef]
  3. Valdueza-Felip, S.; Ajay, A.; Redaelli, L.; Chauvat, M.P.; Ruterana, P.; Cremel, T.; Jiménez-Rodríguez, M.; Kheng, K.; Monroy, E. P-i-n InGaN homojunctions (10–40% In) synthesized by plasma-assisted molecular beam epitaxy with extended photoresponse to 600nm. Sol. Energy Mater. Sol. Cells 2017, 160, 355–360. [Google Scholar] [CrossRef] [Green Version]
  4. Boudaoud, C.; Hamdoune, A.; Allam, Z. Simulation and optimization of a tandem solar cell based on InGaN. Math. Comput. Simul. 2020, 167, 194–201. [Google Scholar] [CrossRef]
  5. Islam, S.M.; Protasenko, V.; Rouvimov, S.; Xing, H.; Jena, D. High-quality InN films on GaN using graded InGaN buffers by MBE. Jpn. J. Appl. Phys. 2016, 55, 05FD12. [Google Scholar] [CrossRef] [Green Version]
  6. Bazioti, C.; Papadomanolaki, E.; Kehagias, T.; Walther, T.; Smalc-Koziorowska, J.; Pavlidou, E.; Komninou, P.; Karakostas, T.; Iliopoulos, E.; Dimitrakopulos, G.P. Defects, strain relaxation, and compositional grading in high indium content InGaN epilayers grown by molecular beam epitaxy. J. Appl. Phys. 2015, 118, 155301. [Google Scholar] [CrossRef] [Green Version]
  7. Pantzas, K.; El Gmili, Y.; Dickerson, J.; Gautier, S.; Largeau, L.; Mauguin, O.; Patriarche, G.; Suresh, S.; Moudakir, T.; Bishop, C.; et al. Semibulk InGaN: A novel approach for thick, single phase, epitaxial InGaN layers grown by MOVPE. J. Cryst. Growth 2013, 370, 57–62. [Google Scholar] [CrossRef]
  8. Hernández-Gutiérrez, C.A.; Morales-Acevedo, A.; Cardona, D.; Contreras-Puente, G.; López-López, M. Analysis of the performance of InxGa1−xN based solar cells. SN Appl. Sci. 2019, 1, 628. [Google Scholar] [CrossRef] [Green Version]
  9. Liliental-Weber, Z.; Yu, K.M.; Hawkridge, M.; Bedair, S.; Berman, A.E.; Emara, A.; Khanal, D.R.; Wu, J.; Domagala, J.; Bak-Misiuk, J. Structural perfection of InGaN layers and its relation to photoluminescence. Phys. Status Solidi Curr. Top. Solid State Phys. 2009, 6, 2626–2631. [Google Scholar] [CrossRef]
  10. Dagotto, E.; Hotta, T.; Moreo, A. Colossal magnetoresistant materials: The key role of phase separation. Phys. Rep. 2001, 344, 1–153. [Google Scholar] [CrossRef]
  11. Nagaev, É.L. Lanthanum manganites and other giant-magnetoresistance magnetic conductors. Uspekhi Fiz. Nauk 1996, 166, 857–858. [Google Scholar] [CrossRef]
  12. Khomskii, D. Phase separation, percolation and giant isotope effect in manganites. Phys. B Condens. Matter 2000, 280, 325–330. [Google Scholar] [CrossRef] [Green Version]
  13. Bagavath, C.; Kumar, J. Investigations on the structural, optical and electrical properties of InxGa1−xN thin films. Mater. Chem. Phys. 2019, 234, 318–322. [Google Scholar] [CrossRef]
  14. Rao, M.; Kim, D.; Mahajan, S. Compositional dependence of phase separation in InGaN layers. Appl. Phys. Lett. 2004, 85, 1961–1963. [Google Scholar] [CrossRef]
  15. Li, N.; Wang, S.-J.; Park, E.-H.; Feng, Z.C.; Tsai, H.-L.; Yang, J.-R.; Ferguson, I. Suppression of phase separation in InGaN layers grown on lattice-matched ZnO substrates. J. Cryst. Growth 2009, 311, 4628–4631. [Google Scholar] [CrossRef]
  16. Chen, Z.Z.; Qin, Z.X.; Hu, X.D.; Yu, T.J.; Yang, Z.J.; Tong, Y.Z.; Ding, X.M.; Zhang, G.Y. Study of photoluminescence and absorption in phase-separation InGaN films. Phys. B Condens. Matter 2004, 344, 292–296. [Google Scholar] [CrossRef]
  17. El Huni, W.; Karrakchou, S.; Halfaya, Y.; Arif, M.; Jordan, M.B.; Puybaret, R.; Ayari, T.; Ennakrachi, H.; Bishop, C.; Gautier, S.; et al. Nanopyramid-based absorber to boost the efficiency of InGaN solar cells. Sol. Energy 2019, 190, 93–103. [Google Scholar] [CrossRef]
  18. Gherasoiu, I.; Yu, K.M.; Reichertz, L.A.; Walukiewicz, W. InGaN doping for high carrier concentration in plasma-assisted molecular beam epitaxy. Phys. Status Solidi 2014, 11, 381–384. [Google Scholar] [CrossRef]
  19. Khettou, A.; Zeydi, I.; Chellali, M.; Ben Arbia, M.; Mansouri, S.; Helal, H.; Maaref, H. Simulation and optimization of InGaN Schottky solar cells to enhance the interface quality. Superlattices Microstruct. 2020, 142, 106539. [Google Scholar] [CrossRef]
  20. Wu, J.; Walukiewicz, W. Band gaps of InN and group III nitride alloys. Superlattices Microstruct. 2003, 34, 63–75. [Google Scholar] [CrossRef]
  21. Tangi, M.; De, A.; Shivaprasad, S.M. Role of dislocations and carrier concentration in limiting the electron mobility of InN films grown by plasma assisted molecular beam epitaxy. J. Appl. Phys. 2018, 123, 015701. [Google Scholar] [CrossRef]
  22. Ould Saad Hamady, S.; Adaine, A.; Fressengeas, N. Numerical simulation of InGaN Schottky solar cell. Mater. Sci. Semicond. Process. 2016, 41, 219–225. [Google Scholar] [CrossRef] [Green Version]
  23. Ma, N.; Wang, X.Q.; Liu, S.T.; Chen, G.; Pan, J.H.; Feng, L.; Xu, F.J.; Tang, N.; Shen, B. Hole mobility in wurtzite InN. Appl. Phys. Lett. 2011, 98, 18–21. [Google Scholar] [CrossRef]
  24. Ohnishi, K.; Amano, Y.; Fujimoto, N.; Nitta, S.; Watanabe, H.; Honda, Y.; Amano, H. Electrical properties and structural defects of p-type GaN layers grown by halide vapor phase epitaxy. J. Cryst. Growth 2021, 566–567, 126173. [Google Scholar] [CrossRef]
  25. Morkoç, H. General Properties of Nitrides. In Handbook of Nitride Semiconductors and Devices; Wiley Online Books: Hoboken, NJ, USA, 2008; pp. 1–129. [Google Scholar]
  26. Madi, L.; Bouchama, I.; Bouarissa, N. Effect of light wavelengths on the non-polar InGaN-based thin film solar cells performances using one-dimensional modeling. J. Sci. Adv. Mater. Devices 2019, 4, 509–514. [Google Scholar] [CrossRef]
  27. Tsai, Y.-C.; Bayram, C. Band Alignments of Ternary Wurtzite and Zincblende III-Nitrides Investigated by Hybrid Density Functional Theory. ACS Omega 2020, 5, 3917–3923. [Google Scholar] [CrossRef] [Green Version]
  28. Davydov, V.; Klochikhin, A.; Seisyan, R.; Emtsev, V.; Ivanov, S.; Bechstedt, F.; Furthmüller, J.; Harima, H.; Mudryi, A.; Aderhold, J.; et al. Absorption and Emission of Hexagonal InN. Evidence of Narrow Fundamental Band Gap. Phys. Status Solidi B. 2002, 80, r1–r3. [Google Scholar] [CrossRef]
  29. Tran, B.T.; Chang, E.Y.; Trinh, H.D.; Lee, C.T.; Sahoo, K.C.; Lin, K.L.; Huang, M.C.; Yu, H.W.; Luong, T.T.; Chung, C.C.; et al. Fabrication and characterization of n-In 0.4Ga 0.6N/p-Si solar cell. Sol. Energy Mater. Sol. Cells 2012, 102, 208–211. [Google Scholar] [CrossRef]
  30. Xing, Z.; Yang, W.; Yuan, Z.; Li, X.; Wu, Y.; Long, J.; Jin, S.; Zhao, Y.; Liu, T.; Bian, L.; et al. Growth and Characterization of High In-content InGaN grown by MBE using Metal Modulated Epitaxy Technique (MME). J. Cryst. Growth 2019, 516, 57–62. [Google Scholar] [CrossRef]
  31. Itoh, T.; Kobayashi, A.; Ueno, K.; Ohta, J.; Fujioka, H. Fabrication of InGaN thin-film transistors using pulsed sputtering deposition. Sci. Rep. 2016, 6, 29500. [Google Scholar] [CrossRef]
  32. Lin, S.K.; Wu, K.T.; Huang, C.P.; Liang, C.T.; Chang, Y.H.; Chen, Y.F.; Chang, P.H.; Chen, N.C.; Chang, C.A.; Peng, H.C.; et al. Electron transport in In-rich InxGa1−xN films. J. Appl. Phys. 2005, 97, 046101. [Google Scholar] [CrossRef]
  33. Yadav, B.S.; Mohanta, P.; Srinivasa, R.S.; Major, S.S. Electrical and optical properties of transparent conducting InxGa1−xN alloy films deposited by reactive co-sputtering of GaAs and Indium. Thin Solid Film. 2014, 555, 179–184. [Google Scholar] [CrossRef]
  34. Pantha, B.N.; Wang, H.; Khan, N.; Lin, J.Y.; Jiang, H.X. Origin of background electron concentration in InxGa1−xN alloys. Phys. Rev. B 2011, 84, 075327. [Google Scholar] [CrossRef] [Green Version]
  35. Hsu, L.; Jones, R.E.; Li, S.X.; Yu, K.M.; Walukiewicz, W. Electron mobility in InN and III-N alloys. J. Appl. Phys. 2007, 102, 73705. [Google Scholar] [CrossRef]
  36. Hu, Y.; Hernández-Gutiérrez, C.A.; Solís-Cisneros, H.I.; Santana, G.; Kudriatsev, Y.; Camas-Anzueto, J.L.; López-López, M. Blue luminescence origin and Mg acceptor saturation in highly doped zincblende GaN with Mg. J. Alloy. Compd. 2022, 897, 163133. [Google Scholar] [CrossRef]
  37. Wierer, J.J.; Fischer, A.J.; Koleske, D.D. The impact of piezoelectric polarization and nonradiative recombination on the performance of (0001) face GaN/InGaN photovoltaic devices. Appl. Phys. Lett. 2010, 96, 051107. [Google Scholar] [CrossRef]
  38. Chen, X.; Matthews, K.D.; Hao, D.; Schaff, W.J.; Eastman, L.F.; Walukiewicz, W.; Ager, J.W.; Yu, K.M. Characterization of Mg-doped InGaN and InAlN alloys grown by MBE for solar applications. In Proceedings of the 2008 33rd IEEE Photovoltaic Specialists Conference, San Diego, CA, USA, 11–16 May 2008; pp. 1–6. [Google Scholar]
  39. Pantha, B.N.; Sedhain, A.; Li, J.; Lin, J.Y.; Jiang, H.X. Electrical and optical properties of p-type InGaN. Appl. Phys. Lett. 2009, 95, 261904. [Google Scholar] [CrossRef] [Green Version]
  40. Kuo, D.H.; Tuan, T.T.A.; Li, C.C.; Yen, W.C. Electrical and structural properties of Mg-doped InxGa1−xN (x ≤ 0.1) and p-InGaN/n-GaN junction diode made all by RF reactive sputtering. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2015, 193, 13–19. [Google Scholar] [CrossRef]
  41. Sang, L.; Liao, M.; Ikeda, N.; Koide, Y.; Sumiya, M. Enhanced performance of InGaN solar cell by using a super-thin AlN interlayer. Appl. Phys. Lett. 2011, 99, 161109. [Google Scholar] [CrossRef]
  42. Kumakura, K.; Makimoto, T.; Kobayashi, N. High hole concentrations in Mg-doped InGaN grown by MOVPE. J. Cryst. Growth 2000, 221, 267–270. [Google Scholar] [CrossRef]
  43. Kumakura, K.; Makimoto, T.; Kobayashi, N. Activation Energy and Electrical Activity of Mg in Mg-doped InxGa1−xN (x < 0.2). Jpn. J. Appl. Phys 2000, 39, 337. [Google Scholar]
  44. Chen, P.C.; Chen, C.H.; Chang, S.J.; Su, Y.K.; Chang, P.C.; Huang, B.R. High hole concentration of p-type InGaN epitaxial layers grown by MOCVD. In Proceedings of the Thin Solid Films, Singapore, 14 August 2006; Volume 498, pp. 113–117. [Google Scholar]
  45. Sang, L.; Liao, M.; Koide, Y.; Sumiya, M. InGaN-based thin film solar cells: Epitaxy, structural design, and photovoltaic properties. J. Appl. Phys. 2015, 117, 105706. [Google Scholar] [CrossRef]
  46. Chang, C.A.; Tang, T.Y.; Chang, P.H.; Chen, N.C.; Liang, C. Te Magnesium doping of In-rich InGaN. Jpn. J. Appl. Phys. Part 1 Regul. Pap. Short Notes Rev. Pap. 2007, 46, 2840–2843. [Google Scholar] [CrossRef]
  47. Zhang, L.; Wang, R.; Liu, Z.; Cheng, Z.; Tong, X.; Xu, J.; Zhang, S.; Zhang, Y.; Chen, F. Regulation of hole concentration and mobility and first-principle analysis of mg-doping in InGaN grown by MOCVD. Materials 2021, 14, 5339. [Google Scholar] [CrossRef]
  48. Suski, T.; Staszczak, G.; Grzanka, S.; Czernecki, R.; Litwin-Staszewska, E.; Piotrzkowski, R.; Dmowski, L.H.; Khachapuridze, A.; Kryśko, M.; Perlin, P.; et al. Hole carrier concentration and photoluminescence in magnesium doped InGaN and GaN grown on sapphire and GaN misoriented substrates. J. Appl. Phys. 2010, 108, 023516. [Google Scholar] [CrossRef]
  49. Wang, X.; Che, S.B.; Ishitani, Y.; Yoshikawa, A. Hole mobility in Mg-doped p-type InN films. Appl. Phys. Lett. 2008, 92, 132108. [Google Scholar] [CrossRef]
  50. Wang, K.; Araki, T.; Yu, K.M.; Katsuki, T.; Mayer, M.A.; Alarcon-Llado, E.; Ager, J.W.; Walukiewicz, W.; Nanishi, Y. P-type InGaN across the entire alloy composition range. Appl. Phys. Lett. 2013, 102, 102111. [Google Scholar] [CrossRef]
  51. Jampana, B.R.; Ferguson, I.T.; Opila, R.L.; Honsberg, C.B. Utilizing polarization induced band bending for InGaN solar cell design. Mater. Res. Soc. Symp. Proc. 2009, 1167, 3–8. [Google Scholar] [CrossRef]
  52. Kumakura, K.; Makimoto, T.; Kobayashi, N. Mg-acceptor activation mechanism and transport characteristics in p-type InGaN grown by metalorganic vapor phase epitaxy. J. Appl. Phys. 2003, 93, 3370–3375. [Google Scholar] [CrossRef]
  53. Hernández-Gutiérrez, C.A.; Casallas-Moreno, Y.L.; Rangel-Kuoppa, V.-T.; Cardona, D.; Hu, Y.; Kudriatsev, Y.; Zambrano-Serrano, M.A.; Gallardo-Hernandez, S.; Lopez-Lopez, M. Study of the heavily p-type doping of cubic GaN with Mg. Sci. Rep. 2020, 10, 16858. [Google Scholar] [CrossRef]
  54. Tsai, Y.C.; Bayram, C. Mitigate self-compensation with high crystal symmetry: A first-principles study of formation and activation of impurities in GaN. Comput. Mater. Sci. 2021, 190, 110283. [Google Scholar] [CrossRef]
  55. Radojcic, R.; Hill, A.E.; Hampshire, M.J. A numerical model of a graded band gap CdSxTer1−x solar cell. Sol. Cells 1981, 4, 109–120. [Google Scholar] [CrossRef]
  56. Sassi, G. Theoretical analysis of solar cells based on graded bandgap structures. J. Appl. Phys. 1983, 54, 5421–5427. [Google Scholar] [CrossRef]
  57. Sarollahi, M.; Zamani, M.; Aldawsari, M.A.; Allaparthi, R.; Alhelais, R.; Refaei, M.A.; Uddin Maruf, M.H.; Ware, M.E. Modeling of Λ graded InxGa1−xN solar cells: Comparison of strained and relaxed features. J. Photonics Energy 2022, 12, 03510. [Google Scholar] [CrossRef]
  58. Morales-Acevedo, A. Effective absorption coefficient for graded bandgap semiconductors and the expected photocurrent density in solar cells. Sol. Energy Mater. Sol. Cells 2009, 93, 41–44. [Google Scholar] [CrossRef]
  59. Kazazis, S.A.; Papadomanolaki, E.; Androulidaki, M.; Kayambaki, M.; Iliopoulos, E. Optical properties of InGaN thin films in the entire composition range. J. Appl. Phys. 2018, 123, 125101. [Google Scholar] [CrossRef]
  60. Park, N.-G. Perovskite solar cells: An emerging photovoltaic technology. Mater. Today 2015, 18, 65–72. [Google Scholar] [CrossRef]
  61. Xu, Q.; Johnson, C.; Disney, C.; Pillai, S. Enhanced Broadband Light Trapping in c-Si Solar Cells Using Nanosphere-Embedded Metallic Grating Structure. IEEE J. Photovolt. 2016, 6, 61–67. [Google Scholar] [CrossRef]
  62. Kumawat, U.K.; Kumar, K.; Bhardwaj, P.; Dhawan, A. Indium-rich InGaN/GaN solar cells with improved performance due to plasmonic and dielectric nanogratings. Energy Sci. Eng. 2019, 7, 2469–2482. [Google Scholar] [CrossRef] [Green Version]
  63. Tsai, Y.-C.; Bayram, C.; Leburton, J.-P. Effect of Auger Electron–Hole Asymmetry on the Efficiency Droop in InGaN Quantum Well Light-Emitting Diodes. IEEE J. Quantum Electron. 2022, 58, 3300109. [Google Scholar] [CrossRef]
  64. Muyeed, S.A.A.; Borovac, D.; Xue, H.; Wei, X.; Song, R.; Tansu, N.; Wierer, J.J. Recombination Rates of InxGa1−xN/AlyGa1−yN/GaN Multiple Quantum Wells Emitting From 640 to 565 nm. IEEE J. Quantum Electron. 2021, 57, 3200207. [Google Scholar] [CrossRef]
  65. Green, M.A.; Dunlop, E.D.; Hohl-Ebinger, J.; Yoshita, M.; Kopidakis, N.; Hao, X. Solar Cell Efficiency Tables (Version 58). Prog. Photovolt. Res. Appl. 2021, 29, 657–667. [Google Scholar] [CrossRef]
Figure 1. (a) Electron concentration as a function of In molar fraction and (b) hole concentration as a function of In molar fraction [3,13,18,28,30,31,32,33,37,38,39,40,41,42,44,45,46,50,51].
Figure 1. (a) Electron concentration as a function of In molar fraction and (b) hole concentration as a function of In molar fraction [3,13,18,28,30,31,32,33,37,38,39,40,41,42,44,45,46,50,51].
Micromachines 13 01828 g001
Figure 2. The AM1.5 solar spectra range absorbed by InGaN ternary alloy in the whole In content range.
Figure 2. The AM1.5 solar spectra range absorbed by InGaN ternary alloy in the whole In content range.
Micromachines 13 01828 g002
Figure 3. The absorption constant (A) for InGaN alloy with different In molar fractions calculated from the Absorption coefficient vs. (E − Eg)1/2 plot [59].
Figure 3. The absorption constant (A) for InGaN alloy with different In molar fractions calculated from the Absorption coefficient vs. (E − Eg)1/2 plot [59].
Micromachines 13 01828 g003
Figure 4. Photocurrent comparison between different materials with graded bandgap and the effect of the thickness in the JSC convergence [58,60,61].
Figure 4. Photocurrent comparison between different materials with graded bandgap and the effect of the thickness in the JSC convergence [58,60,61].
Micromachines 13 01828 g004
Figure 5. Solar cell structure to simulate (left) and band diagram of the structure under dark conditions (right).
Figure 5. Solar cell structure to simulate (left) and band diagram of the structure under dark conditions (right).
Micromachines 13 01828 g005
Figure 6. (a) Photon absorption in the structure and (b) influence of electron concentration in the absorber layer on the electric field.
Figure 6. (a) Photon absorption in the structure and (b) influence of electron concentration in the absorber layer on the electric field.
Micromachines 13 01828 g006
Figure 7. Radiative, Auger and Shockley–Read–Hall recombination rates under 1-sun at AM1.5 standard light conditions.
Figure 7. Radiative, Auger and Shockley–Read–Hall recombination rates under 1-sun at AM1.5 standard light conditions.
Micromachines 13 01828 g007
Figure 8. EQE performance with n1 = 1 × 1017 cm−3 and n2 = 1 × 1019 cm−3 [65].
Figure 8. EQE performance with n1 = 1 × 1017 cm−3 and n2 = 1 × 1019 cm−3 [65].
Micromachines 13 01828 g008
Figure 9. Summary of results for step-graded InGaN-based solar cell. (a) Open-circuit voltage, (b) current density, (c) fill factor, (d) efficiency, and (e) J-V curves for lowest and highest efficiencies for n = 1 × 1017 cm−3 in the absorber layer.
Figure 9. Summary of results for step-graded InGaN-based solar cell. (a) Open-circuit voltage, (b) current density, (c) fill factor, (d) efficiency, and (e) J-V curves for lowest and highest efficiencies for n = 1 × 1017 cm−3 in the absorber layer.
Micromachines 13 01828 g009
Figure 10. Effect of the stress relaxation and piezoelectric mitigation in the power conversion efficiency of the structure.
Figure 10. Effect of the stress relaxation and piezoelectric mitigation in the power conversion efficiency of the structure.
Micromachines 13 01828 g010
Table 1. Parameters of the binary compounds.
Table 1. Parameters of the binary compounds.
ParameterInNGaN
Bandgap0.7 eV [19,20]3.4 eV [13]
Electron mobility250 cm2/Vs [21]300 cm2/Vs [22]
Hole mobility30 cm2/Vs [23]14 cm2/Vs [24]
The effective mass of DOS in the CB0.11 [25]0.22 [25]
The effective mass of DOS in the VB1.63 [25]1.4 [25]
Dielectric constant15.3 [26]8.9 [22]
Table 2. Recombination parameters.
Table 2. Recombination parameters.
ParameterValue
τp (electron lifetime)1.3 ns [62]
τp (hole lifetime)5.4 ns [62]
Auger recombination coefficients3.0 × 10−30 cm6/s [63]
Radiative recombination coefficient1.4 × 10−13 cm3/s [64]
Table 3. Comparison of VOC, JSC, FF, and Efficiency of the solar cell structure without polarization effect.
Table 3. Comparison of VOC, JSC, FF, and Efficiency of the solar cell structure without polarization effect.
Parameteri = 1 × 1016 cm−3i = 1 × 1017 cm−3i = 1 × 1019 cm−3
Absorber Layer Thickness (µm)Absorber Layer Thickness (µm)Absorber Layer Thickness (µm)
0.511.522.50.511.522.50.511.522.5
VOC (V)2.052.402.502.562.602.052.402.502.562.602.712.762.782.802.80
JSC (mA/cm2)46.7249.2850.9350.3252.7246.7249.5751.4050.6652.9511.1516.3520.5724.4427.89
FF62.8462.6063.1660.5961.9062.8462.8163.9161.7062.4077.6680.3884.0185.4887.30
Efficiency60.1974.0480.4278.0685.8560.1974.7282.1380.0285.9121.9636.2648.0358.4868.17
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínez-Revuelta, R.; Solís-Cisneros, H.I.; Trejo-Hernández, R.; Pérez-Patricio, M.; Paniagua-Chávez, M.L.; Grajales-Coutiño, R.; Camas-Anzueto, J.L.; Hernández-Gutiérrez, C.A. Numerical Study of a Solar Cell to Achieve the Highest InGaN Power Conversion Efficiency for the Whole In-Content Range. Micromachines 2022, 13, 1828. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13111828

AMA Style

Martínez-Revuelta R, Solís-Cisneros HI, Trejo-Hernández R, Pérez-Patricio M, Paniagua-Chávez ML, Grajales-Coutiño R, Camas-Anzueto JL, Hernández-Gutiérrez CA. Numerical Study of a Solar Cell to Achieve the Highest InGaN Power Conversion Efficiency for the Whole In-Content Range. Micromachines. 2022; 13(11):1828. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13111828

Chicago/Turabian Style

Martínez-Revuelta, Rubén, Horacio I. Solís-Cisneros, Raúl Trejo-Hernández, Madaín Pérez-Patricio, Martha L. Paniagua-Chávez, Rubén Grajales-Coutiño, Jorge L. Camas-Anzueto, and Carlos A. Hernández-Gutiérrez. 2022. "Numerical Study of a Solar Cell to Achieve the Highest InGaN Power Conversion Efficiency for the Whole In-Content Range" Micromachines 13, no. 11: 1828. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13111828

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