Deborah
Eric
*ab,
Jianliang
Jiang
*b,
Ali
Imran
c and
Abbas Ahmad
Khan
d
aDepartment of Electronics Engineering, Dong-A University, Busan 49315, South Korea. E-mail: ericdeborah@dau.ac.kr
bSchool of Optics and Photonics, Beijing Institute of Technology, Beijing 100081, China. E-mail: jiangjianliang@bit.edu.cn
cSchool of Micro and Nano Electronics, ZJU-Hangzhou Global Scientific and Technological Innovation Center, State Key Labs of Silicon Materials and Micro-Nano Fabrication Center, Zhejiang University, Hangzhou 311200, China
dIMDEA-Nanociencia, Campus de Cantoblanco, 28049 Madrid, Spain
First published on 12th June 2024
It is essential to have an adequately thick active layer to achieve efficient performance in quantum dot intermediate band solar cells (QD-IBSC) utilizing InxGa1−xN with high indium concentrations. The thickness plays a crucial role in maximizing photon absorption and optimizing the overall effectiveness of the solar cell (SC). In this paper, we introduce QD-IBSC with Ga-face (0 0 0 1) applying 1 nm i-GaN interlayers, which will provide strain relaxation to the In0.5Ga0.5N/GaN QD layer for increasing photovoltaic performance. Normally, the coupling among QDs splits the quantized energy level and leads to the formation of minibands within the forbidden region of conventional SC. In particular, the QDs are sensitive to dot regimentation and thus affect the properties of QD-IBSC. The electronic band structure of these QDs is controlled by changing the size of the QD, interdot distances and regimentation. In this paper, optimization of the optical structure of the QD-IBSC is performed by investigating the calculation results of both the maximum number of absorbed photons and the carrier transport property through tunneling simultaneously as a function of the thickness of the i-GaN interlayers. For the calculation, the three-dimensional regimented array of InxGa1−xN QD is analyzed using an envelope function. This work demonstrates Ga-face n–i–p structure (n-GaN/i-GaN:In0.5Ga0.5N:i-GaN/p-GaN) utilizing the 20 periods of 3 nm thick In0.5Ga0.5N QD layers and a GaN layer of 1 nm thickness can achieve a maximum conversion efficiency of 48%.
Ab initio calculations5,24,25 are employed to select various IB materials. These materials possess fundamental properties such as strong sub-bandgap absorption, extended carrier lifetimes, and excellent carrier mobilities. Notably, the III-nitride family26 exhibits these desirable characteristics. Among them, the direct bandgap InxGa1−xN alloy showcases the most promising properties for optoelectronic devices.27 Its tunable bandgap, better drift velocity, high mobility, high absorption coefficient and high tolerance to radiation enable efficient absorption of a broad range of the solar spectrum in SC.28,29 InxGa1−xN exists in two different structures: zinc blende and wurtzite.28 While the wurtzite structure is thermodynamically stable, it lacks an inversion plane perpendicular to the c-axis. This leads to spontaneous and piezoelectric polarization caused by an internal electric field.29 On the other hand, zinc blende possesses high crystal symmetry, resulting in the absence of crystal polarization.30–32 In our investigation, we explore the utilization of wurtzite InxGa1−xN QDs embedded in GaN for IBSCs.
However, the fabrication of InxGa1−xN SC still faces significant challenges.27 A few of them include: (1) the bandgap of InxGa1−xN varies indirectly with the indium composition, which can lead to poor crystallinity and high defect rate that significantly impact the emission process.33 (2) The InxGa1−xN films often experience indium cluster segregation at the surface, caused by a solid phase miscibility gap resulting from a lattice mismatch and differences in enthalpy formation between GaN and InN.34 (3) Epolarization effects also affect the efficiency of optoelectronic devices, as discussed in the subsequent section.
Aouami et al. calculated the behaviour of different geometries for multiple quantum dot solar cells.35 Pérez et al. considered a juxtaposition of two kinds of QDs and determined the energy levels.36 Aissat et al. modelled In0.25Ga0.75N/GaN QDSC with five layers of quantum dots at the temperature of 285 K and concluded that the efficiency of SC is highly dependent on temperature.37 Chen et al. studied the single junction InxGa1−xN SC with three intermediate band gaps. Their research revealed that increasing the bandgap (1 > x > 0) results in a higher number of IBs within the absorber layer, leading to an efficiency of over 50%. While IBs can be obtained at low indium concentrations in InxGa1−xN, higher concentrations often result in clustering, limiting the efficiency to around 12%.38 Deng et al. conducted InxGa1−xN/InN IBSC computations with ortho-rhombic symmetry of QDs embedded in periodic arrays. Without considering polarization fields, they achieved a maximum efficiency of 60.3% for QD size 3.3 nm and interdot size 2 nm.39 Zhang et al. ascertained an indium concentration of 45%, resulting in an efficiency of 61.67% for cubic QD size 3.3 nm and interdot size 2 nm.40 Several studies have reported on simple PIN solar cells with polarization effects and potential solutions.41–44 However, to the best of our knowledge, no research group has reported a comprehensive study on quantum dot intermediate band solar cells (QD-IBSCs) that optimizes both the quantum dot layers and the active layer to reduce the lattice mismatch between InxGa1−xN/GaN layers.
In this study, we investigate and analyse the photovoltaic properties of n–i–p Ga-face (0 0 0 1) In0.5Ga0.5N/GaN QD-IBSC with three-dimensional In0.5Ga0.5N/GaN QDs embedded between n-type and p-type GaN emitters, with a 1 nm interlayer incorporated in the absorber layer. This interlayer aims to mitigate the strain relaxation caused by compressive forces at the interface between In0.5Ga0.5N and GaN. By optimizing the size and distance of the quantum dots, we calculate the position and width of the intermediate band (IB). Additionally, we determine the energies of the quantum dots by solving the Schrödinger equation in the presence of Epolarization.
The wurtzite structure of InxGa1−xN gives rise to a pronounced polarization effect due to the non-centric symmetry of charges. This Epolarization polarization effect, along with potential barriers and band bending, significantly influences the behaviour of devices.47 The characteristics of InxGa1−xN alloy exhibit a nonlinear dependence on the alloy composition, resulting in complex properties.48 The wurtzite structure encompasses two types of polarization. The first type, known as Psp, is an inherent property of wurtzite semiconductors and arises from the absence of an inversion plane perpendicular to the c-axis.49 This phenomenon contributes to variations in dislocation density and surface morphology.46 Importantly, the Psp values for InN and GaN are nearly equal.49 The calculation of Psp can be determined using the following relation:50
Psp(InxGa1−xN) = xPsp(InN) + (1 − x)Psp(GaN) − bInGax(1 − x) | (1) |
The presence of a second type of polarization, referred to as Ppz, can be observed in strained epitaxial layers. This polarization arises from the lattice-relaxation strain that occurs when two different materials are grown on the same substrate. In the case of wurtzite nitrides, the growth under biaxial strain induces piezoelectric polarization.50 The magnitude of Ppz depends on the composition of the InxGa1−xN alloy and is related to its binary constituents, InN and GaN, through the following relation:51
Ppz(InxGa1−xN) = xPpz(InN) + (1 − x)Ppz(GaN) | (2) |
In our specific case, the growth of InN QDs on a GaN layer with lattice mismatch induces strain effects due to the relaxation of the mismatched layers. A 1 nm interlayer of i-GaN is incorporated to mitigate this strain. The total polarization (PT = Epolarization) is determined by the combined contributions of Ppz and Psp. Since the layers are in compression, the two polarizations act in opposite directions (as shown in the inset in Fig. 1(b)), resulting in a reduced net polarization. The presence of a built-in electric field and lower polarization field facilitates the movement of more carriers across the interfaces from the n-type to the p-type layer. Fig. 1(b) displays the calculated PT at the interfaces, along with the directions of the spontaneous polarization (Psp) and piezoelectric polarization (Ppz) at the interface. The structures demonstrate the lattice orientations and the corresponding vectors of Psp.52
The following list of assumptions is used to carry out limiting efficiency calculations.
1. The solar cell absorbs blackbody radiation at a temperature of Ts = 6000 K and ambient Ta = 300 K and emits blackbody radiation at ambient Ta = 300 K.
2. Only radiative transitions occur between the bands.
3. All photons above the lowest energy gap are absorbed, and no high-energy photon is used in a low-energy process.
4. Carrier mobility is infinite; consequently, the quasi-Fermi energy levels are constant throughout the cell.
5. Only one electron–hole pair is created per photon.
6. A perfect mirror is located on the back of the cell so that radiation makes a double pass and can only escape through the front illumination area.
7. The net photon flux (number of incident photons minus the number of emitted photons) equals the number of charge carrier pairs collected at the contacts.9
In the case of intermediate band solar cells (IBSCs), the detailed balance assumptions mentioned above are applied with slight modifications as follows:
Firstly, it is assumed that photons with photon energy Eph < Eg are absorbed, even though they fall below the energy bandgap. This is possible because an additional bandgap within the forbidden region has been introduced, which is isolated and has zero density of states. Secondly, the original bandgap Eg (Eg = E1 + E2) is split into two separate energy levels, E1 and E2. This splitting facilitates the absorption of photons with energy Eph < Eg. Therefore, in addition to absorbing photons with Eph ≥ Eg using the conventional mechanism, IBSCs can also absorb photons with Eph < Eg (assumption 3). As a result of absorbing photons with Eph < Eg, the absorptivity of the cell increases. However, it should be noted that only one additional electron–hole pair contributes to the short-circuit current density, Jsc (assumption 5). Moreover, the intermediate band (IB) is partially filled, and the valence band offset is considered to be negligible. It is important to mention that the output voltage is not calculated based on the IB (assumption 7).4
Taking the aforementioned factors into account, the electronic state of quantum dots (QDs) is determined by utilization of the Schrödinger equation, employing the Kronig–Penney model. Our previous work53 extensively discussed the theoretical framework behind this calculation. The position and width of the QDs are further optimized using the envelope function approximation. Simultaneously, the electronic properties are computed by solving Poisson's equation. To integrate these two calculations, we employ COMSOL Multiphysics, a software platform that combines the Schrödinger equation and electrostatics. This integration enables us to model the behavior of charge carriers in quantum-confined systems. Notably, the electric potential derived from the electrostatics is incorporated into the potential energy term of the Schrödinger equation within the COMSOL Multiphysics framework.
The solar cell and sun are considered blackbody systems. By applying the Roosbroeck–Shockley formula, the photon flux within the energy range of E1 to E2 can be determined using the following calculations,20
(3) |
μ = EFC − EFV = 3.39 eV, μ1 = EFIB − EFV = 1.89 eV, μ2 = EFC − EFIB = 1.5 eV, | (4) |
(5) |
(6) |
(7) |
The efficiency (η) can be calculated as:
(8) |
Numerical experimentation has been conducted to simulate the QD-IBSC device and analyze the behavior of stacked quantum dot layers. The calculation of the band structure employs a tight-binding model that superimposes wave functions for the individual quantum dots. The finite element method technique is utilized to solve the model to enhance computational efficiency. The simulation process involves several steps. Firstly, the eigenvalues of the quantum dots are determined by solving the Schrödinger equation.53 Subsequently, the impact of embedding QDs in the intrinsic layer of a Ga-face n–i–p SC under the conditions mentioned above is simulated. The modeling incorporates the material properties specified in Table 1.
Parameters | Values |
---|---|
Bandgap (Eg) | 1.6875 (eV) |
Lifetime (τ) | 6.5 (ns) |
Effective conduction band density of In0.5Ga0.5N | 1.6 × 1018 (cm−3) |
Effective valence band density of In0.5Ga0.5N | 9 × 1018 (cm−3) |
Effective electron mass of In0.5Ga0.5N | 0.16 m0 |
Effective hole mass of In0.5Ga0.5N | 0.585 m0 |
Electron affinity | 5.29875 (eV) |
Elastic constant (c13) | 100 (GPa) |
Elastic constant (c33) | 299 (GPa) |
Piezoelectric Const. (e31) | 0.04 (C m−2) |
Piezoelectric Const. (e31) | 0.88 (C m−2) |
Optical modes and the distribution of the electric field are also considered. The Fresnel model is applied to optimize the reflection and transmission of light. Boundaries are fine-tuned using Floquet's periodicity condition. Moreover, the influence of thin enough material strain is considered as per the equation described in Section 2 of the referenced article.
The three-dimensional doping profile of the proposed structure with excited state energy is shown in Fig. 2(a). The blue color shows the concentration of excess electrons, whereas the red indicates the excess holes. The green color indicates the intrinsic level with no doping. The inset shows the excited state energies of the QDs. Fig. 2(b) shows the energy band diagram for the proposed structure. Ec and Ev are the CB and VB energy levels, respectively. EIB is the IB energy level created when QDs are placed within the intrinsic level of n–i–p IBSC. These bands have their Fermi levels. Efc and Efv are the quasi-Fermi levels for the CB and VB. The intermediate band has its Fermi level EfIB which is separated from the quasi-Fermi levels of the CB and VB. This IB is half-filled with electrons by optimizing the distance between QDs and spacing between barrier regions.53Eg is the bandgap of GaN (3.39 eV) and EInN is the bandgap for InN (0.7 eV). E1 and E2 are the energy gaps between Ev and EIB, EIB and EC, respectively. Lqd is the distance between two adjacent QDs, and Lbr is the width of the barrier. Upon illumination, photons are absorbed; if Eph (energy from photons) ≤ Eg, this energy photon is absorbed and creates an electron–hole pair. Additional electron–hole pairs are created other than the conventional ones, which results in higher efficiency.53 Additionally, low-energy photons below Eg are absorbed, thus increasing the photocurrent.
Fig. 3(a)–(c) shows the 3 × 3 layer of QDs with their localized wave functions for the QD size of 3 nm and their interdot distance of 2 nm. Three eigenvalues are inside the conduction band at 1.3479, 1.8472, and 2.6289 eV. The first state is the pure quantum dot state. The wave function for this state is zero at the center of the quantum dot and vanishes at the boundary. The second and third states are weakly confined concerning the first state.57 The band formation of IB strongly depends on the dot sizes53 and interdot spacing between each QD. The wave functions of the QDs will overlap for certain QD sizes and interdot spacings such that energy level splitting occurs. Small confinement in barriers increases the tunneling probability of the wavefunctions.53 Due to the quantum confinement effect, QD coupling occurs as soon as they are brought near each other. This results in new band formation within the available density of states (Fig. 4). The colors in Fig. 3 show the QD states for the excited states.
Fig. 3 WF of a system: (a) ground state E0 = 1.34 eV, (b) first excited state E1 = 1.85 eV and (c) second excited state E2 = 2.63 eV. |
Fig. 4 Variation of the first optically strongest four states in the conduction band interdot spacing. |
As identical quantum dots (QDs) are brought together, they form a QD array where their wave functions become highly extended, and their energy level becomes very small. Consequently, the energy bands formed by the QDs exhibit continuity rather than discrete levels,27 as shown in Fig. 4. This can be explained using the QD Bloch state, i.e., eikz, where the QD wavenumber ‘k’ lies between −π/L to π/L where L is the measure of one period (Lqd + Lbr). Within this range, certain energy intervals between the intrinsic barriers (IBs) are unoccupied by the QD energies, resulting in energy gaps. These gaps represent regions with zero density of states, where carriers cannot thermally depopulate from the conduction band (CB) to the intrinsic barrier due to the delocalization of wave functions in the ground state energy.11Fig. 4 particularly estimates the variation in miniband formation for interdot spacing and size of the QD. The width of the E0 miniband at Lqd = 1 nm is 403.6 eV; at Lqd = 3 nm it is 60.9 eV and it is lowered as the QD size decreases or the distance between vertically stacked QDs increases. For small QD sizes, the E1, E2 and E3 are so close that they overlap with the CB edge of the barrier material. Increasing the QD size decreases the gap between the ground states of the hole and electrons (IB-VB).
The inclusion of thin interlayers is crucial for absorbing excess indium and relieving compressive strain. These GaN interlayers need to strike a balance between being thick enough to be effective and thin enough to facilitate carrier transport through tunneling. The In0.5Ga0.5N QD layers should be thick enough to capture the incident light beam while remaining fully strained. Fig. 5(a) exhibits increasing interlayer effects on the electrical parameters of QD-IBSC. Critical thickness plays a very vital role in strain layers. The barrier height at higher indium concentrations causes current transport through tunneling. The tunneling current diminishes even when the barrier thickness remains constant. The thickness of the GaN interlayer decreases the power conversion efficiency, as shown in Fig. 5(b).
Fig. 5 (a) J–V characteristics and (b) efficiency of In0.5Ga0.5N/GaN QD-IBSC as a function of interlayers with QD layer = 20, Lqd = 3 nm, and Lbr = 3 nm, respectively. |
Fig. 6(a) depicts the relationship between the Jsc and η as a function of the number of QD layers. It is observed that the increasing slope of Jsc reaches a maximum value of 34.2 mA cm−2, while the efficiency peaks at 20 layers. Additionally, the Voc also experiences an increase in the proposed structure. During illumination, a significant number of photons are absorbed in the top n-layer. Due to the higher mobility of electrons compared to holes, the electrons are transported from the n-layer to the p-layer. The energy band bending depends on the interfaces and the electric field induced within the materials, specifically, GaN and In0.5Ga0.5N. The total induced polarization (Epolarization), as discussed in the device theory section, is reduced. Moreover, the built-in electric field (Ebuiltin) aligns in the same direction, facilitating the movement of carrier charges from the n-layer to the p-layer. This phenomenon contributes to the increased efficiency depicted in Fig. 6(b). Therefore, it can be concluded that the optimized configuration for maximum efficiency consists of up to 20 layers according to the proposed structure. It is worth noting that the literature suggests that increasing the number of QD layers introduces strain.5,38–40 However, the inclusion of i-GaN interlayers absorbs the strain generated at the interfaces, leading to an enhancement in both Jsc and η.
Fig. 6 (a) J–V characteristics and (b) efficiency of In0.5Ga0.5N/GaN QD-IBSC as a function of QD layers with Lqd = 3 nm, Lbr = 3 nm, respectively. |
Fig. 7(a) and (b) provide insights into the J–V characteristics and efficiency of the QD-IBSC with respect to different parameters. It is observed that both the J–V characteristics and efficiency show an increasing trend as the QD size increases. The maximum efficiency achieved is 45% under the conditions of Lbr = 3 nm, Lqd = 3 nm, and x = 0.5. For smaller QD sizes, the IB widens, leading to an increased carrier concentration and current. The Voc remains constant for two values, then increases, and becomes constant again. The size of the quantum dot significantly influences the performance of the QD-IBSC. Larger QD sizes result in a lower position of the IB, allowing more space for accumulating bands and the creation of additional IBs. However, there is a limit to the number of IBs that can be formed. In this particular device, three bands are created within the forbidden region, leading to the generation of additional electron–hole pairs and enhancing the device's efficiency. In Fig. 7(c), it is demonstrated that the energy is maximum for the smallest QD size. The wave functions of the electrons are localized within the QDs, and the quantum confinement effect causes discrete energy levels rather than continuous energy levels. The sub gap decreases as the QD size increases due to this quantum confinement effect within the QDs.
Fig. 8(a) shows that an increase in the interdot size leads to a decrease in the Jsc. This reduction is attributed to the localization of QDs as the Lbr increases. The weaker coupling of the wave functions results in individual QD behaviour, rather than a collective band behaviour. Additionally, the interdot size influences the width of the IB. Larger interdot sizes lead to higher sub-bandgap transition energies,53 resulting in lower overall efficiency. Fig. 8(b) illustrates the optimum efficiency of 49% obtained with specific parameters: Lqd of 3 nm, Lbr of 1 nm, and x of 0.5. It shows a decrease in efficiency, which drops abruptly and significantly affects the performance of the QD-IBSC. This decrease is primarily attributed to recombination rates, especially radiative recombination, becoming dominant for small interdot distances. The reduction in bandgap at an indium composition of 0.5 contributes to this effect. On the other hand, Fig. 8(c) demonstrates the relationship between interdot distance and energy. When QDs are stacked at specific distances and sizes, wave function delocalization occurs, forming an IB. The interdot distance variation affects the bandwidth of the IBs, where larger interdot distances result in narrower bandwidths and discrete energy levels. A summary of this work is presented in Table 2.
Fig. 8 (a) J–V characteristics, (b) efficiency and (c) E0, E1, E2 of In0.5Ga0.5N/GaN QD-IBSC as a function of distance between two QDs for x = 0.5, QD layers = 20, Lqd = 3 nm and WL = 1 nm. |
Concentration | QD layer | QD size (nm) | Interdot (nm) | J sc (mA cm−2) | V oc (V) | FF (%) | η (%) |
---|---|---|---|---|---|---|---|
0.5 | 5 | 3 | 3 | 20.27 | 2.44 | 74.39 | 36.669 |
10 | 28.4 | 2.45 | 57.14 | 39.76 | |||
15 | 33.35 | 2.464 | 60.79 | 46.95 | |||
20 | 34.1 | 2.451 | 57.5 | 48.01 |
This journal is © The Royal Society of Chemistry 2024 |