Open Access
13 September 2018 Optimization of nonhomogeneous indium-gallium-nitride Schottky-barrier thin-film solar cells
Tom H. Anderson, Akhlesh Lakhtakia, Peter B. Monk
Author Affiliations +
Abstract
A two-dimensional model was developed to simulate the optoelectronic characteristics of indium-gallium-nitride (InξGa1  −  ξN), thin-film, Schottky-barrier solar cells. The solar cells comprise a window designed to reduce the reflection of incident light, Schottky-barrier and ohmic front electrodes, an n-doped InξGa1  −  ξN wafer, and a metallic periodically corrugated backreflector. The ratio of indium to gallium in the wafer varies periodically throughout the thickness of absorbing layer of the solar cell. Thus, the resulting InξGa1  −  ξN wafer’s optical and electrical properties are made to vary periodically. This material nonhomogeneity could be physically achieved by varying the fractional composition of indium and gallium during deposition. Empirical models for indium nitride and gallium nitride were combined using Vegard’s law to determine the optical and electrical constitutive properties of the alloy. The nonhomogeneity of the electrical properties of the InξGa1  −  ξN aids in the separation of the excited electron–hole pairs, whereas the periodicities of optical properties and the backreflector enable the incident light to couple to multiple guided wave modes. The profile of the resulting charge-carrier-generation rate when the solar cell is illuminated by the AM1.5G spectrum was calculated using the rigorous coupled-wave approach. The steady-state drift-diffusion equations were solved using COMSOL, which employs finite-volume methods, to calculate the current density as a function of the voltage. Midband Shockley–Read–Hall, Auger, and radiative recombination rates were taken to be the dominant methods of recombination. The model was used to study the effects of the solar-cell geometry and the shape of the periodic material nonhomogeneity on efficiency. The solar-cell efficiency was optimized using the differential evolution algorithm.

1.

Introduction

Alloys of indium gallium nitride (InξGa1ξN) can be tailored to possess a wide range of bandgaps, from 0.70 to 3.42 eV, by varying the relative proportions of indium and gallium through the parameter ξ(0,1).1 Pure indium nitride (i.e., ξ=1) has a bandgap of 0.7 eV,2,3 whereas gallium nitride (i.e., ξ=0) has a bandgap of 3.42 eV. It should be noted that InξGa1ξN with high indium content (i.e., ξ0.3) currently suffers from poor electrical characteristics, background n-doping due to Fermi pinning above the conduction-band edge,4 and a bandgap that is greater than expected.5,6 These problems are exacerbated by p-doping of InξGa1ξN.7

Solar cells can be designed to use an in-built potential provided by a Schottky-barrier junction, which can occur at a metal/semiconductor interface.8,9 By partnering n-doped InξGa1ξN with a metal possessing a large work function ϕ—as opposed to, say, employing the more usual p-i-n junction—the problems associated with p-doping of the material are avoided. Furthermore, the deposition process is simplified, as only one dopant element is required. Hence, the reduced fabrication costs could offset the lower efficiencies of Schottky-barrier thin-film solar cells.

Theoretical studies,7,10 which corroborate an earlier experimental study,11 suggest that InξGa1ξN Schottky-barrier solar cells with relatively high efficiency could be designed. Anderson et al.12 investigated the efficiency of InξGa1ξN Schottky-barrier solar cells with periodic variation of the indium-to-gallium ratio. This involved the solution of both the frequency-domain Maxwell postulates in the optical regime and the carrier drift-diffusion equations using the commercial finite-element package COMSOL (V5.2a), in order to simulate the efficiencies of a variety of designs. The efficiency was found to increase significantly on the incorporation of periodic nonhomogeneity with a specific profile.

For the traditional amorphous-silicon p-i-n-junction solar cells, the incorporation of a periodically nonhomogeneous intrinsic layer (i.e., i layer), along with a metallic periodically corrugated backreflector (PCBR), can improve overall efficiency by up to 17%.13 For a Schottky-barrier solar cell made from InξGa1ξN, the inclusion of these features was shown to increase the total efficiency by up to 26.8%.12 In neither case, however, was comprehensive optimization of the design parameters conducted. The improvements seen are likely due to the following reasons:

  • i. The excitation of guided wave modes, including surface-plasmon-polariton waves1416 and waveguide modes,17 is made possible by the inclusion of the metallic PCBR.1821 Both types of phenomena intensify the optical electric field inside the photon-absorbing regions of the solar cell, which leads to an increase in the electron–hole pair generation rate.

  • ii. The combination of a periodically nonhomogeneous semiconductor and a PCBR enables the excitation of an increased number of guided wave modes.20,22,23 More pathways become available for the incident photons to be absorbed, thereby increasing the charge-carrier-generation rate.

  • iii. The drift-diffusion equations include terms pertaining to gradients in the electrical constitutive properties of the materials in the solar cell. The material nonhomogeneity will facilitate the separation of electrons and holes, and it may also suppress recombination.24,25

The aim of this paper is to expand on the previous work on InξGa1ξN by providing a comprehensive optimization of the device parameters in order to maximize efficiency. The optical calculations were undertaken using the rigorous coupled-wave approach (RCWA),16 whereas the electrical calculations were undertaken using COMSOL (V5.3a).26 Optical absorption could have been maximized if only optical models had been used, but the missing influence of the varying electrical properties would have made optimization of efficiency impossible.13 For example, if electrical modeling is omitted, the optical absorption can be maximized by minimizing the bandgap, but this would result in a solar cell with a small open-circuit voltage and therefore, quite likely, low efficiency.27

The plan of this paper is as follows. The design of the chosen solar cell is summarized in Sec. 2.1, with further details available elsewhere.13,28 The optical and electrical constitutive properties used in the simulation are presented in Sec. 2.2, and the computational models employed are described in Sec. 2.3. Numerical results are presented in Sec. 3. Closing remarks are presented in Sec. 4.

2.

Summary of the Two-Dimensional (2-D) Model

2.1.

Solar-Cell Design

The model is described in detail in Ref. 12. For the sake of completeness, a summary is included here. The simulated Schottky-barrier solar cell is schematically shown in Fig. 1. As the solar cell is translationally invariant in the y direction, the simulation is reduced to two dimensions (i.e., the xz plane) without approximation. In the remainder of this paper, the term width refers to the extent along the x axis, whereas the term thickness refers to the extent along the z axis.

Fig. 1

Schematic illustration of the reference unit cell R of the Schottky-barrier solar cell.

JPE_8_3_034501_f001.png

The solar cell comprises a planar antireflection window, a layer containing electrodes, a wafer of InξGa1ξN, and a layer containing a backreflector. Each of these layers is of uniform thickness. Insolation occurs at normal incidence to the solar cell through the antireflection window, with the wave vector of the incident light aligned with the positive z axis.

The device is periodic along the x axis with period Lx and has a thickness Lw+Lc+Lz+Lr. The reference unit cell of the device is the region R={(x,z)|Lx/2<x<Lx/2,LwLc<z<Lz+Lr}. A planar antireflection window, made from flint glass,29 occupies the region LwLc<z<Lc in R. The region 0<z<Lz is occupied by n-doped InξGa1ξN, forming both Schottky-barrier and ohmic junctions with the metal electrodes in the region Lc<z<0 in R. For optical calculations, the ohmic contact and backreflector were assumed to be silver,30 whereas the Schottky-barrier contact was assumed to be platinum.31 It must be noted that the electrical properties of silver were not used. The Schottky-barrier electrode of width Ls is centered in R at x=0 along the x axis. The two ohmic electrodes, each of width Lo/2, are centered at x=±(LxLo/2)/2 in R. Note that Lo+Ls<Lx and so the electrodes are electrically isolated. It should also be noted that, due to the periodicity of the design, there are an equal number of ohmic and Schottky-barrier electrodes in the solar cell. The gaps between the electrodes, Lc<z<0 and either (LxLo)/2<x<Ls/2 or Ls/2<x<(LxLo)/2, are occupied by flint glass.

The region Lz<z<Lz+Lr in R contains both silver and flint glass. The backreflector is made of two silver slabs welded together. The first slab is optically thick and occupies the region Lz+LrLm<z<Lz+Lr. The second slab occupies the region {(x,z)|ζLx/2<x<ζLx/2,Lz+Ld<z<Lz+LrLm}R, where ζ(0,1) is the duty cycle. Thus, Lg=Lr(Ld+Lm) is the corrugation height. The remainder of the region Lz<z<Lz+Lr in R is occupied by flint glass which electrically insulates silver from InξGa1ξN.

Absorption of the normally incident solar flux with AM1.5G spectrum32 was calculated by solving the frequency-domain Maxwell postulates.16 The semiconductor charge-carrier drift-diffusion equations model the spatial distributions of the electron density and the hole density.33,34 Because of the nonhomogeneity of the semiconductor (i.e., InξGa1ξN), the effective dc electric field acting on

  • a. electrons includes a contribution from gradients in the electron affinity, and

  • b. holes includes contributions from gradients in both the electron affinity and the bandgap.

Direct, midgap Shockley–Read–Hall, and Auger recombination were all included in our simulation. The current density J, which is averaged over either the Schottky-barrier electrode (or, identically, both of the ohmic electrodes), was calculated for a range of values of the external biasing voltage Vext.

2.2.

Material Parameters

For a specific bandgap Eg0, the fractional concentration of indium ξ is given as

Eq. (1)

ξ(z)=b+(EgGaNEgInN)4b[Eg0(z)EgGaN]+(b+EgGaNEgInN)22b,
where the bowing parameter b=1.43  eV,12,35 and the bandgaps EgInN=0.7  eV of InN and EgGaN=3.42  eV of GaN.

2.2.1.

Optical parameters

The optical refractive index nopt of InξGa1ξN depends on the free-space wavelength λ0 and was modeled using two equations. The real part of nopt is provided by the Adachi model as7

Eq. (2)

Re{nopt(ξ,λ0)}=Re(AA(ξ){[Eg0(ξ)Eγ(λ0)]2[21+Eγ(λ0)Eg0(ξ)1Eγ(λ0)Eg0(ξ)]+BA(ξ)}),
where AA(ξ) and BA(ξ) are interpolated from the corresponding parameters for InN and GaN provided in Table 1. The photon energy is denoted by Eγ(λ0)=2πc0/λ0, where =1.054571800×1034  m2kgs1 is the reduced Planck constant and c0=2.99792485×108  ms1 is the speed of light in free space.

Table 1

Electronic data used for GaN and InN. The composition of InξGa1−ξN was estimated using Eq. (1), with linear interpolation used to estimate data for the semiconductor-filled region 0<z<Lz with bandgaps not presented here in all cases, except for the electron affinity χ0 which uses Eq. (6).

SymbolUnitGaNInN
BandgapEg*eV3.420.7
Electron affinityχ0eV4.15.6
Density of states (conduction band)NCcm32.3×10189.1×1017
Density of states (valence band)NVcm34.6×10195.3×1019
Electron mobility 1μn(1)cm2V1s12951030
Electron mobility 2μn(2)cm2V1s1146014150
Caughey–Thomas doping power (electrons)δn0.710.6959
Caughey–Thomas critical doping density (electrons)Nncritcm37.7×10162.07×1016
Hole mobility 1μp(1)cm2V1s133
Hole mobility 2μp(2)cm2V1s1170340
Caughey–Thomas doping power (holes)δp22
Caughey–Thomas critical doping density (holes)Npcritcm31×10188×1017
Auger recombination factor (electrons)Cncm6s11.5×10301.5×1030
Auger recombination factor (holes)Cpcm6s11.5×10301.5×1030
Direct recombination factorCradcm3s11.1×1082×1010
Slotboom reference energyErefeV9×1039×103
Slotboom reference concentrationNrefcm31×10171×1017
Conduction-band fractionαoc0.90.9
Adachi refractive-index parameter AAAA9.3113.55
Adachi refractive-index parameter BABA3.032.05

The imaginary part of the optical refractive index nopt was modeled as

Eq. (3)

Im{nopt(ξ,λ0)}=λ04παopt(ξ,λ0).
The absorption coefficient αopt was modeled as7

Eq. (4)

αopt(ξ,λ0)=105C(ξ)[Eγ(ξ)Eg0(ξ)]+D(ξ)[Eγ(ξ)Eg0(ξ)]2  nm1,
wherein the constants

Eq. (5)

C(ξ)=(3.52518.28ξ+40.22ξ237.52ξ3+12.77ξ4)  eV1D(ξ)=(0.6651+3.616ξ2.460ξ2)  eV2},
come from interpolation of parameters given by Brown et al.36

2.2.2.

Electrical parameters

The Schottky-barrier work function matched that of platinum in our simulations. Thus, ϕ=5.93  eV.37 For a specific value of ξ, the electrical properties were modeled using either quadratic or linear (i.e., Vegard’s law38) interpolation of data for InN and GaN.

The electron affinity

Eq. (6)

χ0(z)=ξ(z)χInN+[1ξ(z)]χGaNbξ(z)[1ξ(z)],z(0,Lz),
was modeled using the same quadratic fit as the bandgap, where χInN and χGaN are the electron affinities of InN and GaN, respectively. All other parameters presented in the first column of Table 1 were modeled using Vegard’s law of linear interpolation.

The narrowing of the bandgap associated with doping was incorporated through the Slotboom model.39 An empirical low-field mobility model—called either the Caughey–Thomas7 or the Arora26 mobility model—was used for the variations of the electron mobility and the hole mobility. Details of these models are available elsewhere.7,12

The bandgap of InξGa1ξN was taken to vary periodically in the thickness direction of the solar cell, described as

Eq. (7)

Eg0(z)=Eg*A(1{12[sin(2πzLp2πϕ)+1]}α),
where Eg* is the baseline (maximum) bandgap, A is the amplitude, Lp=Lz/κ˜ is the period with the bandgap nonhomogeneity ratio κ˜>0, ϕ is a phase shift, and α is a shaping parameter.

2.3.

Computational Model

The problem of calculating the total efficiency of the solar cell was decoupled into two separate calculations. First, the RCWA16 was used to calculate the spectrally integrated photon-absorption rate, which is ideally equal to the charge-carrier-generation rate. This was then coupled with a 2-D finite-element electronic model that was implemented in the COMSOL (V5.3a) software package.26 In the remainder of this paper, terms in small capitals are COMSOL terms.

2.3.1.

Differential evolution algorithm

The differential evolution algorithm (DEA)40 was used to optimize the solar-cell design. Given N¯ parameters in the optimization problem, an initial population P0 of NP members in the parameter-search space SRN¯ was chosen randomly in [0.5,1] with a uniform distribution. After the cost function C:SR of the problem had been evaluated at each of these points, the DEA produced a new population P1 of NP members in the parameter search space S to test. This process was iterated until the change in absorptance was less than 1% or until a set amount of time had passed.

By representing the current population Pj as a matrix with each of its NP columns being vectors in S, the next population can be written as

Eq. (8)

Pj+1=M1[Pj+F(v*1Pj)+F(P˜j(1)P˜j(2))]+M2Pj,
where v*S is the optimal parameter vector found at that stage; 1 is the vector of 1’s; is the outer product; P˜j(1) and P˜j(2) are versions of Pj where the columns have been randomly interchanged; the parameter F(0,2] is the step size to be taken by the DEA at each iteration; M1 and M2=11M1 are filter matrices of 1’s and 0’s generated by the DEA, with M1 having approximately a fraction CR of 1’s, where CR is termed the crossover fraction; and ∘ is the Shur product (elementwise multiplication).

The cost function C was taken to be either the efficiency η or the optical short-circuit current density JSCOpt. The population number was set to NP=30, the crossover fraction was set to CR=0.6, and the step size F was set to be randomly distributed in [0.5, 1] uniformly. Allowing F to vary randomly with each iteration has been termed dither, and has been shown to improve convergence for many problems.41

2.3.2.

RCWA algorithm

Suppose that the face z=LcLw of the solar cell is illuminated by a normally incident plane wave with electric field phasor

Eq. (9)

E̲inc(x,z,λ0)=Eo2(x^̲+y^̲)exp(i2πλ0z).
As a result of the metallic backreflector being periodically corrugated, the x-dependences of the electric and magnetic field phasors must be represented by Fourier series everywhere as

Eq. (10)

E̲(x,z,λ0)=n=n=e̲(n)(z,λ0)exp(iκ(n)x),|z|<,|x|<,
and

Eq. (11)

H̲(x,z,λ0)=n=n=h̲(n)(z,λ0)exp(iκ(n)x),|z|<,|x|<,
where i=1, κ(n)=n(2πLx), and e̲(n)=ex(n)x^̲+ey(n)y^̲+ez(n)z^̲ as well as h̲(n)=hx(n)x^̲+hy(n)y^̲+hz(n)z^̲ are Fourier coefficients. Likewise, the optical permittivity ϵ(x,z,λ0)=ϵ0nopt2(x,z,λ0) everywhere has to be represented by the Fourier series

Eq. (12)

ϵ(x,z,λ0)=n=n=ϵ(n)(z,λ0)exp(iκ(n)x),|z|<,|x|<,
where ϵ0 is the permittivity of free space.

Computational tractability requires truncation so that n{Nt,,Nt}, Nt0. Column vectors

Eq. (13)

e˘σ(z,λ0)=[eσ(Nt)(z,λ0),eσ(Nt+1)(z,λ0),,eσ(Nt1)(z,λ0),eσ(Nt)(z,λ0)]T,σ{x,y,z},
and

Eq. (14)

h˘σ(z,λ0)=[hσ(Nt)(z,λ0),hσ(Nt+1)(z,λ0),,hσ(Nt1)(z,λ0),hσ(Nt)(z,λ0)]T,σ{x,y,z},
were set up, the superscript T denoting the transpose. Furthermore, the matrices

Eq. (15)

K˘=diag[κ(Nt),κ(Nt+1),,κ(Nt1),κ(Nt)]
and

Eq. (16)

ϵ˘(z,λ0)=[ϵ(0)(z,λ0)ϵ(1)(z,λ0)ϵ(2Nt+1)(z,λ0)ϵ(2Nt)(z,λ0)ϵ(1)(z,λ0)ϵ(0)(z,λ0)ϵ(2Nt+2)(z,λ0)ϵ(2Nt+1)(z,λ0)ϵ(2Nt1)(z,λ0)ϵ(2Nt2,λ0)(z,λ0)ϵ(0)(z,λ0)ϵ(1)(z,λ0)ϵ(2Nt)(z,λ0)ϵ(2Nt1)(z,λ0)ϵ(1)(z,λ0)ϵ(0)(z,λ0)]
were set up. The frequency-domain Maxwell curl postulates then yielded the matrix ordinary differential equation

Eq. (17)

ddzf˘(z,λ0)=iP˘(z,λ0)f˘(z,λ0),
where the 4(2Nt+1)-column vector

Eq. (18)

f˘(z,λ0)=[e˘x(z,λ0)e˘y(z,λ0)h˘x(z,λ0)h˘y(z,λ0)]
and the 4(2Nt+1)×4(2Nt+1) matrix

Eq. (19)

P˘(z,λ0)=ω[0˘0˘0˘μ0I˘0˘0˘μ0I˘0˘0˘ϵ˘(z,λ0)0˘0˘ϵ˘(z,λ0)0˘0˘0˘]+1ω[0˘0˘0˘K˘[ϵ˘(z,λ0)]1K˘0˘0˘0˘0˘0˘μ01K˘K˘0˘0˘0˘0˘0˘0˘]
contains μ0 as the permeability of free space, 0˘ as the (2Nt+1)×(2Nt+1) null matrix, and I˘ as the (2Nt+1)×(2Nt+1) identity matrix.

The solar cell was discretized along the z axis.16 This effectively broke the domain R into a cascade of slices. Each slice was homogeneous along the z axis, but it was either homogeneous or periodically nonhomogeneous along the x axis. Equation (17) was then solved using a stepping algorithm to give an approximation for f˘ in each slice. Finally, the Fourier coefficients of the z components of the electric and magnetic field phasors were obtained from

Eq. (20)

e˘z(z,λ0)=ω1[ϵ˘(z,λ0)]1K˘h˘y(z,λ0)h˘z(z,λ0)=(ωμ0)1K˘e˘y(z,λ0)}.
Thus, the electric field phasor was determined throughout the solar cell.

The spectrally integrated number of absorbed photons per unit volume per unit time is given as

Eq. (21)

Nph(x,z)=η0λ0minλ0maxIm{ϵ(x,z,λ0)}|E̲(x,z,λ0)Eo|2S(λ0)dλ0,
where η0=μ0/ϵ0 is the intrinsic impedance of free space and S(λ0) is the AM1.5G solar spectrum.32 With the assumption that the absorption of every photon in the InξGa1ξN layer releases an electron–hole pair, the charge-carrier-generation rate can be calculated as

Eq. (22)

G(x,z)=Nph(x,z),
everywhere in that layer. Whereas λ0min=350  nm, λ0max=1240  nmeV/Eg,min is the maximum wavelength that can contribute to the optical short-circuit current density

Eq. (23)

JSCOpt=qe1Lx0LzLx/2Lx/2G(x,z)dxdz,
where Eg,min (in eV) is the minimum bandgap present in the solar cell and qe=1.6×1019  C is the elementary charge.

The integral in Eq. (21) was approximated using the trapezoidal rule43 with sampling at wavelengths spaced at 2-nm intervals. The integral in Eq. (23) was also approximated using the trapezoidal rule. The sampling resolution was regular in both directions, with δx=Lx/100, and δz=Lz/200.

The optical short-circuit current density provides a rough benchmark for the device efficiency and is used by many optics researchers who simulate solar cells.42 However, as recombination is neglected, JSCOpt is necessarily larger than the actually attainable short-circuit current density JSC, which is the electronically simulated current density that flows when the solar cell is illuminated and no external bias is applied (i.e., when Vext=0). For the results presented here, calculating only JSCOpt would have been inadequate as the electrical constitutive properties were also significantly varied.

2.3.3.

Adaptive-Nt implementation

The calculated value of JSCOpt varies with Nt. An adaptive method was implemented to estimate when Nt is sufficiently large. Equation (21) was evaluated using the trapezoidal rule.43 At the first value of λ0 sampled, λ0=λ0min, JSCOpt was calculated with Nt1 and Nt2=Nt1+2. If the magnitude ΔJSCOpt=|JSCOpt(Nt2)JSCOpt(Nt1)| of the difference between the two calculated values of JSCOpt was greater than a specified tolerance, then Nt1 was set equal to Nt2 and Nt2 was increased by two. This iterative procedure was continued until ΔJSCOpt was less than the specified tolerance for two subsequent comparisons. A maximum value of Nt100 was enforced in order to force to the calculation to terminate within a reasonable duration. After a successful calculation, the next value of λ0 was selected, and Nt1=2Nt2/4 and Nt2=Nt1+2 were chosen for the next calculation, where is the ceiling function.

2.3.4.

Solution of drift-diffusion equations

The charge-carrier-generation rate G(x,z), calculated using the RCWA, was processed using an external Matlab™ code and then used as the input, via user-defined generation, for the COMSOL electrical model. Recombination was incorporated via Auger, direct and trap-assisted (midgap Shockley–Read–Hall) phenomena, using parameters as provided in Table 1.

Due to the symmetry in the simulation, only the right half of the domain R (i.e., 0xLx/2) was electrically simulated, with insulator interfaces applied down the lines of symmetry.

Fermi–Dirac carrier statistics were employed along with a finite volume (constant shape function) discretization, as this inherently conserves current throughout the solar cell.26 COMSOL utilizes a Scharfetter–Gummel upwinding scheme. The free triangular, Delaunay mesh has a maximum element size of 15 nm. Further details can be found in Ref. 12.

The semiconductor module of COMSOL was used to calculate the current densities flowing through the ohmic, and therefore also Ideal Schottky, electrodes. A prescribed external voltage Vext was applied between these electrodes. The current density flowing through the Schottky-barrier electrode was modeled using thermionic currents, with standard Richardson coefficients of An*=110  AK2cm2 and Ap*=90  AK2cm2.7,26 By sweeping Vext from 0 V up to a value where JSC drops to zero, the JSC-Vext curve was produced. This enabled calculation of the maximum attainable value of the efficiency η.

3.

Numerical Simulation Results

3.1.

Optimization Study

The defined problem has 15 parameters, shown in Table 2, all of which influence the charge-carrier-generation rate G(x,z) and the efficiency η of the solar cell. The choice of four of these parameters was guided by either physical constraints or they were found to strongly affect the optical response of the solar cell while affecting the electrical characteristics only indirectly, i.e., by changing the spatial profile of the charge-carrier-generation rate.

  • A smaller value of Ld was seen to further increase the number of photons absorbed in the InξGa1ξN; however, a minimum value of Ld=40  nm was chosen: (i) to model a solar cell where the electrical insulation of the backreflector is maintained, and (ii) to provide a suitable surface for InξGa1ξN deposition.

  • The effect of the minimum thickness Lm of the metallic layer has been omitted from the results. This is because, given a sufficiently thick silver layer, only a trivial amount of the incident light will be transmitted through the solar cell. With Lm chosen to be more than twice the penetration depth of silver across the majority of the AM1.5G spectrum, changes in solar-cell performance for modest perturbations in Lm from this value are minimal.

  • A smaller value of Lc was seen to further increase the number of photons absorbed in the InξGa1ξN. A minimum value of Lc=50  nm was chosen so the contacts are more physically realistic: thinner contacts would likely suffer from uneven deposition and exhibit increased series resistance.

  • A strong peak in efficiency was seen when Lw75  nm, irrespective of the other parameters. This corresponds to the glass coating acting as a quarter-wavelength antireflection coating.

For all data reported here, the values Ld=50  nm, Lm=100  nm, Lc=50  nm, and Lw=75  nm were fixed.

Table 2

Summary of parameters used for simulation.

ParameterNameSymbolValue/Range
Device periodLx[100,1500] nm
Thickness of n-doped InξGa1ξN layerLz{20,800} nm
Minimum thickness of insulation windowLd40 nm
Corrugation heightLg[1,200] nm
Corrugation duty cycleζ[0.01,0.99]
Minimum thickness of metalLm100 nm
Electrode-region thicknessLc50 nm
Ohmic-electrode widthLo50+ nm
Schottky-barrier electrode widthLs50+ nm
Antireflection-window thicknessLw75 nm
Pre-doping bandgapEg*[0.7,3] eV
Bandgap nonhomogeneity amplitudeA[0,3.42Eg*]  eV
Bandgap nonhomogeneity phaseϕ[0, 1]
Bandgap nonhomogeneity shaping parameterα[24,28]
Bandgap nonhomogeneity ratioκ˜[0.01, 5.5]

The remaining 11 parameters were allowed to vary within the following ranges: Lx[100,1500]  nm, Lz[20,800]  nm, Lg[1,200]  nm, ζ[0.01,0.99], Lo[50,LxLs]  nm, Ls[50,LxLo]  nm, Eg*[0.7,3]  eV, A[0,3.42Eg*]  eV, ϕ[0,1), α[24,28], and κ˜[0.01,5.5]. These ranges, along with the chosen values of the optical parameters, are summarized in Table 2.

The maximum obtained efficiency η=11.13% was found at: Lx=694  nm, Lg=125  nm, ζ=0.012, Lo/Lx=0.16, Ls/Lx=0.01, A=0.74025, κ˜=3.05, ϕ=0.711, α=13.3614, Eg*=1.17 eV, and Lz=735 nm. These values were computed after 10 DEA population evolutions and provide an estimate for the maximum efficiency, as well as allowing the upcoming interpretation of the results. Further iterations would provide greater confidence in the conclusions, at the cost of greater computation time.

3.1.1.

Results of optimization study

Figure 2 shows η plotted against JSCOpt, with each data point corresponding to one DEA population member. The maximum value of JSCOpt was calculated to be 37.042  mAcm2, but the maximum efficiency was observed to occur at JSCOpt=31.442  mAcm2. Whereas a larger value of JSCOpt increases the likelihood of obtaining a high value of η, the former does not predict the latter. Indeed, the designs with values of JSCOpt>25  mAcm2 produced efficiencies ranging from <1% up to over 11% in Fig. 2(a). In part, this is caused by a device with a large optical short-circuit current density JSCOpt not automatically producing a large short-circuit current density JSC. This behavior is shown by the droop at higher values of JSCOpt in Fig. 2(b): as the optical short-circuit current density increases, recombination in the solar cell increases. These observations highlight the importance of conducting full optoelectronic simulations when modeling solar cells, especially when parameters with a strong electrical effect, such as the bandgap, are allowed to vary.

Fig. 2

(a) Comparison of the solar-cell efficiency η to the optical short-circuit current density JSCOpt, showing that an increase in the latter does not necessarily translate into an increase in the former. (b) Relationship between optical short-circuit current density and simulated short-circuit current density. For all data points, Ld=50  nm, Lm=100  nm, Lc=50  nm, and Lw=75  nm were fixed, but the other parameters were varied.

JPE_8_3_034501_f002.png

3.1.2.

Details of optimization study

The results from the optimization study show how the different parameters affect the efficiency η of the solar cell. Figures 3 and 4 show the projections of the entire parameter space onto the sets of axes containing the efficiency and each of the optimization parameters. Parameters that have a strong effect on the efficiency have most of their points strongly clustered around the design with the highest efficiency. Each point is colored with the value of JscOpt.

Fig. 3

Results projected onto the plane containing η and (a) Lz, (b) Eg*, (c) A, (d) κ˜, (e) ϕ, (f) α, (g) Lo/Lx, and (h) Ls/Lx for a solar cell with Ld=50  nm, Lm=100  nm, Lc=50  nm, and Lw=75  nm. The large points highlight the location of the device with the maximum efficiency.

JPE_8_3_034501_f003.png

Fig. 4

(a) Bandgap of most efficient design. Rest, as Fig. 3 except that the results are projected onto the plane containing η and (b) Lx, (c) Lg, and (d) ζ.

JPE_8_3_034501_f004.png

In Fig. 3(a), the thickness of the solar cell is seen to have a moderate effect on the resulting efficiency. The peak visible around Lz700  nm sees light clustering, with some reasonably efficient solar cells with η>7%, also produced when Lz is less than half the optimal value. Figure 3(b) shows that Eg* strongly affects the resulting solar cell efficiency. The peak around 1.2 eV lies in the region predicted by the Shockley–Queisser limit. Solar cells with narrower bandgaps do produce solar cells with higher optical short-circuit current densities, but the reduction in open-circuit voltage dramatically reduces the efficiency. Figure 3(c) shows that a nonzero amplitude A can substantially increase the efficiency. While the structure of the peak is not well resolved, all solar cells with A<0.5  eV had efficiencies less than half that of the maximum attained efficiency. Further increase of A beyond 1 eV slowly reduces the attainable efficiency. Figures 3(g) and 3(h) show that a small Schottky-barrier electrode and a slightly larger ohmic electrode are required to maximize the efficiency.

A previous study12 has suggested that optimal values of κ˜ are integer multiples of 1.5, which conclusion is not contradicted by the data; see Fig. 3(d). The strong peak around ϕ=0.75 in Fig. 3(e) is also in line with previous work.12,13 At this value of ϕ, a wide bandgap is produced near to the electrodes. This seems of paramount importance for producing high efficiency solar cells with nonhomogeneous bandgaps. The best bandgap profile is shown in Fig. 4(a).

Finally, in Fig. 4, the effects of the PCBR are shown. A period of Lx>500  nm is seen to significantly increase solar cell efficiency. This because the majority of short-wavelength light is absorbed far from the PCBR and so scattering effects are minimal. The grating amplitude and duty cycle are not seen to have strong effects on the solar-cell efficiency—the points do not cluster strongly in Figs. 4(c) and 4(d).

3.2.

Detailed Study

The simulated values of the major variables for the highest-efficiency device are shown in Fig. 5. The first two subfigures, 5(a) and 5(b), show the generation rate and the recombination rate within the solar cell when the external voltage is zero. Light is incident from below in Fig. 5. The bands of low generation and low recombination at z{230,570,710}  nm correspond to the locations where the bandgap perturation is large, i.e., the bandgap is much wider here. The majority of the photons absorbed in the first 200 nm of the solar cell are collected before recombining, whereas the majority of those absorbed in the rear 500 nm recombine. The exception to this trend are the photons that are absorbed where the bandgap peaks. These are quickly swept out of these regions by the effective electric field produced by the gradient in the bandgap and electron affinity.

Fig. 5

Properties of the optimal solar cell at short-circuit condition: (a) generation rate in log10(cm3s1), (b) recombination rate in log10(cm3s1), (c) current density in log10(mAcm3), (d) current density in z-direction in log10(mAcm3), (e) electron density in log10(cm3), and (f) hole density in log10(cm3).

JPE_8_3_034501_f005.png

Figures 5(c) and 5(d) show the current density produced in the solar cell. By comparing the two figures, it is seen that in the region of the narrow ohmic contact, at the outer edges of the plots, the current density is strongly perpendicular to the contact. The current density in the vicinity of the Schottky barrier (at the center of the plots) is lower, but it is also perpendicular to the contact. The current density toward the back of the solar cell is dramatically lower, supporting the earlier analysis that the majority of the excited carriers in this region recombine.

Figures 5(e) and 5(f) show the electron and hole densities at the short-circuit condition. The electrons are the majority carriers in the majority of the solar cell. In the vicinity of the Schottky barrier, there is a high concentration of holes. In the regions where the bandgap is large, both carrier densities are very low, which has the effect of drastically reducing recombination in these areas. Unfortunately, it also acts to limit current flow across these bands: it may therefore be beneficial in future studies to include lower-bandgap pathways to increase charge-carrier extraction from further back in the solar cell.

Finally, Fig. 6 shows the resulting JV curve for the optimal device. The short-circuit current is 21.79  mAcm2, the open-circuit voltage is 0.683 V, and the fill factor is 0.7479.

Fig. 6

The resulting JV curve for the most efficient solar cell design.

JPE_8_3_034501_f006.png

4.

Closing Remarks

A combined optoelectronic model was developed to enable to optimization of InξGa1ξN-based Schottky-barrier solar cells. These solar cells possessed a PCBR, a layer of InξGa1ξN, metallic electrodes, and an antireflection coating. The optimization was conducted using the DEA. The AM1.5G solar spectrum was used to illuminate the solar cell at normal incidence.

With a solely optical model, it was shown that the optical short-circuit current density JSCOpt of the design is strongly dependent on the thicknesses of the materials that are applied to the surface. A 75-nm thick layer of flint glass acts as a quarter-wavelength antireflection coating,44,45 maximizing JSCOpt. Minimizing the thickness of the front metallic electrodes also maximizes JSCOpt by reducing reflection.

An optimization study, using the DEA and a full optoelectronic model, produced a design for an InξGa1ξN-based Schottky-barrier solar cell with a simulated efficiency of 11.13%. This design included a periodically nonhomogeneous bandgap, with just over three full periods. The minimum bandgap was 1.17 eV and the maximum bandgap was 1.91 eV. The phase of the periodic nonhomogeneity was such that the InξGa1ξN close to the electrodes has a wide bandgap.

While experimental work is needed to test the veracity of the models employed, it has been shown that InξGa1ξN-based Schottky-barrier solar cells with a high efficiency may be producible if a periodic material nonhomogeneity is included.

Acknowledgments

This paper was based in part on a paper entitled, “Optimal indium-gallium-nitride Schottky-barrier thin-film solar cells,” presented at the SPIE Conference, “Next Generation Technologies for Solar Energy Conversion,” held August 5–11, 2017, in San Diego, California, USA.46 The authors thank F. Ahmad (Pennsylvania State University) for assistance with Figs. 25. The research of T.H. Anderson and P.B. Monk was partially supported by the US National Science Foundation under Grant No. DMS-1619904. The research of A. Lakhtakia was partially supported by the US National Science Foundation under Grant No. DMS-1619901. A. Lakhtakia also thanks the Charles Godfrey Binder Endowment at the Pennsylvania State University for ongoing support of his research.

References

1. 

J. Wu et al., “Superior radiation resistance of InξGa1ξN alloys: full-solar-spectrum photovoltaic material system,” J. Appl. Phys., 94 (10), 6477 –6482 (2003). https://doi.org/10.1063/1.1618353 JAPIAU 0021-8979 Google Scholar

2. 

J. Wu et al., “Unusual properties of the fundamental band gap of InN,” Appl. Phys. Lett., 80 (21), 3967 –3969 (2002). https://doi.org/10.1063/1.1482786 APPLAB 0003-6951 Google Scholar

3. 

Y. Saito et al., “Growth temperature dependence of indium nitride crystalline quality grown by RF-MBE,” Phys. Status Solidi B, 234 (3), 796 –800 (2002). https://doi.org/10.1002/1521-3951(200212)234:3<796::AID-PSSB796>3.0.CO;2-H Google Scholar

4. 

S. X. Li et al., “Fermi-level stabilization energy in group III nitrides,” Phys. Rev. B, 71 (16), 161201 (2005). https://doi.org/10.1103/PhysRevB.71.161201 Google Scholar

5. 

K. Osamura et al., “Fundamental absorption edge in GaN, InN and their alloys,” Solid State Commun., 11 (5), 617 –621 (1972). https://doi.org/10.1016/0038-1098(72)90474-7 SSCOA4 0038-1098 Google Scholar

6. 

T. Yodo et al., “Strong band edge luminescence from InN films grown on Si substrates by electron cyclotron resonance-assisted molecular beam epitaxy,” Appl. Phys. Lett., 80 (6), 968 –970 (2002). https://doi.org/10.1063/1.1450255 APPLAB 0003-6951 Google Scholar

7. 

S. O. S. Hamady, A. Adaine and N. Fressengeas, “Numerical simulation of InGaN Schottky solar cell,” Mater. Sci. Semicond. Process., 41 (1), 219 –225 (2016). https://doi.org/10.1016/j.mssp.2015.09.001 MSSPFQ 1369-8001 Google Scholar

8. 

N. K. Swami, S. Srivastava and H. M. Ghule, “The role of the interfacial layer in Schottky-barrier solar cells,” J. Phys. D Appl. Phys., 12 (5), 765 –771 (1979). https://doi.org/10.1088/0022-3727/12/5/018 Google Scholar

9. 

J. P. Colinge and C. A. Colinge, Physics of Semiconductor Devices, Kluwer Academic, New York (2002). Google Scholar

10. 

P. Mahala et al., “Metal/InGaN Schottky junction solar cells: an analytical approach,” Appl. Phys. A, 118 (4), 1459 –1468 (2015). https://doi.org/10.1007/s00339-014-8910-5 Google Scholar

11. 

J.-J. Xue et al., “Au/Pt/InGaN/GaN heterostructure Schottky prototype solar cell,” Chin. Phys. Lett., 26 (9), 098102 (2009). https://doi.org/10.1088/0256-307X/26/9/098102 CPLEEU 0256-307X Google Scholar

12. 

T. H. Anderson, T. G. Mackay and A. Lakhtakia, “Enhanced efficiency of Schottky-barrier solar cell with periodically nonhomogeneous indium gallium nitride layer,” J. Photonics Energy, 7 (1), 014502 (2017). https://doi.org/10.1117/1.JPE.7.014502 Google Scholar

13. 

T. H. Anderson et al., “Combined optical-electrical finite-element simulations of thin-film solar cells with homogeneous and nonhomogeneous intrinsic layers,” J. Photonics Energy, 6 (2), 025502 (2016). https://doi.org/10.1117/1.JPE.6.025502 Google Scholar

14. 

L. M. Anderson, “Parallel-processing with surface plasmons, a new strategy for converting the broad solar spectrum,” in Proc. 16th IEEE Photovoltaic Specialists Conf., 371 –377 (1982). Google Scholar

15. 

L. M. Anderson, “Harnessing surface plasmons for solar energy conversion,” Proc. SPIE, 408 172 –178 (1983). https://doi.org/10.1117/12.935723 PSISDG 0277-786X Google Scholar

16. 

Jr. J. A. Polo, T. G. Mackay and A. Lakhtakia, Electromagnetic Surface Waves: A Modern Perspective, Elsevier, Waltham, Massachusetts (2013). Google Scholar

17. 

T. Khaleque and R. Magnusson, “Light management through guided-mode resonances in thin-film silicon solar cells,” J. Nanophotonics, 8 (1), 083995 (2014). https://doi.org/10.1117/1.JNP.8.083995 1934-2608 Google Scholar

18. 

P. Sheng, A. N. Bloch and R. S. Stepleman, “Wavelength-selective absorption enhancement in thin-film solar cells,” Appl. Phys. Lett., 43 579 –581 (1983). https://doi.org/10.1063/1.94432 APPLAB 0003-6951 Google Scholar

19. 

C. Heine and R. H. Morf, “Submicrometer gratings for solar energy applications,” Appl. Opt., 34 2476 –2482 (1995). https://doi.org/10.1364/AO.34.002476 APOPAI 0003-6935 Google Scholar

20. 

L. Liu et al., “Experimental excitation of multiple surface-plasmon-polariton waves and waveguide modes in a one-dimensional photonic crystal atop a two-dimensional metal grating,” J. Nanophotonics, 9 (1), 093593 (2015). https://doi.org/10.1117/1.JNP.9.093593 1934-2608 Google Scholar

21. 

M. Solano et al., “Optimization of the absorption efficiency of an amorphous-silicon thin-film tandem solar cell backed by a metallic surface-relief grating,” Appl. Opt., 52 966 –979 (2013). https://doi.org/10.1364/AO.52.000966 APOPAI 0003-6935 Google Scholar

22. 

F. Ahmad et al., “On optical-absorption peaks in a nonhomogeneous thin-film solar cell with a two-dimensional periodically corrugated metallic backreflector,” J. Nanophotonics, 12 (1), 016017 (2018). https://doi.org/10.1117/1.JNP.12.016017 1934-2608 Google Scholar

23. 

M. Faryad and A. Lakhtakia, “Enhancement of light absorption efficiency of amorphous-silicon thin-film tandem solar cell due to multiple surface-plasmon-polariton waves in the near-infrared spectral regime,” Opt. Eng., 52 087106 (2013). https://doi.org/10.1117/1.OE.52.8.087106 Google Scholar

24. 

M. I. Kabir et al., “Amorphous silicon single-junction thin-film solar cell exceeding 10% efficiency by design optimization,” Int. J. Photoenergy, 2012 (1), 1 –7 (2012). https://doi.org/10.1155/2012/460919 Google Scholar

25. 

S. M. Iftiquar et al., “Single- and multiple-junction p-i-n type amorphous silicon solar cells with p-a-Si1xCx:H and nc-Si:H films,” Photodiodes—From Fundamentals to Applications, 289 –311 InTech, Rijeka, Croatia (2012). Google Scholar

26. 

( (2017) www.comsol.com/ June ). 2017). Google Scholar

27. 

B. J. Civiletti et al., “Optimization approach for optical absorption in three-dimensional structures including solar cells,” Opt. Eng., 57 (5), 057101 (2018). https://doi.org/10.1117/1.OE.57.5.057101 Google Scholar

28. 

T. H. Anderson et al., “Towards numerical simulation of nonhomogeneous thin-film silicon solar cells,” Proc. SPIE, 8981 (1), 898115 (2014). https://doi.org/10.1117/12.2041663 PSISDG 0277-786X Google Scholar

29. 

I. H. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am., 55 (10), 1205 –1208 (1965). https://doi.org/10.1364/JOSA.55.001205 JOSAAH 0030-3941 Google Scholar

30. 

E. D. Palik, Handbook of Optical Constants of Solids, Academic Press, Boston, Massachusetts (1985). Google Scholar

31. 

W. S. M. Werner, K. Glantschnig and C. Ambrosch-Draxl, “Optical constants and inelastic electron-scattering data for 17 elemental metals,” J. Phys. Chem. Ref. Data, 38 (4), 1013 –1092 (2009). https://doi.org/10.1063/1.3243762 JPCRBU 0047-2689 Google Scholar

32. 

National Renewable Energy Laboratory, “Reference Solar Spectral Irradiance: Air Mass 1.5,” (2018) https://www.nrel.gov/grid/solar-resource/spectra-am1.5.html April ). 2018). Google Scholar

33. 

J. Nelson, The Physics of Solar Cells, Imperial College Press, London (2003). Google Scholar

34. 

S. J. Fonash, Solar Cell Device Physics, 2nd ed.Academic Press, Burlington, Massachusetts (2010). Google Scholar

35. 

J. Wu and W. Walukiewicz, “Band gaps of InN and group III nitride alloys,” Superlatt. Microstruct., 34 (1), 63 –75 (2003). https://doi.org/10.1016/j.spmi.2004.03.069 SUMIEK 0749-6036 Google Scholar

36. 

G. F. Brown et al., “Finite element simulations of compositionally graded InGaN solar cells,” Solar Energy Mater. Solar Cells, 94 (3), 478 –483 (2010). https://doi.org/10.1016/j.solmat.2009.11.010 SEMCEQ 0927-0248 Google Scholar

37. 

D. R. Lide, CRC Handbook of Chemistry and Physics, 89th ed.CRC Press, Boca Raton, Florida (2008). Google Scholar

38. 

L. Vegard, “XV. Die Röntgenstrahlen im Dienste der Erforschung der Materie,” Z. Kristallogr., 67 (1–6), 239 –259 (1928). https://doi.org/10.1524/zkri.1928.67.1.239 ZEKRDZ 0044-2968 Google Scholar

39. 

J. W. Slotboom and H. C. de Graaff, “Measurements of bandgap narrowing in Si bipolar transistors,” Solid-State Electron., 19 (10), 857 –862 (1976). https://doi.org/10.1016/0038-1101(76)90043-5 Google Scholar

41. 

S. Das and P. N. Suganthan, “Differential evolution: a survey of the state-of-the-art,” IEEE Trans. Evol. Comput., 15 (1), 4 –31 (2011). https://doi.org/10.1109/TEVC.2010.2059031 ITEVF5 1089-778X Google Scholar

42. 

N. Anttu et al., “Absorption and transmission of light in III-V nanowire arrays for tandem solar cell applications,” Nanotechnology, 28 (20), 205203 (2017). https://doi.org/10.1088/1361-6528/aa6aee NNOTER 0957-4484 Google Scholar

43. 

Y. Jaluria, Computer Methods for Engineering, Taylor and Francis, Washington, DC (1996). Google Scholar

44. 

D. Chen, “Anti-reflection (AR) coatings made by sol–gel processes: a review,” Sol. Energy Mater. Sol. Cells, 68 (3), 313 –336 (2001). https://doi.org/10.1016/S0927-0248(00)00365-2 SEMCEQ 0927-0248 Google Scholar

45. 

J. Moghal et al., “High-performance, single-layer antireflective optical coatings comprising mesoporous silica nanoparticles,” ACS Appl. Mater. Interfaces, 4 (2), 854 –859 (2012). https://doi.org/10.1021/am201494m AAMICK 1944-8244 Google Scholar

46. 

T. H. Anderson, A. Lakhtakia and P. B. Monk, “Optimal indium-gallium-nitride Schottky-barrier thin-film solar cells,” Proc. SPIE, 10368 (1), 103680A (2017). https://doi.org/10.1117/12.2272739 PSISDG 0277-786X Google Scholar

Biography

Tom H. Anderson received his MSc degree from the University of York in 2012, for work on modeling of nonlocal transport in Tokamak plasmas. He received his PhD from the University of Edinburgh, United Kingdom, in 2016, for his thesis, “Optoelectronic Simulations of Nonhomogeneous Solar Cells.” He is presently affiliated with the University of Delaware. His current research interests include optical and electrical modeling of solar cells, numerics, plasma physics, and plasmonics.

Akhlesh Lakhtakia is an Evan Pugh University professor and the Charles Godfrey Binder professor of engineering science and mechanics at the Pennsylvania State University. His current research interests include surface multiplasmonics, solar cells, sculptured thin films, mimumes, bioreplication, and forensic science. He has been elected a fellow of OSA, SPIE, IoP, AAAS, APS, IEEE, RSC, and RSA. He received the 2010 SPIE Technical Achievement Award and the 2016 Walston Chubb Award for Innovation.

Peter B. Monk is a Unidel professor in the Department of Mathematical Sciences at the University of Delaware. He is the author of Finite Element Methods for Maxwell’s Equations and coauthor with F. Cakoni and D. Colton of The Linear Sampling Method in Inverse Electromagnetic Scattering.

© 2018 Society of Photo-Optical Instrumentation Engineers (SPIE) 1947-7988/2018/$25.00 © 2018 SPIE
Tom H. Anderson, Akhlesh Lakhtakia, and Peter B. Monk "Optimization of nonhomogeneous indium-gallium-nitride Schottky-barrier thin-film solar cells," Journal of Photonics for Energy 8(3), 034501 (13 September 2018). https://doi.org/10.1117/1.JPE.8.034501
Received: 12 May 2018; Accepted: 13 August 2018; Published: 13 September 2018
Lens.org Logo
CITATIONS
Cited by 12 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Solar cells

Gallium

Indium

Electrodes

Electrons

Gallium nitride

Indium nitride


CHORUS Article. This article was made freely available starting 13 September 2019

Back to Top