Open Access
4 May 2018 Optimization approach for optical absorption in three-dimensional structures including solar cells
Benjamin J. Civiletti, Thomas H. Anderson, Faiz Ahmad, Peter B. Monk, Akhlesh Lakhtakia
Author Affiliations +
Abstract
The rigorous coupled-wave approach (RCWA) and the differential evolution algorithm (DEA) were coupled in a practicable approach to maximize absorption in optical structures with three-dimensional morphology. As a model problem, optimal values of four geometric parameters and the bandgaps of three i-layers were found for an amorphous-silicon, multiterminal, thin-film tandem solar cell comprising three p  –  i  –  n junctions with a metallic hexagonally corrugated backreflector. When the optical short-circuit current density was chosen as the figure of merit to be maximized, only the bandgap of the topmost i-layer was significant and the remaining six parameters played minor roles. While this configuration would absorb light very well, it would have poor electrical performance. This issue arises because the optimization problem allows for the thicknesses and bandgaps of the semiconductor layers to change. We therefore devised another figure of merit that takes into account bandgap changes by estimating the open-circuit voltage. The resulting configuration was found to be optimal with respect to all seven variable parameters. The RCWA + DEA optimization approach is applicable to other types of photovoltaic solar cells as well as optical absorbers, with the choice of the figure of merit being vital to a successful outcome.

1.

Introduction

Three items are needed to numerically optimize the design of an optical absorber such as a thin-film solar cell. The first item is a fast solver that can predict the performance of the device in a variety of configurations. The second item is an optimization code that can mitigate the effect of local minima without excessive computational time. The third item is a figure of merit that adequately captures the desired performance characteristics of the device so that a good design will emerge. Some of these choices are explored in this paper.

The rigorous coupled-wave approach (RCWA)1,2 can be used to model the optical performance of thin-film optical absorbers, as has been shown for solar cells with metallic backreflectors that are periodically corrugated along one direction.35 Indeed, the RCWA provides accurate results with high computational speed for boundary-value problems involving structures that are invariant only along, say, the y-axis and therefore are quasi-two-dimensional.6,7 Furthermore, the RCWA can be coupled with the differential evolution algorithm (DEA)8,9 for optimization.5 However, the computational requirements of RCWA increase significantly when the backreflector is periodically corrugated in two directions, i.e., the boundary-value problem is fully three-dimensional (3-D) in nature.10,11 Additional design parameters enter the optimization process thereby to increase the computational burden further.

The dimensions of the unit cell of an optical absorber with a PCBR directly affect optical absorption.5 Optimization of thin-film solar cells with two-dimensionally corrugated backreflectors for maximum absorption has not been reported heretofore, to our knowledge. As a preliminary study showed that it is becoming a practicable proposition with commonly available computational resources, we decided to implement the RCWA + DEA approach to optimize a fully 3-D absorbing structure.12 To demonstrate this approach, we report here the maximization of optical absorption in an idealized thin-film tandem solar cell fabricated over a periodically corrugated backreflector (PCBR) with hillock-shaped corrugations arranged on a hexagonal lattice. The active region of the chosen solar cell comprises three electrically isolated pin junctions. The semiconductor layers were taken to have the bandgap-dependent optical properties of amorphous silicon.13 Silver 14 is a good choice for the PCBR because its plasmonic nature can be harnessed to launch surface-plasmon-polariton (SPP) waves inside the device and thereby enhance the optical electric field and optical absorption.1517 With the foregoing choices, our results indicate that maximization of the optical short-circuit current density, the standard figure of merit,1820 does not result in a desirable design. Instead, we found that the maximum power density is a better figure of merit.

The plan of this paper is as follows: the optical boundary-value problem that is solved to determine the spectrally integrated number of absorbed photons per unit volume per unit time Nph is presented in Sec. 2. The numerical techniques adopted for this work are presented in Sec. 3: the 3-D implementation of the RCWA is briefly described in Sec. 3.1, Sec. 3.2 discusses the diagonalization of a matrix that emerges in the RCWA implementation, and Sec. 3.3 briefly describes the DEA. Numerical results are provided in Sec. 4, Sec. 4.3 discusses the convergence of the numerical methods, whereas Sec. 4.4 briefly compares these numerical results to a thin-film tandem solar cell with a bisinusoidal PCBR. Closing remarks are presented in Sec. 5.

The free-space wavenumber, angular frequency, and intrinsic impedance of free space are denoted by k0=2π/λ0, ω=k0c0, and η0=μ0/ϵ0, respectively, where λ0 is the free-space wave length, μ0 is the permeability of free space, ϵ0 is the permittivity of free space, and c0=1/ϵ0μ0 is the speed of light in free space, respectively. Vectors are underlined, column vectors and matrices associated with the RCWA are in boldface with breve notation, and the Cartesian unit vectors are identified as x^_, y^_, and z^_. The imaginary unit is denoted by i=1.

2.

Model Boundary-Value Problem

We considered the boundary-value problem shown schematically in Fig. 1, which also defines the thicknesses Ld, Lg, and Lm. The device occupies the region

Eq. (1)

X:{(x,y,z)|<x<,<y<,0<z<Ld+Lg+Lm},
with the half spaces z<0 and z>Ld+Lg+Lm occupied by air. The reference unit cell is identified as R:{(x,y,z)|Lx/2<x<Lx/2,Ly/2<y<Ly/2,0<z<Ld+Lg+Lm}, the backreflector (which also functions as an electrode in a solar cell) being doubly periodic with period Lx along the x-axis and period Ly along the y-axis.

Fig. 1

(a) Schematic of the model boundary-value problem in the plane y=0. The n-type semiconductor layers are blue, the p-type semiconductor layers are red, and the i-type are gray. White regions are occupied by a material with real relative permittivity ϵw(λ0). (b) Schematic of the grating region in the plane z=zg[Ld,Ld+Lg].

OE_57_5_057101_f001.png

The region 0<z<Ld comprises an antireflection window and three pin junctions that are electrically isolated from each other by two windows, as shown in Fig. 1(a). The relative permittivity ϵd(x,y,z,λ0) of this multilayered material depends on λ0. The layers are identified in the figure. All windows are made of a material of relative permittivity ϵw(λ0). The λ0-dependent relative permittivity of each semiconductor layer depends on the bandgap chosen for that layer.

The region Ld+Lg<z<Ld+Lg+Lm is occupied by a metal with relative permittivity ϵm(λ0). The region Ld<z<Ld+Lg, henceforth termed the grating region, contains a periodically undulating surface with period Lx along the x-axis and period Ly along the y-axis. The unit cell in the xy plane was chosen to form a two-dimensional (2-D) rectangular lattice that is equivalent to a hexagonal lattice. If the side of the regular hexagons in this lattice is denoted by Lh, then Lx=Lh and Ly=3Lh for the rectangular lattice.

The grating region is defined by hillocks arranged as in Fig. 1(b). Each hillock is a frustum of a sphere of radius Rsph. The base of the hillock is a circle of radius Rg, and the height of the hillock equals Lg, as shown in Fig. 1(a); accordingly

Eq. (2)

Rsph=12(Rg2Lg+Lg).
The intersection of the plane z=zg[Ld,Ld+Lg] and a hillock is a circle of radius

Eq. (3)

Rˇ=(zgLd)[2Rsph(zgLd)].
The relative permittivity ϵg(x,y,z,λ0)=ϵg(x±Lx,y±Ly,z,λ0) in the grating region is

Eq. (4)

ϵg(x,y,z,λ0)=ϵm(λ0)[ϵm(λ0)ϵw(λ0)]U(x,y,z),
with

Eq. (5)

U(x,y,z)={1,DRˇ0,D<Rˇ,
where D is the minimum distance between the point (x,y,z) and the centers (0,0,zg), (Lx,0,zg), (Lx/2,Ly/2,zg), (Lx,Ly,zg), and (0,Ly,zg) of the circle and four quarter circles as shown in Fig. 1(b).

3.

Numerical Techniques Used

3.1.

Rigorous Coupled-Wave Approach

The RCWA was used to the calculate the electric field phasor everywhere inside the chosen device as a result of illumination by a monochromatic plane wave normally incident on the plane z=0 from the half space z<0. The electric field phasor of the incident plane wave was taken as

Eq. (6)

E_inc(x,y,z,λ0)=Eox^_+y^_2exp(ik0z),
where Eo=415πVm1.

As a result of the PCBR being doubly periodic, the x- and y-dependences of the electric and magnetic field phasors have to be represented everywhere by Fourier series as

Eq. (7)

E_(x,y,z,λ0)=m=m=n=n=e_(m,n)(z,λ0)exp{i[kx(m)x+ky(n)y]}
and

Eq. (8)

H_(x,y,z,λ0)=m=m=n=n=h_(m,n)(z,λ0)exp{i[kx(m)x+ky(n)y]},
where kx(m)=m(2π/Lx), ky(n)=n(2π/Ly), and e_(m,n)=ex(m,n)x^_+ey(m,n)y^_+ez(m,n)z^_ as well as h_(m,n)=hx(m,n)x^_+hy(m,n)y^_+hz(m,n)z^_ are Fourier coefficients. Likewise, the relative permittivity ϵrel(x,y,z,λ0) everywhere is represented by the Fourier series

Eq. (9)

ϵrel(x,y,z,λ0)=m=m=n=n=ϵrel(m,n)(z,λ0)exp{i[kx(m)x+ky(n)y]},
where ϵrel(m,n)(z,λ0) are Fourier coefficients. Substitution of Eqs. (7)–(9) in the frequency-domain Maxwell curl postulates yields the matrix ordinary differential equation [Ref. 21, Chap. 3]

Eq. (10)

ddzf˘(z,λ0)=iP˘(z,λ0)f˘(z,λ0),
where the column vector f˘(z,λ0) contains the x- and y-directed components of the Fourier coefficients of the electric and magnetic field phasors.

Detailed descriptions of the algorithm to solve Eq. (10) are available elsewhere.12,21 In brief, computational tractability requires the foregoing expansions to be truncated to include only m{Mt,,Mt} and n{Nt,,Nt}, with Mt0 and Nt0. The region R is partitioned into a sufficiently large number of slices along the z-axis. Each slice is taken to be homogeneous along the z-axis but may be periodically nonhomogeneous along the x- and y-axes. Thus, the matrix P˘(z,λ0) is assumed to be piecewise uniform in z. Boundary conditions are enforced on the planes z=0 and z=Ld+Lg+Lm to match the fields to the incident, reflected, and transmitted fields, as appropriate. A stable marching algorithm is then used to determine the Fourier coefficients of the electric and magnetic field phasors in each slice.21 Finally, the z-directed components of the electric and magnetic field phasors in the device can be obtained through algebraic equations arising during the derivation of Eq. (10). Thus, the electric field phasor E_(x,y,z,λ0) can be determined throughout the solar cell.

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

Eq. (11)

Nph(x,y,z)=1c0λ0minλ0maxIm{ϵrel(x,y,z,λ0)}|E_(x,y,z,λ0)Eo|2S(λ0)dλ0,
where is the reduced Planck constant, S(λ0) is the AM1.5G solar spectrum,22 λ0min=400  nm, and λ0max=1100  nm. With the assumption that the absorption of every photon in a semiconductor layer releases an electron-hole pair, the charge-carrier-generation rate G(x,y,z) equals Nph(x,y,z), which can be determined at any location in the nine semiconductor layers. The integral on the right side of Eq. (11) was approximated using the trapezoidal rule23 with the integrand uniformly sampled every 10 nm.

3.2.

Diagonalization of P˘(z,λ0)

The numerically stable RCWA algorithm requires that P˘(z,λ0) be diagonalizable24 in each slice for every λ0[λ0min,λ0max] [Ref. 21, Sec. 3.8]. As the relative permittivity is not uniform in any slice in the grating region, the built-in function eig in MATLAB® (version R2016b) was used to compute the eigenvalues and eignenvectors of P˘. In all other slices, the eigenvalues and eigenvectors of P˘ were determined analytically to increase the computational speed.

A superindex

Eq. (12)

τ=m(2Nt+1)+n,m[Mt,Mt],n[Nt,Nt]
is defined for convenience, where τt=2MtNt+Mt+Nt.21 In any slice in which the relative permittivity ϵrel(x,y,z,λ0) is independent of x and y, the 2(2τt+1) distinct eigenvalues of P˘ are given by

Eq. (13)

gτ±=±k02ϵrel[k˘x(τ)]2[k˘y(τ)]2,τ{τt,,τt},
with each eigenvalue having a geometric multiplicity of 2, k˘x(τ)=kx(m), and k˘y(τ)=ky(n). Half of the 4(2τt+1) eigenvectors are

Eq. (14)

vτ1±=[0,,0,(gτ±)1{ωμ01ωϵ0ϵrel[k˘x(τ)]2},0,,0,k˘x(τ)k˘y(τ)ωϵ0ϵrelgτ±,0,,0,1,0,,0]T,τ{τt,,τt}.

In the column vector on the right side of Eq. (14), the nonzero entries occur in the following locations (counting from the top): τ, τ+2τt+1, and τ+6τt+3. The remaining 2(2τt+1) eigenvectors are

Eq. (15)

vτ2±=[0,,0,k˘x(τ)k˘y(τ)ωϵ0ϵrelgτ±,0,,0,(gτ±)1{ωμ01ωϵ0ϵrel[k˘y(τ)]2},0,,0,1,0,,0]Tτ{τt,,τt}.
The nonzero entries occur in the column vector on the right side of Eq. (15) in the following locations (counting from the top): τ, τ+2τt+1, and τ+4τt+2.

3.3.

Differential Evolution Algorithm

We employed the DEA to maximize a chosen figure of merit over a variety of optical and electrical parameters numbering N¯. The DEA has been used previously5 for seeking optimal designs of PCBRs that are invariant along the y-axis, as the algorithm is well suited to search a large space of candidate solutions. The number of candidate solutions depends on the number of parameters and discretization of parameter ranges for the optimization.

To maximize the figure of merit C:SRN¯R, we wish to find an optimal point Vopt={v1,v2,,vN¯}S, where S is a search space of all possible parameter combinations.

We note that the DEA requires that all parameter ranges be discretized, so the search is conducted over a finite number of possible outcomes. The DEA requires specification of the crossover probability CR(0,1), a differential weight α(0,2), and the number of random points NP. Details of the algorithm have been provided elsewhere.5,8,9

The DEA is very useful for solving complicated optimization problems, but it does not guarantee convergence to a global extremum.25 However, stochastic sampling of the search space helps to avoid local maxima.

4.

Numerical Results and Discussion

In this paper, we have chosen to maximize optical absorption. The figures of merit defined later in this section take into account all optical effects such as the excitation of SPP waves and waveguide modes.26 This allows a tradeoff between the various optical phenomena without prejudicing one mechanism over another. However, the parameter space is chosen so that the excitation of SPP waves and waveguide modes can be supported.

For all numerical results in this paper, the window layers were chosen to be made of aluminum-doped zinc oxide (AZO). For the two windows between pin junctions, the thicknesses were fixed so that d2d=d1d=20  nm. The relative permittivity ϵw(λ0) of AZO was taken from a standard source.27 The minimum thickness of the PCBR was fixed at Lm=150  nm. The metal was chosen to be silver, whose relative permittivity ϵm also depends on λ0.14

The bandgaps Ei, {1,2,3}, of the i-layers in the triple-junction solar cell were kept variable in the range [1.3, 1.95] eV, but their thicknesses d1i=d2i=d3i=200  nm were kept fixed. The thicknesses of all three n-layers and all three p-layers were also kept fixed dn=dp=20  nm. The bandgaps of all three n-layers were fixed as En=1.8  eV, {1,2,3}. The bandgaps of the p-layers were fixed as follows: E1p=1.8  eV and E2p=E3p=1.95  eV.

By introducing C or Ge into the lattice, a material is formed, but the i-layers are still of the a-Si:H_GeC family. This process changes the bandgap, where the λ0-dependent relative permittivity of the material is obtained by an analytical model.13,17 The electrical properties of the material also change, and can be found by applying Vegard’s law to known values.28 As the electrical properties have no effect on our optical model, they have no role in this study.

Furthermore, the lattice parameter Lh[200,800]  nm, the antireflection-window’s thickness dw[10,130]  nm, the base radius Rg[10,400]  nm, and the corrugation height Lg[0,300]  nm were allowed to vary. Thus, the dimension of the search space S was N¯=7, and we sought an optimal design over a candidate space of 9×1011 possible configurations. We used parameter values CR=0.7, α=0.8, and NP=70 for optimization.

4.1.

Optimization for Optical Short-Circuit Current Density

The figure of merit C for the DEA optimization was initially chosen to be standard figure of merit for optical modeling of solar cells:29 the optical short-circuit current density

Eq. (16)

JSCOpt=qeLxLyRscNph(x,y,z)dxdydz,
where Rsc is the portion of the reference unit cell R occupied by the nine semiconductor layers in the solar cell and qe=1.6×1019  C is the elementary charge. This figure of merit will maximize the number of photons absorbed in the solar cell, but disregards all electrical properties. In this section, we proceed to show that this results in a poor design.

Plots of JSCOpt against the bandgaps Ei, {1,2,3}, for points sampled by the optimization exercise are shown in Fig. 2. In this figure as well as in Fig. 3, the data points from DEA are projected onto the plane formed by the variable being investigated and the figure of merit. The bandgap of the topmost i-layer is the most influential parameter that controls JSCOpt. While attempting to maximize JSCOpt, the DEA minimized the bandgap in this layer, and we see that many parameter sets with E3iopt=1.3  eV at the boundary of the constraint set were evaluated. As E3i increases from 1.3 to 1.95 eV, JSCOpt decreases throughout most of this interval as shown in Fig. 2(a).

Fig. 2

JSCOpt in relation to (a) E3i, (b) E2i, and (c) E1i. Each marker (·) represents a choice of parameters by DEA as the algorithm progresses. Larger values of JSCOpt are desirable.

OE_57_5_057101_f002.png

Fig. 3

Psup in relation to (a) Lh, (b) dw, (c) ζ=Rg/Lh, and (d) Lg. Each marker (·) represents a choice of parameters by DEA as the algorithm progresses. The large marker indicates the maximum value of Psup achieved.

OE_57_5_057101_f003.png

Maximization of JSCOpt showed that minimizing E3i was the most important factor compared with the remaining N¯1 variable parameters (i.e., E1i, E2i, Lh, dw, Rg, and Lg). This phenomenon is clearly exemplified in Fig. 2(a), where, for any fixed value of E3i, the remaining N¯1 variable parameters contribute only to a ±2  mAcm2 variation in JSCOpt. Unfortunately, such a configuration is likely to be electrically inefficient: with a narrow bandgap, more charge carriers are excited, but the operating voltage of the solar cell will be reduced.30 Contrast to Figs. 2(b) and 2(c) wherein the variations of JSCOpt with the bandgaps E2i and E1i, respectively, are shown. We see that while JSCOpt is maximum when E1i=E2i=1.35eV, there are values of JSCOpt ranging from 18 to 27  mAcm2.

Parenthetically, when the thicknesses of the i-layers were included as variables in an optimization exercise, the DEA simply focused on the maximization of those thicknesses. The resulting configuration would also have poor electrical performance. Although this type of solar cell would absorb more light, an excited charge carrier would have to travel further to reach an electrode, thereby increasing recombination and decreasing efficiency. We note that Fig. 2 demonstrates the choice of figure of merit is very important. We only included these results to contrast them with numerical results in Fig. 5, since all the chosen parameters in the optimization exercise should affect the solar-cell performance.

4.2.

Optimization for Maximum Power Density

To improve the optimal design of the chosen solar cell without including a full electrical model, we devised a figure of merit for the DEA that penalizes the effect of minimizing the bandgap of any of the i-layers. We defined the power density

Eq. (17)

Psup=1LxLy=13EiRNph(x,y,z)dxdydz,
as the figure of merit, with RR being the region occupied by the ’th pin junction. Let us note that Psup is a theoretical upper bound on the maximum extractable power density of the solar cell. Furthermore, the summation over the index indicates that the tandem solar cell is configured in the multiterminal format. Note that this power density is to be computed solely from the absorption of photons and the material bandgap. This estimates the maximum electrical power density but does not involve any electrical modeling (e.g., recombination and mobility of electrons and holes). A useful extension of our approach would be to include an electrical model, but that extension lies beyond the scope of this paper.

Figures 3 and 4 show the results of the DEA optimization for the variable geometric and bandgap parameters of the solar cell—namely, the lattice parameter Lh, the antireflection-window thickness dw, the duty cycle ζ=Rg/Lh, the corrugation height Lg, and the bandgaps E1i, E2i, and E3i. All seven of these parameters influence the figure of merit Psup defined by Eq. (17). Indeed, a steady increase in Psup is seen in Fig. 3(a) on the interval 200Lh600  nm; a very sharp increase in Psup is evident in Fig. 3(b) as dw increases from 10 to 80 nm; Psup increases steeply in Fig. 3(c) as ζ approaches 0.35 with a drop off thereafter; and Psup peaks in the neighborhood of Lg=200  nm in Fig. 3(d). Also shown is the hillock base radius Rg in Fig. 4(d). We note the contrast between the behavior seen in Figs. 2(a) and 4(c). In the latter optimization exercise, the effect of the other parameters amounts to a variation of ±5  mWcm2.

Fig. 4

Psup in relation to (a) E1i, (b) E2i, (c) E3i, and (d) Rg. Each marker (·) represents a choice of parameters by DEA as the algorithm progresses. The large marker indicates the maximum value of Psup achieved.

OE_57_5_057101_f004.png

The optimal values found are as follows: E1i=1.35  eV, E2i=1.95  eV, E3i=1.65  eV, Lh=642  nm, dw=89  nm, ζ=0.36, and Lg=231  nm. These values yielded a maximum Psup=43.66  mWcm2. These optimal parameters, found by maximizing Psup, contrast sharply with our findings in Sec. 4.1, wherein maximization of JSCOpt was dominated by the minimization of E3i with the remaining N¯1 parameters having very little effect.

The optimization exercise yielded two distinct categories of unit cells with relatively high Psup values. The first comprises configurations for which Lh lies in the interval [200, 300] nm, and the second comprises configurations for which Lh[550,650]  nm. For the first category, the optimal value of Lg is 40  nm with a base radius Rg of 50 nm. The second type of configuration has much deeper corrugations, with Psup maximized when Lg takes values near 200 nm and Rg around 230 nm. This phenomenon is evidenced by two distinct peaks in Figs. 3(a)3(d) and 4(d).

4.3.

Convergence of RCWA and DEA

To ensure convergence of the optical short-circuit density, a representative configuration for the unit cell was used to determine an appropriate choice of Nt and Mt. We let Nt vary in the set {2,3,4,5,6}, and defined Mt=3Nt, where · is the ceiling function. After determining that Psup changed by 1% for two successive values of Nt and Mt, the number of Fourier modes was fixed for all numerical results reported in this paper, and taken to be Nt=2, Mt=4. Since Eq. (16) has the charge-carrier generation rate integrated over the nine semiconductor layers, the susceptibility of JSCOpt to the effect of Gibbs’ phenomena6 on the electric field in the region z[0,Ld] is negligible. Hence, the electric field converges everywhere in the semiconductor layers, even for relatively small values of Nt and Mt. We performed 50 DEA iterations in the optimization exercise. Figure 5 shows the convergence of Psup with the number of DEA iterations. We see that the best value of Psup does not change after the first 32 DEA iterations. As with any stochastic optimization method, it is always possible that further iteration would result in improvement of the computed maximum power density.

Fig. 5

The best value of Psup versus number of DEA iterations when optimizing for maximum power density as in Sec. 4.2.

OE_57_5_057101_f005.png

4.4.

Comparison with Bisinusoidal PCBR

A third DEA-based optimization exercise was performed with the PCBR taken to be bisinusoidally corrugated.11 The relative permittivity ϵg(x,y,z,λ0)=ϵg(x±Lx,y±Ly,z,λ0) in the grating region is

Eq. (18)

ϵg(x,y,z,λ0)=ϵm(λ0)[ϵm(λ0)ϵw(λ0)]U[zg1(x)]U[zg2(y)],
with the corrugation-shape functions

Eq. (19)

g1(x)=Ld+Lg[1sin(2πxLx)],

Eq. (20)

g2(y)=Ld+Lg[1sin(2πyLy)].

The DEA parameters were kept the same as in Sec. 4, and we set Lx=Ly for the computation. The optimization exercise yielded an optimal configuration with dw=87  nm, Lg=110  nm, Lx=607  nm, Rg=212  nm, E3i=1.7  eV, E2i=1.55  eV, and E1i=1.3  eV. The maximal power density achieved was Psup=42.84  mWcm2, which is slightly lower than 43.66  mWcm2 obtained for the hexagonally corrugated PCBR in Sec. 4.2. However, the average Psup over all configurations visited by the DEA for the bisinusoidally corrugated PCBR is 38.17  mWcm2 but 35.06  mWcm2 for the hexagonally corrugated PCBR. There were many more configurations with relatively poor Psup values for the hexagonal case.

5.

Conclusions

As a model problem to demonstrate the practicability of the RCWA + DEA approach to design efficient optically absorbing 3-D structures, optimal values of four geometric parameters and the bandgaps of three i-layers were found for an idealized, multiterminal, thin-film tandem solar cell comprising three pin junctions with a silver PCBR with hillock-shaped corrugations arranged on a hexagonal lattice. The figure of merit for the DEA was either (i) the optical short-circuit current density in Sec. 4.1 or (ii) the power density in Sec. 4.2. Thus, two different optimization exercises using the RCWA + DEA approach were performed.

As the optical short-circuit current density takes into account only the optical constitutive properties of the solar cell, maximization of that quantity resulted in a poor design. In particular, we determined that only the bandgap of the topmost i-layer (i.e., E3i) was significant to the optimization of the optical short-circuit current density, and the remaining six parameters played minor roles. Although photon absorption in the topmost pin junction was maximized thereby, this configuration would have poor electrical performance.

This is because, when the thicknesses of the i-layers were included in the optimization, the DEA simply focused on the maximization of those thicknesses. Although increasing those thicknesses can enhance light absorption, the electrical performance may be sacrificed, thereby reducing efficiency. The design of thin-film solar cells must balance optical and electrical performances.31,32

To avoid configurations with potentially poor electrical performance, we next used the power density—thereby weighting the optical short-circuit current density—as a figure of merit. The resulting configuration was optimal with respect to all seven design parameters. Another optimization exercise was then performed on a similar tandem solar cell but with a bisinusoidally corrugated PCBR. In this case, the configurationally averaged power density tested by the DEA was about 10% higher than with the hexagonally corrugated PCBR. In the future, we plan to supplement the optical model by an electrical drift-diffusion model32 and then optimize the overall electrical performance of the solar cell.

In closing, let us emphasize that the triple-junction tandem solar cell was chosen as a model problem to show here the capabilities of the RCWA + DEA approach developed for 3-D optically absorbing structures. Our approach can be extended not only to other types of photovoltaic solar cells33,34 but also to optical absorbers35,36 with 3-D morphology.

Acknowledgments

This paper was substantially based on a paper titled “Optimization of charge-carrier generation in amorphous-silicon thin-film tandem solar cell backed by two-dimensional metallic surface-relief grating,” presented at the SPIE conference Next Generation Technologies for Solar Energy Conversion VIII, held August 5 to 11, 2017 in San Diego, California, United States. The research of B.J. Civiletti, T.H. Anderson, and P.B. Monk is partially supported by the US National Science Foundation (NSF) under grant number DMS-1619904. The research of F. Ahmed and A. Lakhtakia is partially supported by the US NSF under grant number DMS-1619901. A. Lakhtakia thanks the Charles Godfrey Binder Endowment at the Pennsylvania State University for ongoing support of his research.

References

1. 

M. G. Moharam et al., “Formulation for stable and efficient implementation of the rigorous coupled-wave analysis of binary gratings,” J. Opt. Soc. Am. A, 12 1068 –1076 (1995). https://doi.org/10.1364/JOSAA.12.001068 JOAOD6 0740-3232 Google Scholar

2. 

N. Chateau and J.-P. Hugonin, “Algorithm for the rigorous coupled-wave analysis of grating diffraction,” J. Opt. Soc. Am. A, 11 1321 –1331 (1994). https://doi.org/10.1364/JOSAA.11.001321 JOAOD6 0740-3232 Google Scholar

3. 

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

4. 

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

5. 

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

6. 

M. Weismann, D. F. G. Gallagher and N. C. Panoiu, “Accurate near-field calculation in the rigorous coupled-wave analysis method,” J. Opt., 17 125612 (2015). https://doi.org/10.1088/2040-8978/17/12/125612 Google Scholar

7. 

M. V. Shuba et al., “Adequacy of the rigorous coupled-wave approach for thin-film silicon solar cells with periodically corrugated metallic backreflectors: spectral analysis,” J. Opt. Soc. Am. A, 32 1222 –1230 (2015). https://doi.org/10.1364/JOSAA.32.001222 JOAOD6 0740-3232 Google Scholar

8. 

R. Storn and K. Price, “Differential evolution—a simple and efficient heuristic for global optimization over continuous spaces,” J. Global Optim., 11 341 –359 (1997). https://doi.org/10.1023/A:1008202821328 JGOPEO 0925-5001 Google Scholar

9. 

K. Price, R. M. Storn and J. A. Lampinen, Differential Evolution: A Practical Approach to Global Optimization, Springer, New York (2005). Google Scholar

10. 

M. Onishi, K. Crabtree and R. A. Chipman, “Formulation of rigorous coupled-wave theory for gratings in bianisotropic media,” J. Opt. Soc. Am. A, 28 1747 –1758 (2011). https://doi.org/10.1364/JOSAA.28.001747 JOAOD6 0740-3232 Google Scholar

11. 

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 016017 (2018). https://doi.org/10.1117/1.JNP.12.016017 1934-2608 Google Scholar

12. 

B. J. Civiletti et al., “Optimization of charge-carrier generation in amorphous-silicon thin-film tandem solar cell backed by two-dimensional metallic surface-relief grating,” Proc. SPIE, 10368 1036809 (2017). https://doi.org/10.1117/12.2274077 PSISDG 0277-786X Google Scholar

13. 

A. S. Ferlauto et al., “Analytical model for the optical functions of amorphous semiconductors from the near-infrared to ultraviolet: applications in thin film photovoltaics,” J. Appl. Phys., 92 2424 –2436 (2002). https://doi.org/10.1063/1.1497462 JAPIAU 0021-8979 Google Scholar

14. 

Handbook of Optical Constants of Solids, Academic Press, Boston, Massachusetts (1985). 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. 

M. A. Green and S. Pillai, “Harnessing plasmonics for solar cells,” Nat. Photonics, 6 130 –132 (2012). https://doi.org/10.1038/nphoton.2012.30 NPAHBY 1749-4885 Google Scholar

17. 

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., 5253 087106129801 (20132014). https://doi.org/10.1117/1.OE.52.8.087106 Google Scholar

18. 

S. Zhou et al., “Optimizing two-level heirarchical particles for thin-film solar cells,” Opt. Express, 21 A285 –A294 (2013). https://doi.org/10.1364/OE.21.00A285 OPEXFF 1094-4087 Google Scholar

19. 

A. Naqavi et al., “Enhanced light trapping in realistic thin film solar cells using one dimensional gratings,” Proc. SPIE, 8065 80650A (2012). https://doi.org/10.1117/12.883186 PSISDG 0277-786X Google Scholar

20. 

L. C. Andreani, A. Bozzola and M. Liscidini, “Light trapping in thin-film solar cells: towards the Lambertian limit,” Proc. SPIE, 8438 84380C (2012). https://doi.org/10.1117/12.922090 PSISDG 0277-786X Google Scholar

21. 

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

22. 

National Renewable Energy Laboratory, “Reference Solar Spectral Irradiance: Air Mass 1.5,” (2018) http://rredc.nrel.gov/solar/spectra/am1.5/ April ). 2018). Google Scholar

23. 

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

24. 

H. Hochstadt, Differential Equations: A Modern Approach, Dover Press, New York (1975). Google Scholar

25. 

R. Knobloch, J. Mlynek and R. Srb, “Improving convergence properties of a differential evolution algorithm,” AIP Conf. Proc., 1789 030005 (2016). https://doi.org/10.1063/1.4968451 Google Scholar

26. 

L. Liu et al., “Planar light concentration in micro-Si solar cells enabled by a metallic grating–photonic crystal architecture,” ACS Photonics, 3 (4), 604 –610 (2016). http://doi.org/10.1021/acsphotonics.5b00706 KPSJAS 0374-4884 Google Scholar

27. 

X.-Y. Gao, L. Liang and Q.-G. Lin, “Analysis of the optical constants of aluminum-doped zinc-oxide films by using the single-oscillator model,” J. Korean Phys. Soc., 57 710 –714 (2010). https://doi.org/10.3938/jkps.57.710 KPSJAS 0374-4884 Google Scholar

28. 

L. Vegard, “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

29. 

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

30. 

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

31. 

M. Faryad et al., “Optical and electrical modeling of an amorphous-silicon tandem solar cell with nonhomogeneous intrinsic layers and a periodically corrugated backreflector,” Proc. SPIE, 8823 882306 (2013). https://doi.org/10.1117/12.2024712 PSISDG 0277-786X Google Scholar

32. 

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 025502 (2016). https://doi.org/10.1117/1.JPE.6.025502 Google Scholar

33. 

J. Zhu et al., “Nanodome solar cells with efficient light management and self-cleaning,” Nano Lett., 10 2342 –2348 (2010). https://doi.org/10.1021/nl9034237 NALEFD 1530-6984 Google Scholar

34. 

R. Kapadia et al., “Nanopillar photovoltaics: materials, processes, and devices,” Nano Energy, 1 132 –144 (2012). https://doi.org/10.1016/j.nanoen.2011.11.002 Google Scholar

35. 

N. Liu et al., “Infrared perfect absorber and its application as plasmonic sensor,” Nano Lett., 10 1979 –1984 (2010). https://doi.org/10.1021/nl9041033 NALEFD 1530-6984 Google Scholar

36. 

S. Butun and K. Aydin, “Structurally tunable resonant absorption bands in ultrathin broadband plasmonic absorbers,” Opt. Express, 22 19457 –19468 (2014). https://doi.org/10.1364/OE.22.019457 OPEXFF 1094-4087 Google Scholar

Biography

Benjamin J. Civiletti received his MA degree in mathematics from Villanova University in 2014. Currently, he is a PhD student in the Department of Mathematical Sciences at the University of Delaware. His research interests include numerical methods for modeling solar cells.

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, UK, in 2016 for his thesis titled Optoelectronic Simulations of Nonhomogeneous Solar Cells. Currently, he is a postdoctoral researcher at the University of Delaware. His current research interests include optical and electrical modeling of solar cells, numerics, plasma physics, and plasmonics.

Faiz Ahmad received his MSc and MPhil degrees in electronics from Quaid-i-Azam University, Islamabad (Pakistan) in 2010 and 2012, respectively. Currently, he is a PhD student in the Department of Engineering Science and Mechanics at the Pennsylvania State University. His research interests include electromagnetic surface waves and light trapping in thin-film solar cells.

Peter B. Monk is a Unidel professor 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.

Akhlesh Lakhtakia is 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 as a fellow of Optical Society of America, SPIE, Institute of Physics, American Association for the Advancement of Science, American Physical Society, Institute of Electrical and Electronics Engineers, Royal Society of Chemistry, and Royal Society of Arts. He received the 2010 SPIE Technical Achievement Award and the 2016 Walston Chubb Award for Innovation.

© 2018 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2018/$25.00 © 2018 SPIE
Benjamin J. Civiletti, Thomas H. Anderson, Faiz Ahmad, Peter B. Monk, and Akhlesh Lakhtakia "Optimization approach for optical absorption in three-dimensional structures including solar cells," Optical Engineering 57(5), 057101 (4 May 2018). https://doi.org/10.1117/1.OE.57.5.057101
Received: 30 November 2017; Accepted: 30 March 2018; Published: 4 May 2018
Lens.org Logo
CITATIONS
Cited by 20 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Solar cells

Absorption

Optimization (mathematics)

3D modeling

Tandem solar cells

Thin film solar cells

Semiconductors


CHORUS Article. This article was made freely available starting 04 May 2019

Back to Top