Open Access
3 February 2017 Enhanced efficiency of Schottky-barrier solar cell with periodically nonhomogeneous indium gallium nitride layer
Author Affiliations +
Abstract
A two-dimensional finite-element model was developed to simulate the optoelectronic performance of a Schottky-barrier solar cell. The heart of this solar cell is a junction between a metal and a layer of n -doped indium gallium nitride ( In ξ Ga 1 ξ N ) alloy sandwiched between a reflection-reducing front window and a periodically corrugated metallic back reflector. The bandgap of the In ξ Ga 1 ξ N layer was varied periodically in the thickness direction by varying the parameter ξ ( 0,1 ) . First, the frequency-domain Maxwell postulates were solved to determine the spatial profile of photon absorption and, thus, the generation of electron–hole pairs. The AM1.5G solar spectrum was taken to represent the incident solar flux. Next, the drift-diffusion equations were solved for the steady-state electron and hole densities. Numerical results indicate that a corrugated back reflector of a period of 600 nm is optimal for photon absorption when the In ξ Ga 1 ξ N layer is homogeneous. The efficiency of a solar cell with a periodically nonhomogeneous In ξ Ga 1 ξ N

1.

Introduction

A variety of light-trapping strategies capitalizing on structures engineered on the order of the wavelength of solar light have been examined both experimentally and theoretically to enhance the efficiencies of solar cells.1 These strategies include texturing the front surface of the solar cell,2,3 incorporating a periodically corrugated metallic back reflector,46 coating the front surface with an “antireflection” (AR) layer,7,8 embedding nanoparticles inside the light-absorbing layer(s),911 and using light concentrators.1215

An attractive approach for boosting the solar-cell efficiency requires the use of photon-absorbing component materials whose electromagnetic properties are periodically nonhomogeneous along the thickness direction. Recently, it was demonstrated that the incorporation of a periodically nonhomogeneous intrinsic layer (i.e., i layer), along with a periodically corrugated back reflector, in amorphous silicon p-i-n junction solar cells can improve overall efficiency by up to 17%.16 This improvement is likely due to

  • i. the periodic corrugation of the metallic back reflector46,17 facilitating the excitation of surface-plasmon-polariton waves1820 and waveguide modes21 to intensify the electric field inside the semiconductor region, leading to an increase in the electron–hole pair (EHP) generation rate;

  • ii. the periodic nonhomogeneity of the i-layer that may facilitate the excitation of multiple surface-plasmon-polariton waves22 and waveguide modes,17 thereby further boosting the EHP generation rate; and

  • iii. the accompanying spatial gradient in the bandgap that may also aid charge separation and reduce the EHP recombination rate.23,24

The study described herein concerns an especially simple type of solar cell that has received scant attention from researchers to date: a Schottky-barrier solar cell.25 The Schottky barrier is provided by a metal–semiconductor junction.26,27 Recent theoretical studies,28,29 along with an earlier experimental study,30 have suggested that Schottky-barrier solar cells may be a particularly promising proposition if the semiconductor layer (i.e., the absorbing layer) was made from alloys of indium gallium nitride (InξGa1ξN), since the bandgap for these alloys can closely match the range of energies of photons across the entire solar spectrum (i.e., 0.70 to 3.42 eV) by varying the relative proportions of indium and gallium through the parameter ξ(0,1).31 Specifically, indium nitride (i.e., ξ=1) has a bandgap of 0.7 eV32,33 and absorbs efficiently across the infrared regime in the solar spectrum, while gallium nitride (i.e., ξ=0) has a bandgap of 3.42 eV and absorbs efficiently across the near-ultraviolet portion of the solar spectrum.

With current technologies, there are significant challenges to overcome in the production of InξGa1ξN alloys for all values of ξ, particularly for smaller-bandgap alloys. When the proportion of indium is large (i.e., ξ0.3), poor crystal growth plagues the realization of solar-cell applications. Such poor growth results in decreased carrier transport, background n doping due to Fermi pinning above the conduction band edge,34 and a bandgap that is greater than expected.35,36 By using a Schottky-barrier junction with n-doped InξGa1ξN, as opposed to a p-i-n junction, the difficulty of p doping the material is avoided.

In the following sections, a two-dimensional (2-D) numerical simulation is described for a Schottky-barrier solar cell made with InξGa1ξN. It is essential that the coupled processes of optical absorption and electrical-current generation are simultaneously accommodated in the simulation. This is because a focus on the computation of only the optical (i.e., at full quantum efficiency) short-circuit current density, but not of the open-circuit voltage, overplays the importance of the EHP generation rate by not taking the EHP recombination rate into account.16 In our simulation, particular attention is paid to the role of periodic nonhomogeneity of the absorbing material. Thus, the effect of periodically varying the fractional composition parameter ξ of InξGa1ξN in the direction perpendicular to the mean plane of the back reflector is explored, the back reflector being periodically corrugated in one direction. The model is briefly described in Secs. 2.1 and 2.2, with further details being available elsewhere.16,37 Numerical results are presented in Secs. 3.1 and 3.2. Closing remarks are presented in Sec. 4.

2.

Summary of the Model

2.1.

Physical Model

A schematic diagram of the simulated Schottky-barrier solar cell is shown in Fig. 1. The back reflector is corrugated along the x-axis, and the z axis is normal to the plane of the Schottky contact and the mean plane of the back reflector. In the remainder of this paper, the term “width” refers to extent in the x-direction, while “thickness” refers to extent in the z-direction. Insolation is provided via an excitation port at z=LAirLwLc. A planar AR window made of an insulating material occupies the region LwLc<z<Lc. To align with our earlier work, the optical permittivity of this layer was taken to be identical to that of aluminum-doped zinc oxide.38

Fig. 1

Schematic illustration of a Schottky-barrier solar cell. Only one back-reflector period is shown [i.e., (Lx/2)<x<(Lx/2)].

JPE_7_1_014502_f001.png

The region 0<z<Lz is occupied by n-doped InξGa1ξN, forming a Schottky junction and two ohmic junctions with a metal in the region 0<z<Lc. For calculations, the metal is assumed to be silver.39 The Schottky contact of width Ls is centered at x=0. The two ohmic contacts, each of width Lo/2, are centered about x=±(LxLo/2)/2. Note that Lo+Ls<Lx. The two regions between the metal contacts for 0<z<Lc are occupied with the same material as the AR window that occupies LwLc<z<Lc.

A silver back reflector, covered with an insulating window (made of the same material as the window that occupies LwLc<z<Lc), occupies the region Lz<z<Lz+Lr. This back reflector is periodically corrugated in the x-direction with period Lx. The region Lz<z<g(x) is filled with the window material, while the region g(x)<z<Lz+Lr is filled with silver. The corrugation in the unit cell is specified by the function

Eq. (1)

g(x)={Lz+daLgcos(πxζLx),2x/Lx(ζ,ζ)Lz+da,2x/Lx(ζ,ζ),
where Lx is the corrugation period, Lgda is the corrugation height, and ζ(0,1] is the duty cycle. As the corrugation is invariant along the y-axis, our model is 2-D.

Absorption of the normally incident solar flux with AM1.5G spectrum40 is calculated by solving the frequency-domain Maxwell postulates. The semiconductor charge-carrier drift-diffusion equations model the electron and hole density spatial distributions.41,42 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, mid-gap Shockley–Read–Hall, and Auger recombination are all included in our simulation. The current density J, which is averaged over the Schottky contact (or, identically, both of the ohmic contacts), is calculated for a range of external biasing voltages Vext.

By varying the proportion of indium relative to that of gallium, the bandgap of InξGa1ξN can be engineered to take any value from EgInN=0.7  eV (i.e., for InN when ξ=1) continuously through to EgGaN=3.42  eV (i.e., for GaN when ξ=0). Thus, allowing for nonhomogeneity in the z-direction, the InξGa1ξN alloy has bandgap given by

Eq. (2)

Eg0(z)=ξ(z)EgInN+[1ξ(z)]EgGaNbξ(z)[1ξ(z)],z(0,Lz),
where the bowing parameter b=1.43  eV.43 To estimate ξ for a material with a specific bandgap Eg0, Eq. (2) may be solved as

Eq. (3)

ξ(z)=b+(EgGaNEgInN)4b[Eg0(z)EgGaN]+(b+EgGaNEgInN)22b.

The electron affinity χ0 for InξGa1ξN is modeled in an analogous manner to the bandgap. Hence

Eq. (4)

χ0(z)=ξ(z)χInN+[1ξ(z)]χGaNbξ(z)[1ξ(z)],z(0,Lz),
where χInN and χGaN are the electron affinities of InN and GaN, respectively. For all other material parameters of InξGa1ξN, as described in the first column of Table 1, Vegard’s law of linear interpolation44 is assumed to apply, i.e.,

Eq. (5)

τInξGa1ξN(z)=ξ(z)τInN+[1ξ(z)]τGaN,z(0,Lz),
where τ{NC,NV,Cn,Cp,Crad,αoc}. Data for τInN and τGaN are provided in Table 1.

Table 1

Electronic data used for GaN and InN. The composition of InξGa1−ξN was estimated using Eq. (3), 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. (4).

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 narrowing of the bandgap associated with doping was incorporated through the Slotboom model.45 While this model was developed for silicon, similar narrowing behavior under heavy-doping conditions has been observed in GaN.46 When doped, the bandgap of the semiconductor narrows to Eg=Eg0ΔEg, while the electron affinity reduces to χ=χ0αocΔEg, with the conduction-band fraction αoc being a material-specific parameter. Here, the bandgap narrowing is estimated as

Eq. (6)

ΔEg=Eref{ln(Nd+NaNref)+[ln(Nd+NaNref)]2+12},
where Nd and Na are the doping densities of the donor and the acceptor, respectively, and Nref and Eref are the empirically determined reference doping density and reference energy, respectively.

The real part of the optical refractive index of InξGa1ξN is

Eq. (7)

Re{n}=Re(AA(ξ){[Eg(ξ)Eγ]2[21+EγEg(ξ)1EγEg(ξ)]+BA(ξ)}),
per the Adachi model,29 where AA and BA are parameters given in Table 1. The photon energy is Eγ=hc0/λ0, where h=6.62607004×1034  m2kgs1 is the Planck constant, c0=299792485  ms1 is the speed of light in free space, and λ0 is the free-space wavelength. The imaginary part of the refractive index is

Eq. (8)

Im{n}=αoptλ04π.
The absorption coefficient αopt is modeled in Ref. 29

Eq. (9)

αopt(ξ)=105C(ξ)(EγEg)+D(ξ)(EγEg)2nm1,
where the constants

Eq. (10)

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.47

An empirical low-field mobility model—called either the Caughey–Thomas 29 or the Arora 48 mobility model—describes the variations of the electron mobility μn and the hole mobility μp with temperature and doping. Thus

Eq. (11)

μ=μ(1)(TTref)αct,+μ(2)(TTref)βμ(1)(TTref)αct,1+[(TTref)γ(Nd+NaNcrit)]δ,{n,p},
where =n for electrons and =p for holes. In the foregoing equation, μ(2) is the maximum value of the carrier mobility μ when the material is undoped, μ(1) is the minimum value of μ when the material is heavily doped, Ncrit is the critical doping density of the InξGa1ξN alloy, T is the lattice temperature and Tref=300  K is the reference temperature, and αct,, β, γ, and δ are empirically determined parameters. The solar cell was taken to be operating at T=Tref=300  K; therefore, it should be noted that the choices of αct,, β, and γ have no effect on the results. Following Hamady et al.,29 in lieu of experimental data, we assumed that αct,=β=γ=1 for all simulations reported here. With these assumptions, the carrier mobilities simplify to

Eq. (12)

μ=μ(1)+μ(2)μ(1)1+N˜δ,{n,p},
where N˜=(Nd+Na)/Nref is the ratio of doping density to the critical doping density. The electronic data used for InξGa1ξN presented in Table 1 were provided by Hamady et al.,29 who also demonstrated that a relatively large metal work-function Φ for the Schottky-barrier contact improves efficiency. Accordingly, the relatively large value of Φ=6  eV was chosen here.

The bandgap profile of InξGa1ξN for the solar cells simulated in this study is given by

Eq. (13)

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 κ>0, φ is a phase shift, and α is a shaping parameter, which governs the profile’s gradient. Three example bandgap profiles are shown in Fig. 2 for α{0.2,1,10}, when κ=3 and φ=0.75.

Fig. 2

Bandgap profiles given by Eq. (13) for α{0.2,1,10}, when κ=3 and φ=0.75.

JPE_7_1_014502_f002.png

2.2.

Computational Model

A 2-D finite-element optoelectronic model was implemented in the COMSOL Multiphysics (V5.1) software package48 in two major steps, as described now. Let it be noted that terms in block capitals are COMSOL Multiphysics48 terms.

In the first major step, the Electromagnetic Waves, Frequency Domain module was used to calculate the 2-D EHP generation rate as a function of x(Lx/2,Lx/2] and z(0,Lz). Normally incident monochromatic light sampled at 5-nm intervals on the λ0-scale across the AM1.5G spectrum, 50% s polarized and 50% p polarized, was activated by the Periodic ports option at z=LcLwLAir. Diffraction Order ports for diffraction orders {3,2,1,1,2,3} were added. The inclusion of diffraction ports for even higher orders has little impact on the resulting EHP generation rate, predominantly due to strong absorption of shorter wavelength photons by InξGa1ξN, the majority of which are absorbed close to the surface of the device before they can be scattered. The boundaries parallel to the z-axis have Floquet periodicity with the wavevector provided by the periodic ports. The region z>Lz+L+f behind the back reflector was taken to be a perfectly matched layer. The semiconductor region was first meshed with a Mapped mesh with a 10-nm Distribution in both the z- and x-directions and then split into to a triangular distribution using the insert center points conversion. The back-reflector region was covered with an Extra fine, Delaunay, and Free Triangular mesh. Data for a spatial map of the computed EHP generation rate were stored in an external file.

The EHP generation rate can be used to compute the optical short-circuit current density JSCOpt, assuming that every absorbed photon creates an EHP in the InξGa1ξN layer and that no recombination takes place. Neglect of recombination implies that JSCOpt is necessarily larger than the 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).

In the second major step, the Semiconductor module was used to calculate the electron and hole densities in the semiconductor region, and thereby the current densities. Due to the symmetry of the unit cell, only its right half (i.e., 0xLx/2) needs to be electrically simulated. Fermi-Dirac carrier statistics, along with continuous quasi-Fermi levels at any internal boundary, were employed. Finite volume (constant shape function) discretization was employed as this inherently conserves current throughout the solar cell.48 COMSOL uses a Scharfetter–Gummel upwinding scheme for solving the charge carrier transport equations. The Free triangular, Delaunay mesh has a maximum element size of 15 nm.

A potential difference of Vext was applied between the ohmic and Ideal Schottky contacts. Thermionic currents, with standard Richardson coefficients of An*=110  AK2cm2, and Ap*=90  AK2cm2 were applied at the Schottky barrier.29,48 Insulator Interfaces were placed at the remaining external electrical boundaries. External Mathematica™ or MATLAB™ codes were used to calculate the User-Defined Generation from the output of the first major step. Recombination was incorporated via Auger, Direct and Trap-Assisted (Midgap Shockley-Read-Hall) pathways, with parameters as provided in Table 1. To facilitate convergence, the nonhomogeneity, EHP generation, and EHP recombination physics were slowly activated as the solver progressed by use of a continuation parameter. An analytic doping model was used to set the donor concentration to Nd. The Free triangular, Delaunay mesh used has a maximum element size of 15 nm.

3.

Numerical Simulations of InξGa1ξN Schottky-Barrier Solar Cells

3.1.

Design of Periodically Corrugated Back Reflector

The numerical results presented here are for devices with parameters listed in Table 2. A preliminary study was carried out to ascertain reasonable values for the dimensions of the device. Figure 3 presents JSCOpt as a function of the back-reflector period Lx(300,1000)nm for a Schottky-barrier solar cell with a homogeneous InξGa1ξN layer of thickness Lz{200,400,600,860,1000}nm. For this set of simulations, Lg=130  nm, da=135  nm, and ζ=0.5 were determined as optimal for a solar cell with a homogeneous InξGa1ξN layer characterized by Eg0=1.6  eV, Lx=600  nm, and Lz=860  nm. While this optimization is far from exhaustive, significant optical variation was seen to occur for thinner InξGa1ξN solar cells: a 6% variation in JSCOpt was calculated for a solar cell with Lz=200  nm, with a maximum JSCOpt of 24.9  mAcm2 when Lx=480  nm. Importantly, for solar cells with InξGa1ξN layers thicker than about 600 nm, JSCOpt saturates as Lx increases beyond a threshold value. Therefore, Lx=600  nm was chosen to be optimal. While the peak value of JSC of a Schottky-barrier solar cell with a 600-nm-thick InξGa1ξN layer is obtained for larger periods than this (Lx680  nm), JSC falls more rapidly for devices with back reflectors of overly large periods than for those with periods that are too small.

Table 2

Summary of parameters used for the simulations.

SymbolValue
Device widthLx600 nm
n-doped InξGa1ξN thicknessLz600 nm
Back reflector insulating layer thicknessda135 nm
Corrugation heightLg130 nm
Corrugation duty cycleζ0.5
Metal thicknessLm100 nm
Contact region thicknessLc5 nm
Ohmic contact widthLo140 nm
Schottky contact widthLs440 nm
AR window thicknessLw100 nm
Air thicknessLAir500 nm
Predoping bandgapEg0[0.7,3.42] eV
Bandgap-nonhomogeneity amplitudeA0  eV
Bandgap-nonhomogeneity phaseφ[0,1]
Bandgap-nonhomogeneity shaping parameterα(0,32)
Bandgap-nonhomogeneity periodLp>0  nm
Bandgap-nonhomogeneity ratioκ>0

Fig. 3

Optical short-circuit current density JSCOpt as a function of the back-reflector period Lx for a Schottky-barrier solar cell with InξGa1ξN layer of thickness Lz{200,400,600,860,1000}nm. The absorbing layer of InξGa1ξN is homogeneous with ξ=0.6, i.e., the bandgap Eg0=1.45  eV. Thin solid lines show JSCOpt for the equivalent solar cells with a flat back-reflector.

JPE_7_1_014502_f003.png

3.2.

Effect of Periodic Nonhomogeneity of InξGa1ξN Layer

The essential effects of including periodic nonhomogeneity in the thickness direction in the InξGa1ξN layer may be inferred from Fig. 4. For a fixed bandgap-nonhomogeneity amplitude A{0,0.2,0.4,0.6,0.8,1}, the baseline bandgap Eg* is shown to dramatically affect the simulated efficiency

Eq. (14)

η=maxVextVextJ(Vext)1000  W/m2,
which equals the maximum power produced by the solar cell as a function of Vext, divided by the incident AM1.5G solar power flux of 1000  W/m2. The solar spectrum was discretized at 5-nm λ0-intervals; with the minimum λ0 chosen to be 300 nm; the maximum λ0 chosen as min[2400,1240/(Eg*A)] nm; the other bandgap-nonhomogeneity parameters fixed at κ=3, α=2, and φ=0.75. At the minimum bandgap for an InξGa1ξN alloy, i.e., Eg=0.7  V which arises at ξ=1, there is a minimum permissible baseline bandgap Eg* for each value of A. Irrespective of the choice of value of A(0,0.8)eV, a baseline bandgap of Eg*=1.5  eV turns out to be optimal. From Fig. 4, it may be inferred that the difference in the efficiencies with a homogeneous InξGa1ξN layer and a periodically nonhomogeneous InξGa1ξN layer is greatest when

  • a. the bandgap-nonhomogeneity amplitude A is the greatest and

  • b. when the maximum bandgap Eg*1.5  eV.

Fig. 4

Efficiency η as a function of baseline bandgap Eg* for bandgap-nonhomogeneity amplitude A{0,0.2,0.4,0.6,0.8,1}eV, when κ=3, α=2, and φ=0.75.

JPE_7_1_014502_f004.png

Indeed, at Eg*=1.5  eV, the efficiency is 13.26% for A=0 and 16.81% for A=0.8  eV, i.e., the relative increase in efficiency attributable to the periodic nonhomogeneity of the InξGa1ξN layer is 26.8%, which is this paper’s most significant result. For A=1  eV, Eg*=1.7  eV is optimal, yielding an efficiency of 15.8%.

The importance of the periodicity in the bandgap variation when seeking maximal efficiency may be inferred from Figs. 5 and 6. In Fig. 5, the thickness of the InξGa1ξN layer is fixed at Lz=600  nm while the ratio κ varies from 0.5 to 4. The efficiency of the solar cell is seen to reach a maximum value of about 17% at κ=1.4 and κ=3. Between these two values of κ, only a slight decrease in η was found. Significantly, the efficiency for all values of κ>0 is greater than it is when the InξGa1ξN layer is homogeneous.

Fig. 5

Efficiency η as a function of the ratio κ=Lz/Lp for the InξGa1ξN layer when Lz=600  nm, Eg*=1.5  eV, A=0.8  eV, α=2, and φ=0.75. The horizontal red line indicates the efficiency when the InξGa1ξN layer is homogeneous.

JPE_7_1_014502_f005.png

Fig. 6

Efficiency η as a function of thickness Lz of the InξGa1ξN layer, which is either homogeneous (solid black curve) or periodically nonhomogeneous with Lp=200  nm (dashed red curve) or Lp=400  nm (dotted blue curve), when Eg*=1.54  eV, A=0.8  eV, α=5, and φ=0.75. Integer values of the ratio κ=Lz/Lp are identified.

JPE_7_1_014502_f006.png

Figure 6 illustrates the effect of varying the thickness Lz of the InξGa1ξN layer when the nonhomogeneity period Lp is fixed. Irrespective of the period, the solar cell with the nonhomogeneous InξGa1ξN layer performs better than the one with a homogeneous InξGa1ξN layer. Furthermore, the solar cell with the smaller period (Lp=200  nm) is found to perform more efficiently than the one with larger period (Lp=400  nm). Distinct maximums can be seen at κ=1.5 and κ=3, which correspond to Lz=300  nm and Lz=600  nm, respectively. The efficiencies at these points, respectively, are 17.28% and 16.88%, which correspond to relative increases in efficiency of 16.7% and 28.5% as compared to the analogous solar cell with a homogeneous InξGa1ξN layer.

By varying the bandgap-nonhomogeneity shaping parameter α, the spatial profile of the InξGa1ξN nonhomogeneity can be varied. Thus, α=1 holds for a sinusoidal bandgap profile, while smaller and larger values of α yield bandgap profiles with steeper gradients, as illustrated in Fig. 2. For all three values of α in Fig. 7, the efficiency of the solar cell is substantially greater than it is for the corresponding solar cell with a homogeneous InξGa1ξN layer. The maximum increase in efficiency is found for a nearly sinusoidal bandgap profile; indeed, the efficiency increases from 13.2% for a Schottky-barrier solar cell with a homogeneous InξGa1ξN layer to 16.88% for its analog with a nonhomogeneous InξGa1ξN layer with α=2, which is a relative increase in efficiency of 28.5%.

Fig. 7

Efficiency η plotted against bandgap-nonhomogeneity shaping parameter α, when Eg*=1.54  eV, A=0.8  eV, Lz=600  nm, κ=3, and φ=0.75.

JPE_7_1_014502_f007.png

Let us note here that as α becomes much larger or smaller than unity, the spatial gradients of the bandgap increase in magnitude and the peaks in Fig. 2 become narrower. Accordingly, at extreme values of α, the semiclassical carrier-transport equations implemented in our model become less appropriate as quantum processes become significant. Then, it may be necessary to take into account the quantization of allowed energy states between the peaks of Eg0(z) and tunneling through those peaks, especially for large amplitudes A.

4.

Closing Remarks

A 2-D finite-element model was devised to simulate the combined optical and electrical performances of InξGa1ξN Schottky-barrier solar cells. First, it was found that a periodically corrugated back reflector of a period of 600 nm is optimal for photon absorption in a solar cell containing a homogeneous layer of InξGa1ξN. Second, the effects of periodic nonhomogeneity of the InξGa1ξN layer were elucidated. The nonhomogeneity was directed perpendicular to the mean plane of the periodically corrugated back reflector. For the particular model investigated here, the efficiency of a solar cell with a 600-nm-thick layer of InξGa1ξN was found to increase by 26.8% when suitable periodic nonhomogeneity was incorporated. Thus, the incorporation of a periodically nonhomogeneous InξGa1ξN layer in a Schottky-barrier solar cell can substantially increase the efficiency as compared to the analogous solar cell with a homogeneous InξGa1ξN layer. A comprehensive optimization of study of material and design parameters may well yield even greater nonhomogeneity-induced increases in efficiency. However, such an optimization study—which is justified by the substantial efficiency boosts reported herein—represents a major undertaking that lies beyond the scope of this paper. Similarly, it would be of some value to delineate the photon-absorption mechanisms that underpin the boost in the light-to-electricity conversion efficiency that arises following the introduction of periodic nonhomogeneity, but this too is matter for future study.

The feasibility of producing InξGa1ξN alloys with the prescribed spatial variation in ξ to achieve efficiency boosts in Schottky-barrier solar cells is a matter for our experimentalist colleagues to shed light on. On the basis of our numerical study and the relatively huge boosts in efficiency that may be attained, it would be worthwhile for major efforts to be directed toward the production of appropriate InξGa1ξN alloys.

Last, the principal finding of our study is that the introduction of periodic nonhomogeneity can, in principle, substantially boost the efficiency of a Schottky-barrier solar cell. To focus on efficiency, our attention was restricted to only normally incident solar radiation for all cell configurations considered. The influence of the angle of incidence is planned for a future study in which the mechanisms for efficiency boosting will be more fully explored.

Acknowledgments

T.H.A. thanks the Charles Godfrey Binder Endowment for partial financial support during a six-month stay at Pennsylvania State University. T.G.M. acknowledges the support of EPSRC Grant No. EP/M018075/1. A.L. thanks the National Science Foundation for partial financial support under Grant No. DMS-1619901, and he is grateful to the Charles Godfrey Binder Endowment for ongoing support of his research.

References

1. 

R. Singh, G. F. Alapatt and A. Lakhtakia, “Making solar cells a reality in every home: opportunities and challenges for photovoltaic device design,” IEEE J. Electron Devices Soc., 1 129 –144 (2013). http://dx.doi.org/10.1109/JEDS.2013.2280887 Google Scholar

2. 

P. Verlinden et al., “The surface texturization of solar cells: a new method with V-grooves with controllable sidewall angles,” Sol. Energy Mater. Sol. Cells, 26 71 –78 (1992). http://dx.doi.org/10.1016/0927-0248(92)90126-A SEMCEQ 0927-0248 Google Scholar

3. 

N. Yamada et al., “Characterization of antireflection moth-eye film on crystalline silicon photovoltaic module,” Opt. Express, 19 A118 –A125 (2011). http://dx.doi.org/10.1364/OE.19.00A118 OPEXFF 1094-4087 Google Scholar

4. 

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). http://dx.doi.org/10.1063/1.94432 APPLAB 0003-6951 Google Scholar

5. 

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

6. 

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). http://dx.doi.org/10.1364/AO.52.000966 APOPAI 0003-6935 Google Scholar

7. 

S. K. Dhungel et al., “Double-layer antireflection coating of MgF2/SiNx for crystalline silicon solar cells,” J. Korean Phys. Soc., 49 885 –889 (2006). KPSJAS 0374-4884 Google Scholar

8. 

S. A. Boden and D. M. Bagnall, “Sunrise to sunset optimization of thin film antireflective coatings for encapsulated, planar silicon solar cells,” Prog. Photovoltaics Res. Appl., 17 241 –252 (2009). http://dx.doi.org/10.1002/pip.v17:4 PPHOED 1062-7995 Google Scholar

9. 

D. M. Schaadt, B. Feng and E. T. Yu, “Enhanced semiconductor optical absorption via surface plasmon excitation in metal nanoparticles,” Appl. Phys. Lett., 86 063106 (2005). http://dx.doi.org/10.1063/1.1855423 APPLAB 0003-6951 Google Scholar

10. 

J.-Y. Lee and P. Peumans, “The origin of enhanced optical absorption in solar cells with metal nanoparticles embedded in the active layer,” Opt. Express, 18 10078 –10087 (2010). http://dx.doi.org/10.1364/OE.18.010078 OPEXFF 1094-4087 Google Scholar

11. 

P. Shokeen, A. Jain and A. Kapoor, “Silicon nanospheres for directional scattering in thin-film solar cells,” J. Nanophotonics, 10 036013 (2016). http://dx.doi.org/10.1117/1.JNP.10.036013 1934-2608 Google Scholar

12. 

R. A. Sinton et al., “27.5-percent silicon concentrator solar cells,” IEEE Electron Device Lett., 7 567 –569 (1986). http://dx.doi.org/10.1109/EDL.1986.26476 EDLEDZ 0741-3106 Google Scholar

13. 

K. Nishioka et al., “Evaluation of InGaP/InGaAs/Ge triple-junction solar cell and optimization of solar cells structure focusing on series resistance for high-efficiency concentrator photovoltaic systems,” Sol. Energy Mater. Sol. Cells, 90 1308 –1321 (2006). http://dx.doi.org/10.1016/j.solmat.2005.08.003 SEMCEQ 0927-0248 Google Scholar

14. 

M. E. Solano et al., “Buffer layer between a planar optical concentrator and a solar cell,” AIP Adv., 5 097150 (2015). http://dx.doi.org/10.1063/1.4931386 AAIDBI 2158-3226 Google Scholar

15. 

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

16. 

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). http://dx.doi.org/10.1117/1.JPE.6.025502 Google Scholar

17. 

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 093593 (2015). http://dx.doi.org/10.1117/1.JNP.9.093593 1934-2608 Google Scholar

18. 

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). http://dx.doi.org/10.1109/ICIP.2013.6738391 Google Scholar

19. 

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

20. 

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

21. 

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

22. 

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). http://dx.doi.org/10.1117/1.OE.52.8.087106 Google Scholar

23. 

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

24. 

S. M. Iftiquar et al., “Single- and multiple-junction p-i-n type amorphous silicon solar cells with paSi1xCx:H and nc-Si:H films,” Photodiodes—From Fundamentals to Applications, InTech, Rijeka, Croatia (2012). Google Scholar

25. 

P. T. Landsberg and C. Klimpke, “Theory of the Schottky barrier solar cell,” Proc. R. Soc. A: Math. Phys. Eng. Sci., 354 101 –118 (1977). http://dx.doi.org/10.1098/rspa.1977.0058 PRLAAZ 1364-5021 Google Scholar

26. 

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., 12 765 –771 (1979). http://dx.doi.org/10.1088/0022-3727/12/5/018 JPAPBE 0022-3727 Google Scholar

27. 

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

28. 

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

29. 

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

30. 

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

31. 

J. Wu et al., “Superior radiation resistance of InxGa1-xN alloys: full-solar-spectrum photovoltaic material system,” J. Appl. Phys., 94 6477 –6482 (2003). http://dx.doi.org/10.1063/1.1618353 JAPIAU 0021-8979 Google Scholar

32. 

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

33. 

Y. Saito et al., “Growth temperature dependence of Indium Nitride crystalline quality grown by RF-MBE,” Physica Status Solidi B, 234 796 –800 (2002). http://dx.doi.org/10.1002/(ISSN)1521-3951 Google Scholar

34. 

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

35. 

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

36. 

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., 806 968 –970 (2002). http://dx.doi.org/10.1063/1.1450255 APPLAB 0003-6951 Google Scholar

37. 

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

38. 

X. Y. Gao, Y. 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). http://dx.doi.org/10.3938/jkps.57.710 KPSJAS 0374-4884 Google Scholar

39. 

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

40. 

“Reference solar spectral irradiance: air mass 1.5,” (2015) http://rredc.nrel.gov/solar/spectra/am1.5/ July ). 2015). Google Scholar

41. 

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

42. 

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

43. 

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

44. 

L. Vegard, “Die Röntgenstrahlen im Dienste der Erforschung der Materie,” Z. Kristallogr., 67 239 –259 (1928). http://dx.doi.org/10.1524/zkri.1928.67.1.239 Google Scholar

45. 

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

46. 

I.-H. Lee et al., “Band-gap narrowing and potential fluctuation in Si-doped GaN,” Appl. Phys. Lett., 74 102 –104 (1999). http://dx.doi.org/10.1063/1.122964 APPLAB 0003-6951 Google Scholar

47. 

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

48. 

( (2015) http://www.comsol.com/ January ). 2015). Google Scholar

Biography

Tom H. Anderson received his BSc degree in mathematics and physics from the University of Edinburgh in 2011. He received his MSc degree in fusion energy from the University of York in 2012 and his PhD in applied mathematics from the University of Edinburgh in 2016. His research interests include the optical and electrical modeling of thin-film solar cells.

Tom G. Mackay is a reader at the School of Mathematics, the University of Edinburgh, and an adjunct professor in the Department of Engineering Science and Mechanics at Pennsylvania State University. He is a graduate of the Universities of Edinburgh, Glasgow, and Strathclyde and a fellow of the Institute of Physics (UK) and SPIE. His research interests include the electromagnetic theory of complex materials, including homogenized composite materials.

Akhlesh Lakhtakia received degrees from the Banaras Hindu University and the University of Utah. He is the Charles Godfrey Binder professor of engineering at the Pennsylvania State University. His research interests include surface multiplasmonics, bioreplication, forensic science, solar cells, sculptured thin films, and mimumes. He is 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, and Royal Society of Chemistry. He received the 2010 SPIE Technical Achievement Award and the 2016 Walston Chubb Award for Innovation.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Tom H. Anderson, Tom G. Mackay, and Akhlesh Lakhtakia "Enhanced efficiency of Schottky-barrier solar cell with periodically nonhomogeneous indium gallium nitride layer," Journal of Photonics for Energy 7(1), 014502 (3 February 2017). https://doi.org/10.1117/1.JPE.7.014502
Received: 8 December 2016; Accepted: 18 January 2017; Published: 3 February 2017
Lens.org Logo
CITATIONS
Cited by 17 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Gallium

Solar cells

Reflectors

Gallium nitride

Doping

Indium nitride

Energy efficiency


CHORUS Article. This article was made freely available starting 03 February 2018

Back to Top