Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Simulations of sinusoidal nanotextures for coupling light into c-Si thin-film solar cells

Open Access Open Access

Abstract

We numerically study coupling of light into silicon (Si) on glass using different square and hexagonal sinusoidal nanotextures. After describing sinusoidal nanotextures mathematically, we investigate how their design affects coupling of light into Si using a rigorous solver of Maxwell’s equations. We discuss nanotextures with periods between 350 nm and 1050 nm and aspect ratios up to 0.5. The maximally observed gain in the maximal achievable photocurrent density coupled into the Si absorber is 7.0 mA/cm2 and 3.6 mA/cm2 for a layer stack without and with additional antireflective silicon nitride layers, respectively. A promising application is the use as smooth anti-reflective coatings in liquid-phase crystallized Si thin-film solar cells.

© 2016 Optical Society of America

1. Introduction

The absorber thickness of wafer-based crystalline silicon (c-Si) solar cells will continue to decrease, as predicted by the International Technology Roadmap for Photovoltaic [1]. Consequently, also c-Si solar cells based on thin-film approaches are becoming increasingly important with the potential to compete with wafer-based c-Si technologies. Further, they can combine the advantages of thin-film technology with the high material quality of crystalline silicon.

An important approach to build c-Si thin-film solar cells is the transfer method, where a c-Si epitaxial layer is grown onto a porous layer on a Si wafer such that it can be detached from the wafer and transferred onto another substrate [2–4]. For this approach efficiencies of 19% are reported for a 43 µm thick absorber [2]. Further, wafers thinned to several tenths of micrometers are used to make solar cells. While this approach is not directly applicable to industry, it is very valuable as it allows to explore the efficiency limits of such thin solar cells theoretically and experimentally [5–7]. In 2015, 15.7% on a cell with a 10 µm thick absorber was demonstrated [7].

In this work, we focus on structures that are used in thin-film solar cells with a liquid-phase-crystallized silicon (LPC-Si) absorber [8–11]. Because the c-Si absorber in LPC-Si solar cells is present directly on the glass substrate, no handling issues arise in contrast to wafer-based technologies, where the probability for breakage increases when the wafers get thinner and thinner. The current efficiency record for this technology is 12.1% [12]. For making LPC-Si films, an approximately 10 µm thick nanocrystalline Si film is deposited onto a glass substrate and subsequently crystallized by melting using an electron beam or a laser. The crystallization process leads to grain sizes comparable to multi-crystalline silicon used in industry [10].

c-Si thin-film solar cells with an absorber thickness in the range of 10 µm require advanced light management. Besides light trapping, which is very important because of the weak absorptivity of Si due to its indirect bandgap, as much light as possible has to be coupled into the absorber. This in-coupling is especially important, because Si has a high refractive index and hence a large contrast with the surrounding materials. This leads to a high reflectivity, especially when flat interfaces are used. In-coupling of light into LPC-Si devices can be enhanced using nanotextures at the glass-silicon interface that are manufactured with nanoimprint lithography, which is explained in detail in [13,14].

In recent years, 2-dimensional (2D) textures for light trapping have been heavily investigated: often, they are studied at the rear side of solar cells for light trapping [15,16], on the front side for in-coupling [17,18], or on both sides [19,20]. All these 2D textures have sharp edges making them incompatible with LPC-Si technology, because the textures used for in-coupling have to be implemented prior to the crystallization process. It has been shown that sharply-edged nanotextures on the front side induce a very high defect density in the LPC-Si film, which has a detrimental effect on the open circuit voltage and also on the current density – despite a strongly increased absorption due to the nanotexture [11,21]. Hence, in-coupling must be done with gentle nanostructures that allow the crystallization of LPC-Si with a high electric quality.

To combine good in-coupling of the incident light with high material quality, we look at sinusoidal gratings that are very smooth and gentle. In this work we numerically investigate how reflection from Si interfaces can be minimized by using such sinusoidal gratings. After discussing the mathematical description of such gratings and some general properties of two-dimensional hexagonal gratings, we mention important details of our simulations. Then, we discuss the results for two different layer stacks: a simple layer stack, which is very well suited for studying the physics of in-coupling in detail, and a layer stack with additional antireflective layers, as it is used for real LPC-Si solar cells.

2. Mathematical description of sinusoidal nanotextures

According to group theory five types of 2D lattices exist, of which we investigate sinusoidal nanotextures for the two types with the highest symmetry: square and hexagonal.

2.1. Square lattices

A sinusoidal square lattice can be described by multiplying a cosine defined along the x direction with a cosine along the y direction,

fsq(x,y)=cosxcosy,
as illustrated in Fig. 1(a). Replacing one or both cosines by sines does not change the morphology, but merely leads to a lateral shift of the texture. The border of the unit cell chosen for computation is depicted in Fig. 1(a) as well.

 figure: Fig. 1

Fig. 1 Illustrating different sinusoidal gratings in a 3D (top) and a contour plot representation (bottom): (a) a square lattice constructed according to Eq. (1). Three hexagonal lattices constructed according to Eq. (3) for different structure phases ϕ: (b) a positive cosine lattice (“cos”) with ϕ = 0, (c) a sine lattice (“sin”) with ϕ = π/2, and (d) a negative cosine lattice (“−cos”) with ϕ = π. The red quadrilaterals enclose the unit cell.

Download Full Size | PDF

Before we can apply sinusoidal interfaces to our simulations, we have to scale them laterally and vertically such that the period of the texture, i. e. the unit cell side length, is equal to a desired value P and that the maximal peak-to-valley height has the desired value h. For square lattices the period is set via the following substitutions in Eq. (1):

x2πPx,y2πPy.

The maximal peak-to-valley height in Eq. (1) is max(fsq) − min(fsq) = 2. Hence, we can set the nanotexture to the desired maximal peak-valley height h by multiplying Eq. (1) with h/2.

2.2. Hexagonal lattices

Hexagonal sinusoidals are described by multiplying three cosines with each other,

fhex(x,y)=cos[12(x+3y)]cos[12(x3y)]cos(x+ϕ),
where ϕ is a phase. The first two factors in Eq. (3) constitute a rhombus-shaped lattice; multiplying with the third factor leads to the creation of a hexagonal grating. By tuning the phase ϕ we can generate a variety of different hexagonal textures. In this manuscript we look at three examples: ϕ = 0 leads to a texture, which we call positive cosine (“cos”) [Fig. 1(b)], ϕ = π/2 to a sine (“sin”) texture [Fig. 1(c)] and ϕ = π to a negative cosine (“−cos”) texture [Fig. 1(d)]. When defining the hexagonal texture as in Eq. (3), the unit cell can be chosen such that its borders always have a height of exactly fhex(unit cell border) 0, regardless of ϕ. This is illustrated in Figs. 1(b)–1(d) as well.

We also scale the hexagonal nanotextures in order to get the desired period P and peak-to-valley height h. The period can be set with a substitution similar to Eq. (2),

x2π3Px,y2π3Py.

For hexagonal sinusoidals the maximal peak-to-valley height depends on the phase ϕ. The minimum and maximum of Eq. (3) are given by

min(fhex,ϕ)=14[cosϕ3sin(ϕ3+π6)],
min(fhex,ϕ)=14[cosϕ+3cos(ϕ3)].

After some simplifications we obtain for the maximal peak-to-valley height

hminmax(ϕ):=max(fhex,ϕ)max(fhex,ϕ)=334sin(ϕ3+π3).

For the positive and negative cosine structures, we have hminmax(0)=hminmax(π)=9/8, while the value for sine structures is hminmax(π/2)=33/41.299. We can set hexagonal nanotextures to the desired maximal peak-to-valley height h by multiplying with h/hminmax.

We also apply a vertical shift to the nanotextures such that the minimum is always at zero level. This is done to ensure that the layer thickness above and below the nanotextures is constant for all simulations.

3. Investigated models

We study two different layer stacks: firstly, a simple layer stack illustrated in Fig. 2(a), which allows us to study the effect of the different geometries on the in-coupling in detail. It consists of a glass substrate from which the light is incident. It is covered by a sol-gel layer that carries the sinusoidal nanotexture. Here, we account for the experimental conditions, where textures are transferred onto the sol-gel using nanoimprint lithography [8]. The thickness of the sol-gel layer at the lowest point of the nanotexture is 200 nm. The last layer is the c-Si absorber.

 figure: Fig. 2

Fig. 2 The layer stack and the mesh used in this work shown for a positive cosine (“cos”) texture with P = 500 nm and h = 250 nm, (a) for the simple layer stack and (b) a layer stack with additional antireflective (AR) SiOx and SiNx layers.

Download Full Size | PDF

Secondly, we perform simulations on a layer stack with additional anti-reflective (AR) layers as it is used in experimental LPC-Si solar cell structures [10]. As illustrated in Fig. 2(b), onto the front glass a 250 nm thick silicon oxide (SiOx) layer is deposited, which is covered by a sol-gel layer as described above. Onto the sol-gel, a 70 nm thick silicon nitride (SiNx) layer, another 20 nm thick SiOx layer and the Si absorber are deposited.

To simulate coupling of light into the material stack we solve Maxwell’s equations on the 3D unit cell of the grating with appropriate source, material and boundary condition settings, using a time-harmonic finite-element solver (JCMsuite) [22]. As illustrated in Fig. 2, the layer stacks used in the simulations are built from a tetrahedral mesh.

A convergence study was performed to determine the optimal parameter settings for reaching a sufficient accuracy: the mesh side length constraint of the nanotextured sinusoidal interface is set to 50 nm and the maximal edge length of the tetrahedra is 100 nm. Polynomial degrees between 2 and 4 are employed.

Because we want to study exclusively how the nanotextures affect in-coupling of light, scattering effects arising from the back side of a Si absorber with finite thickness must be omitted. Therefore, we treat the front glass and the Si absorber as infinite half spaces, which we realize by using perfectly matched layers (PML) on top and bottom as transparent boundary conditions. The light is incident from the glass half space. All the light that is absorbed in the Si layer as well as light that propagates into the PML covering the Si layer is regarded as coupled into the absorber. We keep the thicknesses of the glass layer and Si layer (above the nanotexture) at 100 nm and 200 nm, respectively. Changing volumes of the thin Si layer due to changing nanotextures do not affect the results, because we treat the Si absorber as an infinite half space. On the sides of the unit cell, we employ periodic boundary conditions.

The complex refractive index data for sol-gel, SiOx and SiNx were obtained with ellipsome-try. SiOx was determined to be completely transparent between 300 nm and 1200 nm, and both SiNx and sol-gel do not absorb at wavelengths longer than 340 nm. Refractive index data for crystalline Si and glass were provided by Helmholtz-Zentrum Berlin. In the simulations, glass is considered to be non-absorptive, hence the imaginary part of its refractive index is omitted.

Besides absorption spectra, we use the maximum achievable photocurrent density Jphmax as a figure of merit. In this work, Jphmax is the short-circuit current density that would be achieved with a solar cell with an infinitely thick c-Si absorber if all the light coupled into the absorber would lead to the generation of electron-hole pairs under illumination with the AM1.5 solar spectrum [23]. It is calculated with

Jphmax=eλ1λ2Ainc(λ)ΦAM1.5(λ)cosθindλ,
where Ainc is the fraction of the incident light that is coupled into the silicon layer, ΦAM1.5 is the photon flux according to the AM1.5 spectrum and θin is the angle between the direction of the incident light and the solar cell normal. The wavelength λ1 corresponds to the absorption onset of the absorber (300 nm) and λ2 corresponds to the bandgap of c-Si (1107 nm).

Ainc is determined via calculating the net energy flow into the absorber,

Ainc=1PinV(Si)SndA,
where Pin is the power incident onto the unit cell, S denotes the Poynting vector, n is the surface normal, and ∂V(Si) denotes the boundary of the Si absorber in the unit cell. Because the sides of the unit cell are constructed with periodic boundaries, the net flow through these surfaces cancels out. Further, no energy flow will occur through the (imaginary) back surface of the Si absorber at infinity. Hence, for calculating Ainc it is sufficient to integrate S n across the nanotextured interface between the dielectric and the Si absorber.

We take reflection losses of the front air-glass interface into account by multiplying all the absorptivities in the different layers with (1 − Rair–glass), where Rair–glass is calculated with the Fresnel equations and has a value of about 4% at normal incidence. In this study, we neglect absorptive losses in the glass layer, as already mentioned above.

4. Results

As mentioned in Section 3, we discuss results for a simple layer stack and one with additional antireflective layers as it is used in experiments. While studying the simple layer stack enables us to carefully investigate how the different sinusoidal nanotextures affect the coupling of the incident light into the Si absorber, the additional AR layers in the other stack affect the coupling as well. Hence, both coupling due to the nanotexture and due to the AR layers occur and might affect each other. Therefore, studying only the effect of the nanotexture on coupling the light into absorber is not possible in the layer stack with AR layers.

In this section, we first discuss results on the simple layer stack for normal and oblique incidence and investigate how the observed results compare to scattering theory for 2D hexagonal gratings. Secondly, we discuss results on the stack with AR layers.

4.1. Simple layer stack (without AR layers)

4.1.1. Normal incidence

Figure 3(a) shows the maximum achievable current density Jphmax for c-Si absorbers on a glass substrate covered with nanotextured sol-gel with different periods P and a constant aspect ratio of a = h/P = 0.5. Light is incident from the glass side, as illustrated in Fig. 2. The results are shown for a square lattice and three different hexagonal lattices: with a positive cosine (“cos”), a sine (“sin”), and a negative cosine (“−cos”) structure. We see that the differences between the four geometries are significant. The negative cosine texture couples light into the absorber strongest. Periods between P = 500 nm and 600 nm show the strongest coupling of light into the absorber for all four geometries.

 figure: Fig. 3

Fig. 3 Simulation results for the simple layer stack without AR layers [see Fig. 2(a)] for the four different geometries presented in Fig. 1 and a constant aspect ratio of a = h/P = 0.5. (a) The maximum achievable current density Jphmax for five different periods and (b) the spectra of the in-coupled light for the three hexagonal structures and P = 500 nm are shown. A flat reference and, in (a), the current achieved with full absorption are shown as well. All results are corrected for the reflection loss at the air-glass interface (about 4%).

Download Full Size | PDF

The square lattice shows a similar trend as the hexagonal sine lattice, but exhibits lower current densities for all periods except 350 nm. Because of its comparatively weak performance, we will not consider the square lattice for the rest of this study. We attribute the stronger performance of the hexagonal textures with respect to the square textures to the six channels into which light can couple per diffraction order. In contrast, for square lattices only four channels per diffraction order are available.

The significant differences between the performance of the three hexagonal structures are explored in Fig. 3(b), which shows spectra of the fraction of light coupled into Si for the three hexagonal geometries at a period of P = 500 nm. The most prominent features are the sharp peaks at approx. 330 nm, 380 nm and 650 nm and the valleys between approx. 400 nm and 600 nm. The negative cosine lattice shows the highest values across the whole spectrum; especially the valley at short wavelengths is most shallow for this geometry.

Figure 4 shows the local distribution of the imaginary part of the electric field energy density 1/4ℑ(E·D*), which is proportional to the absorbed power density [24]. Especially the negative cosine structure, shown in Fig. 4(c), exhibits a very strong focus in absorption with maximum densities almost twice as high as for the positive cosine structure shown in Fig. 4(a). Also for the sine (Fig. 4(b)) a focus is visible, which however is weaker than for the negative cosine. It therefore seems that the negative cosine texture functions as an array of convex microlenses, which might help us to understand why this texture couples light better into the Si than the other textures. However, the connection between the stronger in-coupling and the microlensing effect has yet to be investigated more in depth.

 figure: Fig. 4

Fig. 4 The imaginary part of the electric field energy density 14(ED*), which is proportional to the absorbed power density. The simulation results are shown for the simple layer stack for the three hexagonal geometries with P = 500 nm and h = 250 nm at λ = 500 nm along cross sections through the unit cells, as they are depicted in Figs. 1(b)–1(d).

Download Full Size | PDF

4.1.2. Analysis using diffraction theory

To analyze the origin of the distinct peaks observed in Fig. 3(b) at wavelengths of approx. 330 nm, 380 nm and 650 nm, we look at the Fourier transform of the field reflected back into the glass. Because of the perfect periodicity of the unit cell, light is scattered into well-defined diffraction channels. Figure 5(a) shows the number of channels diffracted back into the glass as a function of the wavelength for a hexagonal grating with period P = 500 nm: for very short wavelength 19 channels are present. At approx. 330 nm, 380 nm and 650 nm 6 channels vanish, such that only one channel (the zeroth diffraction order) is present above 650 nm. In other words, above these wavelengths the third, second, and first diffraction orders vanish, respectively.

 figure: Fig. 5

Fig. 5 (a) The number of diffracted channels reflected back into the glass for a period of P = 500 nm at 330 nm, 380 nm, 650 nm and 1050 nm. To relate these results to the fraction coupled into Si (FCiSi) shown in Fig. 3(b), the curve for the negative cosine (“−cos”) is shown as well. (b) The reciprocal lattice of a hexagonal lattice with period P. The nearest, second-nearest and third-nearest neighbors of an arbitrary point are depicted, which correspond to the first, second and third diffraction orders, respectively.

Download Full Size | PDF

To understand why the diffraction orders vanish at these distinct wavelengths we take a look at the scattering theory for hexagonal two-dimensional gratings. The primitive vectors of the reciprocal lattice for a hexagonal lattice with period P are given by

g1=4π3P(12,32),g2=4π3P(12,+32)

For diffracted waves the k-vector component k parallel to the lattice plain is equal to a vector connecting two points of the reciprocal lattice. As illustrated in Fig. 5(b), the first, second and third diffraction orders corresponds to the nearest, second-nearest, and third-nearest neighbors, respectively:

k1=|g1|=|g2|=4π3Pg,k2=3g,k3=2g.

The angle sinθi into which the i-th refraction order is scattered is given by

sinθi=kik=ki=λ2πn,
with the wavenumber k and the refractive index n of the medium into which the light is scattered (glass, in our case). The i-th diffraction order vanishes when sinθi 1, hence
λi=2πnki=.

Using ki from Eq. (10) in Eq. (12), we obtain λ1 ≈ 650 nm, λ2 ≈ 375 nm and λ3 ≈ 325 nm for a period of = 500 nm, where we approximate the refractive index of glass by nglass 1.5. These λi values are in agreement with the peaks of the spectra shown in Fig. 3(b).

4.1.3. Oblique incidence

It is also very important to study how structures for light management operate under oblique incidence [25, 26]. We study the behavior of our sinusoidal nanotextures at oblique incidence using the simple layer stack, where we take the angle of incidence of light in the glass layer, θglass, as parameter. In contrast to the results for normal incidence discussed above, here we do not take the effects of the air-glass interface into account. According to Snell’s law, the maximum angle θglass for light incident via a flat air-glass interface would be 41°. However, with (nano)structured anti-reflective coatings between air and glass, also angles > 41° can be achieved. We discuss results for the negative cosine (“−cos”) with P = 500 nm and h = 250 nm and for a flat layer stack.

When studying oblique incidence, also the azimuth ψ of the incident wave has to be taken into account. We chose to average across all azimuths with steps of 15°. Because of the rotational symmetry of the negative cosine texture it is sufficient to simulate for three azimuths: ψ = 0°, 15° and 30°. To average for all azimuths, a weighted average was taken where ψ = 0° and 30° accounted for 25% each and ψ = 15° accounted for 50%. Finally, the average for both polarizations was taken, just as for normal incidence.

In Fig. 6 the maximum achievable photocurrent densitiy multiplied with 1/cosθglass is shown, where we take into account that the AM1.5 spectrum is reduced by this cosine factor for oblique incidence. As a reference, we plot the current density achieved when all the light is absorbed. For small angles < 60° the negative cosine clearly outperforms the flat cases. However, for θglass = 75°, the flat structure couples light better into the Si.

 figure: Fig. 6

Fig. 6 The maximum achievable photocurrent densitiy Jphmax multiplied with 1/cosθglass for oblique incidence in the simple layer stack for a negative cosine geometry (“−cos”) with P = 500 nm and h = 250 nm and for a flat layer stack.

Download Full Size | PDF

We can understand why the flat structure functions better than the nanotextured one at θglass = 75° by looking at the refractive index of sol-gel, which is lower than that of glass: e.g. at 600 nm, nglass = 1.51 and nsol-gel = 1.44. While this small difference has only very little effect for normal incidence, at θglass = 75° total internal reflection occurs and only evanescent waves can penetrate into the sol-gel layer. While the sol-gel thickness in the flat sample is 200 nm everywhere, for the nanotextured sample the sol-gel thickness is 200 nm only at the minima of the nanotexture. Hence, for the flat case coupling of the evanescent waves into Si is stronger.

4.2. Layer stack with additional AR layers

As already mentioned in Section 3, experimentally a layer stack with additional SiOx and SiNx AR layers is employed in order to reduce reflective losses. This stack is illustrated in Fig. 2(b). Here, we analyze, how strong this layer stack affects in-coupling, when sinusoidally nanotex-tured interfaces are used.

Figure 7 shows simulation results for negative cosine (“−cos”) nanostructures with a period of P = 500 nm and different aspect ratios a and hence structure heights h. In Fig. 7(a) the maximum achievable photocurrent density Jphmax is shown. For all four different structure heights Jphmax increases when adding the AR layers. This gain is by far largest for the flat stacks with about 5 mA/cm2 and gets smaller with increasing h. At h = 250 nm about 1.7 mA/cm2 are gained when adding the AR layers. Note that the h = 125 nm structure with the complex layer stack yields almost the same Jphmax as the simple layer stack with h = 250 nm. Hence, using the complex layer stack can strongly reduce the required height for effective in-coupling. As a consequence, too high aspect ratios, which could deteriorate the electronic silicon material quality, are not necessary.

 figure: Fig. 7

Fig. 7 (a) Jphmax and (b) fraction of light coupled into Si for the simple layer stack and the complex layer stack with additional AR layers. The results are shown for the negative cosine geometry (P = 500 nm) for different aspect ratios a (or structure heights h).

Download Full Size | PDF

In Fig. 7(b) the fraction coupled into Si is shown for the flat and h = 250 nm cases. We observe that the effect of antireflective layers stack becomes smaller if h becomes larger. For h = 250 nm the complex layer stack is very effective to remove the valley between 400 and 600 nm. But also for wavelengths longer than 650 nm the absorbed fraction is higher than for the simple layer stack.

5. Conclusions and outlook

Liquid-phase crystallized silicon (LPC-Si) thin-film solar cells require advanced light-management concepts to maximise the absorption in the absorber layer. In this study we numerically investigated sinusoidal nanotextures that enhance the fraction of light coupled into the Si absorber from the glass side. This investigation was performed for square and hexagonal sinusoidal nanotextures.

Not only the period P and aspect ratio a of hexagonal sinusoidal gratings determines the fraction of light coupled into Si, but also their morphology: for all periods, the negative cosine (“−cos”) nanotexture performs best. For all nanotextures, the optimal period is between 500 nm and 600 nm. The best structure presented in this paper is a “−cos” of P = 500 nm and a = 0.5 with SiNx and SiOx antireflective (AR) layers. With this structure the maximal achievable photocurrent density Jphmax is increased by 3.6 mA/cm2 with respect to the flat layer stack with AR layers, which is the current standard for the LPC-Si technology. For the simple layer stack without the AR layers, the gain in Jphmax is even 7.0 mA/cm2. We observed that the hexagonal “−cos” nanotextures, which perform strongest, exhibit a microlensing effect with a strongly increased absorbed power density density in its focal region. Except for the shortest period, square sinusoidal structures are outperformed by all three hexagonal geometries.

Also for oblique incidence, the “−cos” nanotexture with P = 500 nm and a = 0.5 enhances in-coupling of light into the silicon layer with respect to the flat reference. However, the small difference in the refractive index of glass and sol-gel reduces the fraction of light coupled into Si at large angles of incidence due to total internal reflection. While currently the sol-gel is required as a carrier of the nano-imprinted texture it might be an advantage to transfer the nanotexture into the glass carrier or the SiOx layer.

This work directly impacts manufacturing efficient c-Si thin-film solar cells, as it enables us to define the smooth nanotextures that couple light into silicon best. However, the subtle balance between optical enhancement versus electrical deterioration due to increasing aspect ratio is yet to be investigated experimentally.

Acknowledgments

We acknowledge Florian Ruske for helping us with the ellipsometry measurements required for obtaining refractive index data. The results were obtained at the Berlin Joint Lab for Optical Simulations for Energy Research (BerOSE) of Helmholtz-Zentrum Berlin fur¨ Materialien und Energie, Zuse Institute Berlin and Freie Universit¨at Berlin. The German Federal Ministry of Education and Research is acknowledged for funding the research activities of the Nano-SIPPE group within the program NanoMatFutur (No. 03X5520) and for funding within program Solar-Nano (13N13164). Further we acknowledge support by the Einstein Foundation Berlin through ECMath within subprojects SE6 and OT5. We thank David Eisenhauer and Grit Köppel for the useful discussions.

References and links

1. International Technology Roadmap for Photovoltaic (ITRPV) – 2014 Results (SEMI Europe, Berlin, Germany, 2015), 6th ed. Revision 1, July 2015.

2. J. H. Petermann, D. Zielke, J. Schmidt, F. Haase, E. G. Rojas, and R. Brendel, “19%-efficient and 43 µm-thick crystalline si solar cell from layer transfer using porous silicon,” Prog. Photovolt: Res. Appl. 20, 1–5 (2012). [CrossRef]  

3. F. Dross, K. Baert, T. Bearda, J. Deckers, V. Depauw, O. El Daif, I. Gordon, A. Gougam, J. Govaerts, S. Granata, R. Labie, X. Loozen, R. Martini, A. Masolin, B. O’Sullivan, Y. Qiu, J. Vaes, D. Van Gestel, J. Van Hoeymissen, A. Vanleenhove, K. Van Nieuwenhuysen, S. Venkatachalam, M. Meuris, and J. Poortmans, “Crystalline thin-foil silicon solar cells: where crystalline quality meets thin-film processing,” Prog. Photovolt: Res. Appl. 20, 770–784 (2012). [CrossRef]  

4. J. Govaerts, C. Trompoukis, H. S. Radhakrishnan, L. Tous, S. N. Granata, E. G. Carnemolla, R. Martini, A. Marchegiani, M. Karim, I. Sharlandziev, T. Bearda, V. Depauw, K. V. Nieuwenhuysen, I. Gordon, J. Szlufczik, and J. Poortmans, “Solar cells from epitaxial foils: An epifoil epiphany,” Energy Procedia 77, 871–880 (2015). [CrossRef]  

5. S. Jeong, M. D. McGehee, and Y. Cui, “All-back-contact ultra-thin siliconnanocone solar cells with 13.7% power conversion efficiency,” Nat. Commun. 4, 2950 (2013). [CrossRef]  

6. A. Ingenito, O. Isabella, and M. Zeman, “Experimental demonstration of 4n2 classical absorption limit in nan-otextured ultrathin solar cells with dielectric omnidirectional back reflector,” ACS Photonics 1, 270–278 (2014). [CrossRef]  

7. M. S. Branham, W.-C. Hsu, S. Yerci, J. Loomis, S. V. Boriskina, B. R. Hoard, S. E. Han, and G. Chen, “15.7% efficient 10-µm-thick crystalline silicon solar cells using periodic nanostructures,” Adv. Mater. 27, 2182–2188 (2015). [CrossRef]   [PubMed]  

8. C. Becker, D. Amkreutz, T. Sontheimer, V. Preidel, D. Lockau, J. Haschke, L. Jogschies, C. Klimm, J. Merkel, P. Plocica, S. Steffens, and B. Rech, “Polycrystalline silicon thin-film solar cells: status and perspectives,” Sol. Energ. Mat. Sol. C. 119, 112–123 (2013). [CrossRef]  

9. D. Amkreutz, J. Haschke, T. Häring, F. Ruske, and B. Rech, “Conversion efficiency and process stability improvement of electron beam crystallized thin film silicon solar cells on glass,” Sol. Energ. Mat. Sol. C. 123, 13–16 (2014). [CrossRef]  

10. J. Haschke, D. Amkreutz, L. Korte, F. Ruske, and B. Rech, “Towards wafer quality crystalline silicon thin-film solar cells on glass,” Sol. Energ. Mat. Sol. C. 128, 190–197 (2014). [CrossRef]  

11. C. Becker, V. Preidel, D. Amkreutz, J. Haschke, and B. Rech, “Double-side textured liquid phase crystallized silicon thin-film solar cells on imprinted glass,” Sol. Energ. Mat. Sol. C. 135, 2–7 (2015). [CrossRef]  

12. T. Frijnts, S. Kühnapfel, S. Ring, O. Gabriel, S. Calnan, J. Haschke, B. Stannowski, B. Rech, and R. Schlatmann, “Analysis of photo-current potentials and losses in thin film crystalline silicon solar cells,” Sol. Energ. Mat. Sol. C. 143, 457–466 (2015). [CrossRef]  

13. C. Becker, P. Wyss, D. Eisenhauer, J. Probst, V. Preidel, M. Hammerschmidt, and S. Burger, “5 × 5 cm2 silicon photonic crystal slabs on glass and plastic foil exhibiting broadband absorption and high-intensity near-fields,” Sci. Rep. 4, 5886 (2014). [CrossRef]  

14. A. J. Wolf, H. Hauser, V. Kübler, C. Walk, O. Hohn, and B. Bläsi, “Origination of nano- and microstructures on large areas by interference lithography,” Microelectron. Eng. 98, 293–296 (2012). [CrossRef]  

15. A. Mellor, I. Tobías, A. Martí, and A. Luque, “A numerical study of bi-periodic binary diffraction gratings for solar cell applications,” Sol. Energ. Mat. Sol. C. 95, 3527–3535 (2011). [CrossRef]  

16. A. Mellor, H. Hauser, C. Wellens, J. Benick, J. Eisenlohr, M. Peters, A. Guttowski, I. Tobías, A. Martí, A. Luque, and B. Bläsi, “Nanoimprinted diffraction gratings for crystalline silicon solar cells: implementation, characteri-zation and simulation,” Opt. Express 21, A295–A304 (2013). [CrossRef]  

17. R. Peretti, G. Gomard, C. Seassal, X. Letartre, and E. Drouard, “Modal approach for tailoring the absorption in a photonic crystal membrane,” J. Appl. Phys. 111, 123114 (2012). [CrossRef]  

18. C. Trompoukis, I. Abdo, R. Cariou, I. Cosme, W. Chen, O. Deparis, A. Dmitriev, E. Drouard, M. Foldyna, E. G. Caurel, I. Gordon, B. Heidari, A. Herman, L. Lalouat, K.-D. Lee, J. Liu, K. Lodewijks, F. Mandorlo, I. Massiot, A. Mayer, V. Mijkovic, J. Muller, R. Orobtchouk, G. Poulain, P. Prod’Homme, P. R. i. Cabarrocas, C. Seassal, J. Poortmans, R. Mertens, O. E. Daif, and V. Depauw, “Photonic nanostructures for advanced light trapping in thin crystalline silicon solar cells,” Phys. Status Solidi A 212, 140–155 (2015). [CrossRef]  

19. O. Isabella, S. Solntsev, D. Caratelli, and M. Zeman, “3-D optical modeling of thin-film silicon solar cells on diffraction gratings,” Prog. Photovolt: Res. Appl. 21, 94–108 (2013). [CrossRef]  

20. H. Sai, T. Matsui, K. Saito, M. Kondo, and I. Yoshida, “Photocurrent enhancement in thin-film silicon solar cells by combination of anti-reflective sub-wavelength structures and light-trapping textures,” Prog. Photovolt: Res. Appl. 23, 1572–1580 (2015). [CrossRef]  

21. V. Preidel, D. Amkreutz, J. Haschke, M. Wollgarten, B. Rech, and C. Becker, “Balance of optical, structural, and electrical properties of textured liquid phase crystallized Si solar cells,” J. Appl. Phys. 117, 225306 (2015). [CrossRef]  

22. J. Pomplun, S. Burger, L. Zschiedrich, and F. Schmidt, “Adaptive finite element method for simulation of optical nano structures,” Phys. Status Solidi B 244, 3419–3434 (2007). [CrossRef]  

23. IEC, “IEC:60904-3: Photovoltaic devices - Part 3: Measurement principles for terrestrial photovoltaic (PV) solar devices with reference spectral irradiance data,” (2008).

24. D. Lockau, “Optical modeling of thin film silicon solar cells with random and periodic light management textures,” Ph.D. thesis, Technische Universität Berlin (2013).

25. A. Naqavi, F.-J. Haug, K. Söderström, C. Battaglia, V. Paeder, T. Scharf, H. P. Herzig, and C. Ballif, “Angular behavior of the absorption limit in thin film silicon solar cells,” Prog. Photovolt: Res. Appl. 22, 1147–1158 (2014). [CrossRef]  

26. X. Li, P.-C. Li, L. Ji, C. Stender, C. McPheeters, S. R. Tatavarti, K. Sablon, and E. T. Yu, “Subwavelength nanostructures integrated with polymer-packaged iii–v solar cells for omnidirectional, broad-spectrum improvement of photovoltaic performance,” Prog. Photovolt: Res. Appl. 23, 1398–1405 (2015). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Illustrating different sinusoidal gratings in a 3D (top) and a contour plot representation (bottom): (a) a square lattice constructed according to Eq. (1). Three hexagonal lattices constructed according to Eq. (3) for different structure phases ϕ: (b) a positive cosine lattice (“cos”) with ϕ = 0, (c) a sine lattice (“sin”) with ϕ = π/2, and (d) a negative cosine lattice (“−cos”) with ϕ = π. The red quadrilaterals enclose the unit cell.
Fig. 2
Fig. 2 The layer stack and the mesh used in this work shown for a positive cosine (“cos”) texture with P = 500 nm and h = 250 nm, (a) for the simple layer stack and (b) a layer stack with additional antireflective (AR) SiOx and SiNx layers.
Fig. 3
Fig. 3 Simulation results for the simple layer stack without AR layers [see Fig. 2(a)] for the four different geometries presented in Fig. 1 and a constant aspect ratio of a = h/P = 0.5. (a) The maximum achievable current density J ph max for five different periods and (b) the spectra of the in-coupled light for the three hexagonal structures and P = 500 nm are shown. A flat reference and, in (a), the current achieved with full absorption are shown as well. All results are corrected for the reflection loss at the air-glass interface (about 4%).
Fig. 4
Fig. 4 The imaginary part of the electric field energy density 1 4 ( E D * ), which is proportional to the absorbed power density. The simulation results are shown for the simple layer stack for the three hexagonal geometries with P = 500 nm and h = 250 nm at λ = 500 nm along cross sections through the unit cells, as they are depicted in Figs. 1(b)–1(d).
Fig. 5
Fig. 5 (a) The number of diffracted channels reflected back into the glass for a period of P = 500 nm at 330 nm, 380 nm, 650 nm and 1050 nm. To relate these results to the fraction coupled into Si (FCiSi) shown in Fig. 3(b), the curve for the negative cosine (“−cos”) is shown as well. (b) The reciprocal lattice of a hexagonal lattice with period P. The nearest, second-nearest and third-nearest neighbors of an arbitrary point are depicted, which correspond to the first, second and third diffraction orders, respectively.
Fig. 6
Fig. 6 The maximum achievable photocurrent densitiy J ph max multiplied with 1/cosθglass for oblique incidence in the simple layer stack for a negative cosine geometry (“−cos”) with P = 500 nm and h = 250 nm and for a flat layer stack.
Fig. 7
Fig. 7 (a) J ph max and (b) fraction of light coupled into Si for the simple layer stack and the complex layer stack with additional AR layers. The results are shown for the negative cosine geometry (P = 500 nm) for different aspect ratios a (or structure heights h).

Equations (13)

Equations on this page are rendered with MathJax. Learn more.

f s q ( x , y ) = cos x cos y ,
x 2 π P x , y 2 π P y .
f hex ( x , y ) = cos [ 1 2 ( x + 3 y ) ] cos [ 1 2 ( x 3 y ) ] cos ( x + ϕ ) ,
x 2 π 3 P x , y 2 π 3 P y .
min ( f hex , ϕ ) = 1 4 [ cos ϕ 3 sin ( ϕ 3 + π 6 ) ] ,
min ( f hex , ϕ ) = 1 4 [ cos ϕ + 3 cos ( ϕ 3 ) ] .
h min max ( ϕ ) : = max ( f hex , ϕ ) max ( f hex , ϕ ) = 3 3 4 sin ( ϕ 3 + π 3 ) .
J ph max = e λ 1 λ 2 A inc ( λ ) Φ AM 1.5 ( λ ) cos θ in d λ ,
A inc = 1 P in V ( Si ) S n d A ,
g 1 = 4 π 3 P ( 1 2 , 3 2 ) , g 2 = 4 π 3 P ( 1 2 , + 3 2 )
k 1 = | g 1 | = | g 2 | = 4 π 3 P g , k 2 = 3 g , k 3 = 2 g .
sin θ i = k i k = k i = λ 2 π n ,
λ i = 2 π n k i = .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.