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

On accurate simulations of thin-film solar cells with a thick glass superstrate

Open Access Open Access

Abstract

The optical response of periodically nanotextured layer stacks with dimensions comparable to the wavelength of the incident light can be computed with rigorous Maxwell solvers, such as the finite element method (FEM). Experimentally, such layer stacks are often prepared on glass superstrates with a thickness, which is orders of magnitude larger than the wavelength. For many applications, light in these thick superstrates can be treated incoherently. The front side of thick superstrate is located far away from the computational domain of the Maxwell solvers. Nonetheless, it has to be considered in order to achieve accurate results. In this contribution, we discuss how solutions of rigorous Maxwell solvers can be corrected for flat front sides of the superstrates with an incoherent a posteriori approach. We test these corrections for hexagonal sinusoidal nanotextured silica-silicon interfaces, which are applied in certain silicon thin-film solar cells. These corrections are determined via a scattering matrix, which contains the full scattering information of the periodically nanotextured structure. A comparison with experimental data reveals that higher-order corrections can predict the measured reflectivity of the samples much better than an often-applied zeroth-order correction.

© 2017 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Numerical Maxwell solvers are well suited to simulate layer stacks of periodically structured thin optical films, which for example are used in thin-film solar cells. Because the computational cost scales with the size of the computational domain, the maximal volume of unit cells used for these simulations is limited.

In practice, such thin-film stacks often are created on glass superstrates with a thickness in the order of millimetres. A superstrate is a substrate, through which the incident light reaches the thin-film stack. Such thick glass layers can always be treated incoherently, because sunlight has a spatial coherence length of about 27 µm [1–3] and a coherence time of about 3 fs [4, 5] at 500 nm wavelength, which corresponds to a disc of about 27 µm radius and 1 µm height.

In recent years, several approaches were discussed to treat optical systems with thick incoherent layers [6–9]. With the finite element method (FEM) [10], such thick layers can be taken into account directly for example with a domain decomposition approach [11]. However, then the layer is treated as a coherent layer, which leads to reflectance/absorptance spectra with a very narrow interference pattern. This contradicts experimental observations. If the sample is not strongly scattering, this interference can be eliminated by combining simulations with well-selected different glass thicknesses [12]. But if the sample scatters strongly, eliminating the interference in the different diffraction orders becomes very cumbersome if not impossible.

A common way to take the incoherent superstrate into account in FEM is to represent the glass layer by a thin layer of a few hundreds of nanometres thickness, which is covered by an infinite half space filled with the same material (glass), as illustrated in Fig. 1(a). The effect of the air-glass interface, which is far off the computational domain, then is applied to the simulation results a posteriori.

 figure: Fig. 1

Fig. 1 (a) Cross section through a periodic unit cell (enclosed by the green box) with a sinusoidal interface [14] and the glass and silicon half spaces above and below the unit cell. The a posteriori corrections described in this paper account for the interaction of light, which is reflected from the unit cell back into the glass half space, with the glass-air interface, depicted in red. The intensities are named as in Section 2: I0,a and I0 are the incident intensities in air and glass, respectively, and Rg is the reflectivity of the air-glass interface; all three variables must be treated for both polarizations.Ioutj is the vector of all intensities emitted from the unit cell in the jth correction order. The glass-air interface reflects the intensities Iinj back towards the unit cell. ρj is the contribution of the jth correction order to the total reflectivity. (b) The sinusoidal texture used in the simulations, which is generated according to Eq. (14). (c) Atomic force microscopy (AFM) measurements of a sinusoidally nanotextured sol-gel layer on glass with 750 nm pitch and about 150 nm texture height, hence an aspect ratio of about a = 20%.

Download Full Size | PDF

In the zeroth-order correction only the initial reflection of the incident light beam at the air-glass interface is taken into account [13, 14]. Interactions of light, which is reflected from the computational domain back into the glass half space, with the glass-air interface are neglected. Higher-order corrections at the glass-air interface may become important, if a significant fraction of the back-reflected light is directed into angles, which are comparable to or larger as the critical angle [15]. Recently we discussed a first-order correction, which takes the first interaction of the back-reflected light with the glass-air interface into account [16, 17]. The first-order correction agrees much better with experimental data than the zeroth-order correction.

In this manuscript we calculate the scattering matrix, which allows to evaluate not only the first order but an arbitrary number of correction orders at negligible computational cost. The number of the correction order denotes, how often light, which is reflected from the computational domain back into the glass superstrate, interacts with the glass-air interface. We discuss the difference between the zeroth and higher-order corrections in dependence of the wavelength. Further, we confirm the validity of this approach with experimental data. The findings presented in this contribution are important for improving the accuracy in optical simulations of nanostructured solar cells on thick glass superstrates.

2. Theory

In numerical simulations, periodically structured thin-film layer stacks are usually treated with periodic boundary conditions on the side faces of the unit cell. Because of the perfect periodicity, far-field reflection and transmission into the top and bottom infinite half spaces, which cover the layer stack, only happens into discrete and well-defined diffraction orders. The reflectivity R of the layer stack in the glass halfspace can be calculated with

Rp(λ)=1|Ei,pg(λ)|2cosθig(λ)n|En,pg(λ)|2cosθng(λ),
for p polarization and similar for s polarization, where the Ei,p and En,pg are replaced by Ei,s and En,sg, respectively. In Eq. (1) the electromagnetic field components Eng and the angles θng are output from a Fourier transform postprocess of the Maxwell solver. The subscript i denotes the incident wave, the superscript g denotes fields and angles in glass, and λ is the vacuum wavelength. The sum is taken over all directions into which the structure reflects.

The zeroth-order correction for reflection accounts only for the initial reflection of the air-glass interface. The reflectivity in air R0 is calculated using

Rp0(λ)=Rp(λ)[1Rpg(λ)]+Rpg(λ),
for p polarization and similar for s polarization. Here, the superscript 0 denotes the zeroth-order correction, Rpg and Rsg are the reflectivities of the air-glass interface for the two polarizations (about 4% at normal incidence), as illustrated in Fig. 1(a) [13, 14].

We obtain higher-order corrections for reflection with the scattering matrix S. Initially, light is incident onto the nanotexture with a k-vector ki with absolute value ki = 2πng(λ)/λ, where ng(λ) is the refractive index of glass. The hexagonal nanotexture reflects light into several diffractions orders with directions knout, where n ∈ {1, 2 … N}. As illustrated in [14], N depends on the wavelength, the period of the nanotexture, its shape (square or hexagonal), the direction of incidence, and the refractive index of the material into which the light is diffracted.

The scattering matrix connects all outgoing directions knout to the directions knin, from which light impinges onto the nanotexture after it is reflected from the glass-air interface back towards the computational domain. They are connected to each other via kn,out=kn,in and kn,out=kn,in, where ‖ denotes the component parallel to the lattice plane of the nanotexture (i.e. in the x y-plane) and ⊥ denotes the perpendicular component, i.e. in z-direction. Note that light incident onto the nanotexture from one of the incident directions kmin with m ∈ {1, 2 … N) leads to the reflection in all directions knout with n ∈ {1, 2 … N}. Because both p and s polarizations have to be taken into account, the dimension of the scattering matrix is 2N ×2N, S2N×2N. In other programs, such as OPTOS [9] or GenPro4 [18], the scattering matrices are defined for non-periodic structures. Their size corresponds to the angular discretization and can be used to steer the accuracy of the calculations.

The incident k-vectors can be calculated using the reciprocal grid, which we illustrated in [14]; its gridpoints Gpq are given by

Gpq=pg1+pg2,
where p, q ∈ ℤ are integers and g1 and g2 are the primitive vectors of the reciprocal grid. For a hexagonal grid with period (pitch) P they are given by
g1=4π3P(12,32),g2=4π3P(12,+32).
Now we find all the possible k-vectors via the condition that their parallel components, are shorter or of equal length than the k-vector of the incident light ki. The parallel components are equivalent to the vectors connecting the origin O of the reciprocal grid to the gridpoints Gpq plus the parallel component of the incident direction, ki. Hence, the condition reads
|kpq,in|=|GpqO+ki|ki.
The respective perpendicular components are given by
kpq,in=ki2|kpq,in|2

Since we are interested in the incoherent propagation of light in the glass superstrate, we have to look at the intensities. Hence, the elements of the scattering matrix are given by

Sba=|Ebag|2cosθbcosθa,
where Eba is the bth outgoing field originating from an incident field distribution where only the ath incident field is different from zero with strength 1. The indices a, b ∈ {1, 2 2N} denote the possible direction and polarization combinations. We further need two vectors that account for reflection and transmission at the glass-air interface: R ∈ ℝ2N and T ∈ ℝ2N. Their components are calculated using the Fresnel equations.

The jth-order correction for the reflection corresponds to the jth interaction of the light, which is reflected from the computational domain into the glass superstrate with the glass-air interface, as illustrated in Fig. 1(a). It is calculated as follows: let Iinj12N be the intensity distribution incident onto the nanotexture in the (j − 1)th-order correction for reflection. The intensities emitted from the nanotexture are given by

Ioutj=SIinj1.
The contribution of the jth-order to the reflection is given by
ρj=IoutjT,
where · denotes the scalar product of two vectors. The intensity distribution incident onto the nanotexture in the jth order Iinj is given by component-wise multiplication of Ioutj with R,
Iin,aj=Iout,ajRa.

We start the iterative process with Iin0, where only the two components corresponding to the ki direction are different from zero. These two components are the incident intensities in p- and s-polarization. They are Ip0=0.5[1Rpg(λ)] and Is0=0.5[1Rsg(λ)], where interface. Rpg and Rsg are the initial reflectivites of the air-glass interface.

The reflectivity of the structure in the th-order correction is given by

R(λ)=j=1ρj(λ)+12[Rpg(λ)+Rsg(λ)].
In this manuscript, we sum iteratively until convergence is reached, i.e. when
ρR1<τ,
with a convergence threshold τ = 10−4.

Note that the zeroth-order correction R0—defined in Eq. (2)—cannot be derived from Eq. (11). R1 with = 1 is equivalent to the first-order correction, which also can be calculated directly without using the scattering matrix [16, 17],

R1=1|Eia|2cosθian[(tnsEs,ng)2+(tnpEp,ng)2]cosθna+Rg.
Here, the Eng vectors are decomposed into s- and p-polarized components before multiplying them with the Fresnel coefficients tns and tnp, respectively, which describe transmission from glass into air. Note that Eia is connected to Eig, which is used in Eq. (1), via transmission from air into glass. For incoherent sunlight, Equation (13) has to be applied to incident s- and p-polarized light, independently.

3. Simulation and experimental details

To test the approach described above, we used the sinusoidal nanotexture denoted as negative cosine texture in earlier work [14, 19], which is illustrated in Fig. 1(b). Mathematically, it is—up to vertical and horizontal scaling—described with

f(x,y)=cos(x)cos[12(x+3y)]cos[12(x3y)].

The simulations required to determine the scattering matrix were performed with the Maxwell solver JCMsuite, which utilizes the finite element method (FEM) [20]. The optical system studied for the simulations in this work is sketched in Fig. 1(a). The optimal dimensions of the tetrahedra and prisms constituting the three-dimensional mesh for the FEM simulations were found to be maximally 50 nm for the sinusoidal interface and 100 nm for the volume elements, as discussed earlier [14]. One advantage of JCMsuite is that all 2N incident electric fields (N directions and two polarizations) can be treated in one simulation, which makes these calculations very efficient.

Experimentally, Köppel and co-workers successfully demonstrated the use of sinusoidal nanotextures as antireflective scheme for thin-film solar cells with a liquid-phase crystallized silicon (LPC-Si) absorber [21, 22]. The nanotextures are fabricated with nanoimprint lithography [8, 23] using high-temperature stable sol-gels [24], which are suited for the LPC process; an atomic force microscopy (AFM) picture of such a texture is shown in Fig. 1(c). Onto the nanotexture an about 10 µm thick nanocrystalline silicon layer is deposited, which subsequently is molten and re-crystallized with a continuous-wave (cw) line laser [25, 26].

To calculate the aspect ratio we divided the texture height, derived from AFM measurements, by the pitch; the error of the aspect ratio was deduced from the AFM error of ±20 nm. The reflectance spectra were obtained with an integrating sphere of 150 nm diameter, which is attached to a PerkinElmer LAMBDA 1050 spectrophotometer, at θin = 8° angle of incidence.

The experimental sample has an LPC-Si layer thickness of about 10 µm, while in the simulations the silicon layer is considered to be infinitely thick. For wavelength shorter than about 600 nm, all light that reaches the 10 µm-thick silicon layer will be absorbed. Hence, we compare simulated and measured reflectance spectra in a 350–600 nm range.

4. Results

Figure 2 shows the angles of the different diffraction orders in glass and air for light, which is diffracted by hexagonal periodic structures with 500 nm and 750 nm pitch at normal incidence. The zeroth diffraction order (θg = θa ≡ 0) is not shown. In glass, the different diffraction orders are present until much longer wavelength than in air. The zeroth order correction [Eqs. (1) and (2)] takes all diffraction orders in glass into account, even if they cannot propagate in air, leading to an overestimated reflectivity.

 figure: Fig. 2

Fig. 2 Angles of the diffraction orders into which a hexagonal periodic structure with a pitch of (a) P = 500 nm and (b) P = 750 nm scatters light at normal incidence. The figure shows the angles in glass and in air. The zeroth diffraction order (θg = θa ≡ 0) is not shown. Experimental data is available in the wavelength range (350–600 nm) – this range is marked by the white boxes. In glass, the diffraction orders are present up to much longer wavelength than in air, which is also illustrated in (c). For P = 750 nm, also the 4th and 5th diffraction orders are present in glass at short wavelengths, but in (b) they are omitted for clarity.

Download Full Size | PDF

Figure 3(a) shows numerical and experimental 1 − R spectra for a nanotextured sample with 500 nm pitch. Large differences are seen between the 0th-order correction [Eq. (2)] and the other two curves. The converged data was obtained iteratively with the scattering matrix approach until the threshold of τ = 10−4 was reached [Eq. (12)]. It is a bit lower than the 1st-order correction [Eq. (13)], which constitutes an upper bound to 1 − R. Both the 1st-order correction and the converged curve agree well with the measurement, in contrast to the 0th-order correction. Also for a pitch of 750 nm, shown in Fig. 3(b), the 0th-order correction shows much worse agreement with the experimental data than the other two numerical curves.

 figure: Fig. 3

Fig. 3 Numerical and experimental 1 − R spectra for the nanotextured layer stacks illustrated in Fig. 1(a) with (a) 500 nm pitch and (b) 750 nm pitch. Numerical results were calculated with the 0th-order correction [Eq. (2)], the 1st-order correction [Eq. (13)], and iterated using the scattering matrix, until convergence was reached for the threshold τ = 10−4 [Eq. (12)]. The 0th-order correction differs strongly from the other curves because not all diffraction orders that are present in glass can propagate into air [see Fig. 2]. Simulation and experimental results are shown for θin = 8°.

Download Full Size | PDF

We take a closer look into the convergence of the scalar scattering matrix approach in Fig. 4, which shows the correction order required to reach convergence with τ = 10−4 for a nanotexture with 750 nm pitch and 25% aspect ratio. Further, the figure shows the number of directions N, into which the nanotexture diffracts light in glass. We observe that always increases towards the edge where the number of directions N decreases. This can be understood as follows: towards such an edge, the angles of the outermost scattered diffraction order become larger and larger, as shown on the right hand side of Fig. 4. Hence, the reflectivites at the glass-silicon and glass-air interfaces increase as well, and less light can leave the glass per correction order – either into air or into silicon, and a larger number of correction orders is required to reach convergence.

 figure: Fig. 4

Fig. 4 The correction order required to reach convergence with τ = 10−4 [red line, see Eq. (12)] for a nanotexture with 750 nm pitch and 25% aspect ratio at normal incidence. Further, the number of directions N is shown, into which the nanotexture diffracts light in glass [cyan dots]. For three wavelengths, these directions are shown on the right. The numbers correspond to the polar angles. More information is given in [14].

Download Full Size | PDF

Figure 5 shows the mean 1 − R for numerical and experimental data as a function of the aspect ratio a of the nanotexture and two different pitches. The mean is taken between 350 and 600 nm wavelength. For both pitches, the first-order correction and the converged curve resemble the experimental data much better than the zeroth-order correction. Even for planar structures (a = 0), R0 and R1 differ by several percent. Figure 5 shows that 1 − R stabilizes for aspect ratios a > 0.4. At 500 nm pitch, 1 − R reaches almost 0.92 while for P = 750 nm, only 0.90 are reached. However, aspect ratios exceeding ≈ 0.3 are difficult to realize experimentally [21].

 figure: Fig. 5

Fig. 5 Simulated and experimental mean 1 − R between 350 and 600 nm wavelength for a nanotexture pitch of (a) P = 500 nm and (b) 750 nm. All results are shown for θin = 8°.

Download Full Size | PDF

5. Summary and outlook

When simulating the optics of periodically nanotextured layer stacks on thick glass superstrates it is imperative to take the front interface into account adequately. Often, one only corrects for initial reflection losses at the air-glass interface, which are about 30% at normal incidence. We show that it is also very important to take higher-order reflections at the glass-air interface into account, which occur for light that is reflected from the nanotextured layer stack back into glass. The incoherent scattering matrix allows to evaluate an arbitrary number of reflection orders at negligible computational cost. However, for a rough estimate also the first-order correction might be sufficient, as recently shown [17]. The first-order correction is easier to evaluate because it can be calculated without the scattering matrix.

In this manuscript we only discuss the correction for glass superstrates with a flat air-glass interface. The model can be easily expanded to superstrates with periodically textured glass-air interfaces—if their period is the same as that of sinusoidal nanotexture. Then, the vectors R and T, as defined in Section 2 have to be replaced by a scattering matrix for the glass-air interface. If that interface has a different period or is randomly textured, both nanotextured interfaces would need to be described with scattering matrices for non-periodic nanotextures [9, 18].

Currently, the method described in this manuscript can be used to correct reflection spectra. However, it cannot correct the absorptivity in thick absorbing layers, such as silicon, yet. Scattering layer stacks, which are represented as scattering matrices, can be coupled to incoherent layers in matrix-based solvers [9, 18]. In a next step, we will investigate combinations between our FEM solver and such approaches in order to calculate full absorption profiles of layer stacks consisting of thick incoherent layers and thin periodically nanotextured systems.

Funding

German Federal Ministry of Education and Research (BMBF) via Nachwuchswettbewerb NanoMatFutur (No. 03X5520); Einstein Foundation Berlin (ECMath project SE6).

Acknowledgments

The results were obtained at the Berlin Joint Lab for Optical Simulations for Energy Research (BerOSE) of Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, Zuse Institute Berlin and Freie Universität Berlin.

References and links

1. M. E. Verdet, Leçons d’Optique Physique (L’Imprimerie Impériale, 1869).

2. G. S. Agarwal, G. Gbur, and E. Wolf, “Coherence properties of sunlight,” Opt. Lett. 29, 459–461 (2004). [CrossRef]   [PubMed]  

3. S. Divitt and L. Novotny, “Spatial coherence of sunlight and its implications for light management in photovoltaics,” Optica 2, 95–103 (2015). [CrossRef]  

4. A. Herman, M. Sarrazin, and O. Deparis, “The fundamental problem of treating light incoherence in photovoltaics and its practical consequences,” New J. Phys. 16, 013022 (2014). [CrossRef]  

5. E. Hecht, Optics (Pearson Higher Education, Harlow, England, 2016), 5th ed.

6. M. C. Troparevsky, A. S. Sabau, A. R. Lupini, and Z. Zhang, “Transfer-matrix formalism for the calculation of optical response in multilayer systems: from coherent to incoherent interference,” Opt. Express 18, 24715–24721 (2010). [CrossRef]   [PubMed]  

7. R. Santbergen, A. H. M. Smets, and M. Zeman, “Optical model for multilayer structures with coherent, partly coherent and incoherent layers,” Opt. Express 21, A262–A267 (2013). [CrossRef]   [PubMed]  

8. 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, characterization and simulation,” Opt. Express 21, A295–A304 (2013). [CrossRef]   [PubMed]  

9. N. Tucher, J. Eisenlohr, H. Gebrewold, P. Kiefel, O. Höhn, H. Hauser, J. C. Goldschmidt, and B. Bläsi, “Optical simulation of photovoltaic modules with multiple textured interfaces using the matrix-based formalism optos,” Opt. Express 24, A1083–A1093 (2016). [CrossRef]   [PubMed]  

10. J. Jin, The Finite Element Method in Electromagnetics (John Wiley & Sons, 2002).

11. A. Schädle, L. Zschiedrich, S. Burger, R. Klose, and F. Schmidt, “Domain decomposition method for Maxwell’s equations: Scattering off periodic structures,” J. Comput. Phys. 226, 477–493 (2007). [CrossRef]  

12. A. Campa, J. Krc, and M. Topic, “Two approaches for incoherent propagation of light in rigorous numerical simulations,” Progress In Electromagnetics Research 137, 187–202 (2013). [CrossRef]  

13. 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]  

14. K. Jäger, C. Barth, M. Hammerschmidt, S. Herrmann, S. Burger, F. Schmidt, and C. Becker, “Simulations of sinusoidal nanotextures for coupling light into c-Si thin-film solar cells,” Opt. Express 24, A569–A580 (2016). [CrossRef]   [PubMed]  

15. D. Lockau, L. Zschiedrich, S. Burger, F. Schmidt, F. Ruske, and B. Rech, “Rigorous optical simulation of light management in crystalline silicon thin film solar cells with rough interface textures,” Proc. SPIE 7933, 79330M (2011). [CrossRef]  

16. K. Jäger, M. Hammerschmidt, G. Köppel, S. Burger, and C. Becker, “On accurate simulations of thin-film solar cells with a thick glass superstrate,” in “Light, Energy and the Environment,” (Optical Society of America, 2016), p. PM3B.5. [CrossRef]  

17. K. Jäger, G. Köppel, D. Eisenhauer, D. Chen, M. Hammerschmidt, S. Burger, and C. Becker, “Optical simulations of advanced light management for liquid-phase crystallized silicon thin-film solar cells,” Proc. SPIE 10356, 103560F (2017).

18. R. Santbergen, T. Meguro, T. Suezaki, G. Koizumi, K. Yamamoto, and M. Zeman, “GenPro4 Optical Model for Solar Cell Simulation and Its Application to Multijunction Solar Cells,” IEEE J. Photovolt. 7, 919–926 (2017). [CrossRef]  

19. K. Jäger, G. Köppel, C. Barth, M. Hammerschmidt, S. Herrmann, S. Burger, F. Schmidt, and C. Becker, “Sinusoidal gratings for optimized light management in c-Si thin-film solar cells,” Proc. SPIE 9898, 989808 (2016). [CrossRef]  

20. 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]  

21. G. Köppel, B. Rech, and C. Becker, “Sinusoidal nanotextures for light management in silicon thin-film solar cells,” Nanoscale 8, 8722–8728 (2016). [CrossRef]   [PubMed]  

22. G. Köppel, D. Amkreutz, P. Sonntag, G. Yang, R. V. Swaaij, O. Isabella, M. Zeman, B. Rech, and C. Becker, “Periodic and Random Substrate Textures for Liquid-Phase Crystallized Silicon Thin-Film Solar Cells,” IEEE J. Photovolt. 7, 85–90 (2017). [CrossRef]  

23. C. Battaglia, J. Escarré, K. Söderström, M. Charrière, M. Despeisse, F. Haug, and C. Ballif, “Nanomoulding of transparent zinc oxide electrodes for efficient light trapping in solar cells,” Nat. Photonics 5, 535 (2011). [CrossRef]  

24. M. Verschuuren and H. van Sprang, “3D Photonic Structures by Sol-Gel Imprint Lithography,” Mater. Res. Soc. Symp. Proc. 1002, 1002 (2007). [CrossRef]  

25. J. Haschke, D. Amkreutz, and B. Rech, “Liquid phase crystallized silicon on glass: Technology, material quality and back contacted heterojunction solar cells, ” Jpn. J. Appl. Phys. 55, 04EA04 (2016). [CrossRef]  

26. C. T. Trinh, N. Preissler, P. Sonntag, M. Muske, K. Jäger, M. Trahms, R. Schlatmann, B. Rech, and D. Amkreutz, “Potential of interdigitated back-contact silicon heterojunction solar cells for liquid phase crystallized silicon on glass with efficiency above 14%,” Sol. Energ. Mat. Sol. C. 174, 187–195 (2018). [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 (5)

Fig. 1
Fig. 1 (a) Cross section through a periodic unit cell (enclosed by the green box) with a sinusoidal interface [14] and the glass and silicon half spaces above and below the unit cell. The a posteriori corrections described in this paper account for the interaction of light, which is reflected from the unit cell back into the glass half space, with the glass-air interface, depicted in red. The intensities are named as in Section 2: I0, a and I0 are the incident intensities in air and glass, respectively, and Rg is the reflectivity of the air-glass interface; all three variables must be treated for both polarizations. I out j is the vector of all intensities emitted from the unit cell in the jth correction order. The glass-air interface reflects the intensities I in j back towards the unit cell. ρj is the contribution of the jth correction order to the total reflectivity. (b) The sinusoidal texture used in the simulations, which is generated according to Eq. (14). (c) Atomic force microscopy (AFM) measurements of a sinusoidally nanotextured sol-gel layer on glass with 750 nm pitch and about 150 nm texture height, hence an aspect ratio of about a = 20%.
Fig. 2
Fig. 2 Angles of the diffraction orders into which a hexagonal periodic structure with a pitch of (a) P = 500 nm and (b) P = 750 nm scatters light at normal incidence. The figure shows the angles in glass and in air. The zeroth diffraction order (θg = θa ≡ 0) is not shown. Experimental data is available in the wavelength range (350–600 nm) – this range is marked by the white boxes. In glass, the diffraction orders are present up to much longer wavelength than in air, which is also illustrated in (c). For P = 750 nm, also the 4th and 5th diffraction orders are present in glass at short wavelengths, but in (b) they are omitted for clarity.
Fig. 3
Fig. 3 Numerical and experimental 1 − R spectra for the nanotextured layer stacks illustrated in Fig. 1(a) with (a) 500 nm pitch and (b) 750 nm pitch. Numerical results were calculated with the 0th-order correction [Eq. (2)], the 1st-order correction [Eq. (13)], and iterated using the scattering matrix, until convergence was reached for the threshold τ = 10−4 [Eq. (12)]. The 0th-order correction differs strongly from the other curves because not all diffraction orders that are present in glass can propagate into air [see Fig. 2]. Simulation and experimental results are shown for θin = 8°.
Fig. 4
Fig. 4 The correction order required to reach convergence with τ = 10−4 [red line, see Eq. (12)] for a nanotexture with 750 nm pitch and 25% aspect ratio at normal incidence. Further, the number of directions N is shown, into which the nanotexture diffracts light in glass [cyan dots]. For three wavelengths, these directions are shown on the right. The numbers correspond to the polar angles. More information is given in [14].
Fig. 5
Fig. 5 Simulated and experimental mean 1 − R between 350 and 600 nm wavelength for a nanotexture pitch of (a) P = 500 nm and (b) 750 nm. All results are shown for θin = 8°.

Equations (14)

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

R p ( λ ) = 1 | E i , p g ( λ ) | 2 cos θ i g ( λ ) n | E n , p g ( λ ) | 2 cos θ n g ( λ ) ,
R p 0 ( λ ) = R p ( λ ) [ 1 R p g ( λ ) ] + R p g ( λ ) ,
G p q = p g 1 + p g 2 ,
g 1 = 4 π 3 P ( 1 2 , 3 2 ) , g 2 = 4 π 3 P ( 1 2 , + 3 2 ) .
| k p q , in | = | G p q O + k i | k i .
k p q , in = k i 2 | k p q , in | 2
S b a = | E b a g | 2 cos θ b cos θ a ,
I out j = S I in j 1 .
ρ j = I out j T ,
I in , a j = I out , a j R a .
R ( λ ) = j = 1 ρ j ( λ ) + 1 2 [ R p g ( λ ) + R s g ( λ ) ] .
ρ R 1 < τ ,
R 1 = 1 | E i a | 2 cos θ i a n [ ( t n s E s , n g ) 2 + ( t n p E p , n g ) 2 ] cos θ n a + R g .
f ( x , y ) = cos ( x ) cos [ 1 2 ( x + 3 y ) ] cos [ 1 2 ( x 3 y ) ] .
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.