The tradeoff between plasmonic enhancement and ... - OSA Publishing

1 downloads 144 Views 2MB Size Report
A. Campa, J. Krč, and M. Topič, “Analysis and optimisation of microcrystalline .... K. Seo, M. Wober, P. Steinvurzel, E. Schonbrun, Y. Dan, T. Ellenbogen, and ...
The tradeoff between plasmonic enhancement and optical loss in silicon nanowire solar cells integrated in a metal back reflector Keya Zhou,1,2,4 Zhongyi Guo,3 Xiaopeng Li,1 Jin-Young Jung,1 Sang-Won Jee,1 KwangTae Park,1 Han-Don Um,1 Ning Wang,1 Jung-Ho Lee1,* 1

Department of Chemical Engineering, Hanyang University, Ansan, Gyeonggi, 426-791, Korea 2 Department of Physics, Harbin Institute of Technology, Harbin 150001, China 3 School of Computer and Information, Hefei University of Technology, Hefei 230009, China 4 [email protected] * [email protected]

Abstract: We perform a systematic numerical study to characterize the tradeoff between the plasmonic enhancement and optical loss in periodically aligned, silicon nanowire (SiNW) arrays integrated with a silver back reflector (Ag BR). Optimizing the embedded depth of the wire bottoms into a silver reflector achieved a highly efficient SiNW solar cell. Compared to the SiNW solar cell employing a flat back reflector, the embedded depth of ~20 nm resulted in the relative increase of ~5% in ultimate solar cell efficiency. ©2012 Optical Society of America OCIS codes: (310.6628) Subwavelength structures, nanostructures; (350.6050) Solar energy.

References and links 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17.

M. A. Green and S. Pillai, “Harnessing plasmonics for solar cells,” Nat. Photonics 6(3), 130–132 (2012). V. E. Ferry, L. A. Sweatlock, D. Pacifici, and H. A. Atwater, “Plasmonic Nanostructure Design for Efficient Light Coupling into Solar Cells,” Nano Lett. 8(12), 4391–4397 (2008). C. Lin and M. L. Povinelli, “The effect of plasmonic particles on solar absorption in vertically aligned silicon nanowire arrays,” Appl. Phys. Lett. 97(7), 071110 (2010). K. R. Catchpole and A. Polman, “Plasmonic solar cells,” Opt. Express 16(26), 21793–21800 (2008). W. Wang, S. Wu, K. Reinhardt, Y. Lu, and S. Chen, “Broadband Light Absorption Enhancement in Thin-Film Silicon Solar Cells,” Nano Lett. 10(6), 2012–2018 (2010). A. Campa, J. Krč, and M. Topič, “Analysis and optimisation of microcrystalline silicon solar cells with periodic sinusoidal textured interfaces by two-dimensional optical simulations,” J. Appl. Phys. 105(8), 083107 (2009). L. Zeng, P. Bermel, Y. Yi, B. Alamariu, K. A. Broderick, J. Liu, C. Hong, X. Duan, J. Joannopoulos, and L. C. Kimerling, “Demonstration of enhanced absorption in thin film Si solar cells with textured photonic crystal back reflector,” Appl. Phys. Lett. 93(22), 221105 (2008). R. Biswas and D. Zhou, “Simulation and modelling of photonic and plasmonic crystal back reflectors for efficient light trapping,” Phys. Status Solidi., A Appl. Mater. Sci. 207(3), 667–670 (2010). J. Springer, A. Poruba, L. Müllerova, M. Vanecek, O. Kluth, and B. Rech, “Absorption loss at nanorough silver back reflector of thin-film silicon solar cells,” J. Appl. Phys. 95(3), 1427–1429 (2004). F. J. Haug, T. Soderstrom, O. Cubero, V. Terrazzoni-Daudrix, and C. Ballif, “Plasmonic absorption in textured silver back reflectors of thin film solar cells,” J. Appl. Phys. 104(6), 064509 (2008). U. W. Paetzold, F. Hallermann, B. E. Pieters, U. Rau, R. Carius, and G. von Plessen, “Localized plasmonic losses at metal back contacts of thin-film silicon solar cells,” Proc. SPIE 7725, 772517, 772517-9 (2010). M. Peters, M. Rüdiger, H. Hauser, M. Hermle, and B. Bläsi, “Diffractive gratings for crystalline silicon solar cells—optimum parameters and loss mechanisms,” Prog. Photovolt. Res. Appl. n/a (2011), doi:10.1002/pip.1151. B. M. Kayes, H. A. Atwater, and N. S. Lewis, “Comparison of the device physics principles of planar and radial p-n junction nanorod solar cells,” J. Appl. Phys. 97(11), 114302 (2005). E. C. Garnett and P. Yang, “Silicon nanowire radial p-n junction solar cells,” J. Am. Chem. Soc. 130(29), 9224– 9225 (2008). E. Garnett and P. Yang, “Light Trapping in Silicon Nanowire Solar Cells,” Nano Lett. 10(3), 1082–1087 (2010). C. Lin and M. L. Povinelli, “Optical absorption enhancement in silicon nanowire arrays with a large lattice constant for photovoltaic applications,” Opt. Express 17(22), 19371–19381 (2009). L. Hu and G. Chen, “Analysis of Optical Absorption in Silicon Nanowire Arrays for Photovoltaic Applications,” Nano Lett. 7(11), 3249–3252 (2007).

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A777

18. M. D. Kelzenberg, S. W. Boettcher, J. A. Petykiewicz, D. B. Turner-Evans, M. C. Putnam, E. L. Warren, J. M. Spurgeon, R. M. Briggs, N. S. Lewis, and H. A. Atwater, “Enhanced absorption and carrier collection in Si wire arrays for photovoltaic applications,” Nat. Mater. 9(3), 239–244 (2010). 19. K.-T. Park, Z. Guo, H.-D. Um, J.-Y. Jung, J. M. Yang, S. K. Lim, Y. S. Kim, and J.-H. Lee, “Optical properties of Si microwires combined with nanoneedles for flexible thin film photovoltaics,” Opt. Express 19(S1 Suppl 1), A41–A50 (2011). 20. Y. Wu, Y. Cui, L. Huynh, C. J. Barrelet, D. C. Bell, and C. M. Lieber, “Controlled Growth and Structures of Molecular-Scale Silicon Nanowires,” Nano Lett. 4(3), 433–436 (2004). 21. Z. Zhang, T. Shimizu, L. Chen, S. Senz, and U. Gösele, “Bottom-Imprint Method for VSS Growth of Epitaxial Silicon Nanowire Arrays with an Aluminium Catalyst,” Adv. Mater. (Deerfield Beach Fla.) 21, 4701–4705 (2009). 22. J. M. Weisse, D. R. Kim, C. H. Lee, and X. Zheng, “Vertical Transfer of Uniform Silicon Nanowire Arrays via Crack Formation,” Nano Lett. 11(3), 1300–1305 (2011). 23. J. Kupec and B. Witzigmann, “Computational electromagnetics for nanowire solar cells,” J. Comput. Electron. 11(2), 153–165 (2012). 24. E. D. Palik, Handbook of Optical Constants of Solids (Academic, Orlando, Fla., 1985). 25. A. W. Snyder and J. D. Love, Optical Waveguide Theory (Chapman & Hall, New York, 1983). 26. K. Seo, M. Wober, P. Steinvurzel, E. Schonbrun, Y. Dan, T. Ellenbogen, and K. B. Crozier, “Multicolored Vertical Silicon Nanowires,” Nano Lett. 11(4), 1851–1856 (2011). 27. B. C. P. Sturmberg, K. B. Dossou, L. C. Botten, A. A. Asatryan, C. G. Poulton, C. M. de Sterke, and R. C. McPhedran, “Modal analysis of enhanced absorption in silicon nanowire arrays,” Opt. Express 19(S5 Suppl 5), A1067–A1081 (2011). 28. J. Kupec and B. Witzigmann, “Dispersion, wave propagation and efficiency analysis of nanowire solar cells,” Opt. Express 17(12), 10399–10410 (2009). 29. C. Genet and T. W. Ebbesen, “Light in tiny holes,” Nature 445(7123), 39–46 (2007). 30. J. Bhattacharya, N. Chakravarty, S. Pattnaik, W. D. Slafer, R. Biswas, and V. L. Dalal, “A photonic-plasmonic structure for enhancing light absorption in thin film solar cells,” Appl. Phys. Lett. 99(13), 131114 (2011). 31. C. Genet, M. P. van Exter, and J. P. Woerdman, “Fano-type interpretation of red shifts and red tails in hole array transmission spectra,” Opt. Commun. 225(4-6), 331–336 (2003).

1. Introduction Plasmonic effects have recently gained much interest in solar cell research and have presented new opportunities for more efficient solar cell designs [1]. Experimental results showing that the photogenerated current can be improved by introducing plasmonic nanostructures into photovoltaic devices have been reported [2–4]. Plasmonic back contacts (called back reflectors, or BR) that employ nanogratings and/or photonic crystals have been adopted in order to utilize the surface plasmonic wave scattering or localized surface plasmonic resonances focused mostly on the interfaces between semiconductors and metal [5–8]. Unfortunately, optical losses are also caused by the metallic back contacts and are an unavoidable side effect that decreases the efficiency of the solar cell operation. Here, we theoretically analyze the tradeoff between the plasmonic enhancement and optical loss that appears when we introduce a metal back reflector into a SiNW solar cell. In thin-film a-Si:H or nanocrystalline silicon solar cells, numerous previous works [9–12] have reported on the optical losses by metallic back contacts. Except for short wavelengths showing some resonant absorption peaks, the optical absorption losses by a rough, nanoscale Ag back reflector was found to be extremely low in the 500−1100 nm wavelength range [9]. Regarding the optical loss in SiNW solar cells, however, little is known about the influence of a metallic back contact. A SiNW array employing a radial p-n junction has been proposed as an attractive candidate for next-generation photovoltaics owing to its low reflection, broadband light absorption, and efficient collection of photogenerated carriers [13–17]. The additional gain of ~4% in conversion efficiency could be estimated by coating a perfect mirror (100% reflectivity) on the backside of a nanowire array in which the NW diameter is 80 nm [17]. Si wire arrays with and without a Ag back reflector have been also investigated using photoelectrochemical characterizations [18]. In previous work, our group experimentally characterized the optical effect of an Al back reflector on a Si-wire-embedded PDMS film using various geometric parameters [19]. Although prior reports clarified that the low reflectivity originated from both the plasmonic effect and the nanostructured Al gratings, more detailed research is necessary to optimize the embedded depth of the NW bottoms into the silver back reflectors, particularly because the relation between the plasmonic

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A778

enhancement (positive impact) and the optical loss (negative impact) depends greatly upon the embedded depth, p, of the NW bottoms, as shown in Fig. 1. Conceptual schematics comparing the distinctive features of light trapping between Si thin films and SiNW solar cells are shown in Fig. 1(a) and 1(b). In a SiNW array, we need to further consider the following three optical routes, which are absent in its thin film counterpart: i) Each Si wire acts as an individual cylindrical waveguide so that the incident light is gradually absorbed along the axial direction while coupling with guided modes inside a wire. The light remains until the wire bottom couples with the metallic back contacts, creating an ohmic loss. ii) Multiple scattering occurs in the SiNWs, and their periodicity excites surface plasmon polaritons (SPPs) travelling along the Ag surfaces that contacted to air or silicon under the phase matching condition. This process also causes light absorption loss at the corresponding wavelengths. iii) Localized surface plasmonic resonances (LSPR) appear in the enhanced local near-field for both the metal contacts and the SiNWs. Additionally, the localized near-field contributes not only to the enhancement of light absorption, but also to the optical loss over the spectral range where such resonances occur.

Fig. 1. Conceptual schematics showing light trapping and optical losses in Ag BR for (a) a Si thin-film solar cell and (b) a Si nanowire solar cell. (c) A three-layer structure consisting of a NW array, a thin nanohole-contact-grating layer, and a thick flat BR. (d, e) Top-view and cross-section of a unit cell. (f) Quarter-view of a unit cell and the location of the boundary conditions. The probe line ‘AB’ (denoted in red) is set along the center of a nanowire, and a probe area ‘C’ (blue) is set close to the bottom of the nanowire.

We theoretically investigated the optical properties of a SiNW array solar cell using the following three different geometries: a bare SiNW array, a SiNW array with a flat Ag BR, and a SiNW array with a wire-embedded Ag BR. In the last formation, the bottoms of a NW array were embedded into a BR, thus forming a nanohole-grating-structure, as shown in Fig. 1(c). A three-dimensional simulation was performed using a finite element method (FEM) implemented in the RF module of COMSOL Multi-Physics, which attempted to determine the optical relation between the plasmonic enhancement in SiNW solar cells and the light losses from the Ag back contact. #173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A779

2. Optical models and simulation approach As depicted in Fig. 1(c), we can divide the overall geometry into three layers: a NW array, a thin nanohole-contact-grating (NCG) layer, and a flat, thick Ag BR. Their corresponding, two-dimensional, top and cross-sectional views are shown in Fig. 1(d) and 1(e), where d , l , a , and p correspond to the wire diameter, wire length, lattice constant, and embedded depth of NWs into the Ag BR, respectively. This geometrical modeling is based on the experimental fabrication of SiNW solar cells [18], in which the SiNW array is fabricated using vapor-liquid-solid (VLS) or vapor-solidsolid (VSS) growth. Note that a SiNW array can also be peeled off a grown substrate after immersion into a polydimethylsiloxane (PDMS) polymer [20–22]. The metallic BR was easily formed via deposition onto the backside of a vertical SiNW array [19]. The surface profile of a detached PDMS bottom face is generally concave in between nanowires because the wire bottoms normally protrude out of the PDMS, which causes the deposited metallic BR to always have a nanohole array grating morphology. In this work, we emulated such a nanograting structure to be a thin NCG layer in contact with an optically thick Ag BR in which the wire bottoms were embedded with a thickness p, as shown in Fig. 1(e). The process formed the nanohole layer. In our simulation, the incident light was set to be an x-polarized plane wave propagating along the −z direction. Perfect magnetic and electric conductors (PMCs and PECs) were used on the sidewalls of a unit cell for generating either symmetric magnetic field or symmetric electric field on the two sides of the sidewalls [23]. On a PMC (or PEC) wall, there was no parallel magnetic field (or electric field) component. As a result, the simulated region was simplified to a quarter of its initial size, as shown in Fig. 1(f). Perfect matched layers (PMLs) were used at the top and bottom of a unit cell, which was capped by the scattering boundaries. The wavelength-dependent optical constants of crystalline silicon (c-Si) and Ag were obtained by fitting the experimental data taken from Ref [24] using a linear interpolation. Our simulation has been carried out by COMSOL Multiphysics in frequency domain in which the Maxwell Equations were used for a sweep on wavelengths. The maximum mesh size was chosen to be 40 nm in air, and 10 nm in Si for accurate calculation. The grids near the Ag/air and Ag/Si interfaces were chosen to be ≤2 nm for observing the plasmonic effects. The portion of light absorbed by SiNW and the energy loss in the Ag BR was calculated via integration within their corresponding volumes: ASi, Ag ( λ ) =

 Q

Si, Ag

( λ , x , y , z ) dV E in

(1)

where E in is the energy of incident light in the given unit cell. Q Si and Q Ag are the electromagnetic power loss densities in the SiNW and Ag BR, respectively, calculated by Q Si, Ag ( λ , x , y , z ) =

2 1 '' ε Si, Ag ( λ ) E ( x , y , z ) 2

(2)

'' ε Si, Ag is the imaginary part of the permittivity of Si or Ag, and E ( x , y , z ) is the electric field calculated by COMSOL Multiphysics.” To evaluate the overall absorption performance of the different geometries in the solar spectrum, we calculated the ultimate efficiency, η UE at Air Mass (AM) 1.5 direct normal and circumsolar conditions using the following equation:

η UE =



λg

310nm

I ( λ ) ASi ( λ )



4000nm

310nm

#173008 - $15.00 USD (C) 2012 OSA

λ dλ λg

I (λ )d λ

(3)

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A780

where I ( λ ) is the spectral irradiance at AM 1.5 direct normal conditions in W/m2/nm. The

λg = 1127nm corresponds to a bandgap energy (1.1 eV) of Si. The lower limit (310 nm) of the integral is 4 eV, and the upper limit (4000 nm) in the denominator indicates the limit of the data available for a solar spectrum. Equation (1) assumes that each photon with energy greater than the bandgap energy photogenerates only one electron-hole pair with an energy equal to λg , whereby the excess energy is converted to heat. Likewise, in a spectrum ranging from 310 to 1127 nm, the total energy losses by the Ag back contact could be expressed as:

η Loss

 =

λg

I ( λ ) A Ag ( λ ) d λ

310nm 4000nm



310nm

I (λ )d λ

(4)

3. Results and discussion We set the length of NWs to 2.33 μm, comparable to the film thickness of thin film solar cells. Ag was chosen as the BR material because it possesses the lowest absorption coefficient among the candidate metals in the visible range; moreover, it is easily deposited onto the backside of SiNW arrays.

3.1 SiNW arrays with a flat Ag BR for small wire diameters and lattice constants We investigated SiNW arrays that had relatively smaller values for both their wire diameters and lattice constants. The wire diameter was set to 60 nm. According to the optical waveguide theory [25], each SiNW can be treated as a step-profile fiber defined by so-called fiber

parameter, V = 2πρ / λ nco2 − ncl2 , where ρ , λ , nco , and ncl are the fiber radius, light wavelength, refractive index of core and cladding layers of a fiber, respectively. In each SiNW, a number of guided modes coexist, but in our case using a small wire diameter, the fiber parameter V is less than the cutoff value of 2.405. Consequently, the fundamental mode, HE11 only exists to make each NW behave as a ‘single-mode NW’. Under illumination, a fundamental mode is excited when phase matching occurs between the air and the SiNW array [26]. For SiNW arrays with a diameter of 60 nm, such phase matching is calculated to occur at a wavelength of ~425 nm. For clarity, the electric field distribution of the HE11 mode at a wavelength of 425 nm is depicted in the inset of Fig. 2(a) using the formulas provided in Ref [25]. In fact, such phenomenon could also be interpreted semi-analytically to couple into a set of so-called key modes of a SiNW array, as shown in Refs [27, 28]. The wavelengthdependent absorptances are plotted in Fig. 2(a) and 2(b) for the SiNW arrays without and with a flat Ag BR, respectively. The guided mode coupling between the air and the SiNW array reveals the presence of an evident peak around ~425 nm in the absorptance spectra. For solar cell applications, it is desirable to use SiNW arrays with a higher filling ratio in order to obtain a high efficiency in energy conversion. Figure 2(c) shows the optical properties of SiNW arrays with the smallest lattice constant of 100 nm, where the blue and black lines correspond to the light absorption in Si with and without a Ag BR, respectively. The absorptance of the Ag substrate in air was also calculated for comparison using the Fresnel equation A = 1 − ( n − i κ − 1) / ( n − i κ + 1) , where n and κ are the real and imaginary parts of the refractive index of Ag, respectively [24]. Given the optical loss by the Ag BR, the broadband optical loss was observed for wavelengths from 450 to 700 nm, where the highest loss occurred at ~600 nm revealing the joule heating converted from ~60% of the incident photon energy. Nevertheless, the presence of the Ag BR improved η UE to 12.2% compared to the bare SiNW array surrounded with vacuum (without Ag BR and a substrate), which gave a η UE of 8.5%. Although the optical loss of only ~1.64% occurred at the air/Ag interface, ~8.1% of the total incident light was estimated to thermally dissipate in the Ag BR, 2

#173008 - $15.00 USD (C) 2012 OSA

2

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A781

where most of the optical loss occurred in the wavelengths from 450 to 700 nm. This feature is clearly different from the loss spectra observed in crystalline silicon thin-film solar cells [9–11], which only show some resonant absorption peaks in shorter wavelengths less than 500 nm.

Fig. 2. (a) Absorptance spectra of SiNW arrays without Ag BR. (b) Absorptance spectra of SiNW arrays integrated with a flat Ag BR. (c) Optical properties of SiNW arrays that have a lattice constant of 100 nm and a wire diameter of 60 nm. The blue and black lines correspond to light absorption by Si with and without Ag BR, respectively. The solid red line corresponds to the absorption loss by the Ag BR with the same geometry, and the green line shows the reflectance. The dashed red line shows the absorption loss at the Ag/air interface calculated by Fresnel equations. (d) The line plots show the light intensities along the probe line ‘AB’ at four typical wavelengths labeled with (i-iv), as denoted in panel c. The insets show the twodimensional electric field distribution as a function of different wavelengths for the probe rectangle ‘C’ depicted in Fig. 1(f).

To further clarify this feature, the electric field distribution along the center of the Si nanowire – i.e., the probe line ‘AB’ shown in Fig. 1(f) – was compared for the four different wavelengths (see Fig. 2(d)). At 400 nm, no optical losses occurred at the Ag back contact since the incident light was quickly absorbed within several hundred nanometers due to multiple scatterings in between the Si nanowires. At 600 nm, the light interacted strongly with the Ag BR because the NW length of 2.33 μm was not long enough to completely absorb the incident light. Note that the electric field intensity at the interface between the Si and Ag BR enhanced the amplitude of the incident light by at least three times its original value. The electric field distribution in the inset (ii) of Fig. 2(d) shows that a strong enhancement occurred in both the Si and Ag regions, which implies that the plasmonic absorption (positive) by the SiNWs accompanies the optical loss (negative) by the Ag BR. At longer wavelengths, such as 800 nm (iii) and 1000 nm (iv), the interactions between the incident light and the Ag BR are insignificant because almost all the incident light was back-reflected due to the low absorptance in both the SiNWs and the Ag BR.

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A782

3.2 SiNW arrays with a flat Ag BR for large wire diameters and lattice constants The ultimate efficiency of an optimized SiNW array with a large lattice constant of ~500 nm and a wire filling ratio of 0.5~0.64 has been reported [16] to be greater than the ultimate efficiency of equally-thin planar film solar cells because of the enhanced field concentration inside the nanowire, as well as the excitation of guided resonance modes. In Fig. 3(a), the light absorptances by SiNW arrays are compared to samples with and without Ag BR for a lattice constant of 500 nm and a wire diameter of 300 nm. Given a refractive index of 3.5 for silicon, the wavelengths of sunlight generally range from 100 to 300 nm (λ = λ0/n) inside silicon. As a result, a number of guided modes can be generated in Si wires that have wire diameters significantly larger than the light wavelength, according to the waveguide theory [25]. Hence, it is reasonable that the wires in Fig. 3(a) show more than one guided mode in each wire, suggesting that the coupling behavior of guided modes between wires is quite complicated. It is noteworthy that such complex absorption spectra could be also understood using a novel semi-analytic modal method, in which those characteristics were predicted by the coupling of incident light into the key modes of SiNW arrays [27, 28]. Many of the evident loss peaks were distributed over the broadband spectrum while overlapping with their absorption resonances, as shown in the red curve of Fig. 3(a). This feature implies that the optical losses by the Ag BR were not only associated with the surface plasmon excitations, but also with the guide-mode excitations in SiNWs. Nevertheless, the SiNWs integrated with a flat Ag BR achieved a higher η UE of 26.7%, compared to the η UE of 22.7% in the bare SiNW array. The total amount of optical loss was calculated to be 3.5% from Eq. (3). Although this value is almost twice the loss portion calculated from the Ag/air interface, it is significantly less than the loss amount (8.1%) of the NW array with a = 100 nm and d = 60 nm. This reduction of optical loss in the wire array with a large lattice constant can be attributed to the enhanced light absorption in the SiNWs due to the guide-mode excitation and multiple scatterings. These processes then directly decrease the light coupling with the Ag BR because the amount of light that is left to interact with the Ag BR decreases. In Fig. 3(b), the electric field distributions along the probe line ‘AB’ shown in Fig. 1(f) are plotted for three wavelengths of (i) 833 nm, (ii) 983 nm, and (iii) 1000 nm. At 833 nm, the energy loss by the Ag BR reached almost 50% of the total incident light, which shows the notable increase in the electric field intensity by a factor of ~15 at the Si/Ag BR interface. At 983 nm, the electric field was observed to increase by a factor of 5 in the Si nanowire compared to the intensity of incident light. At 1000 nm, however, the electric field intensity in the Si nanowire decreased to only ~20% of the value in 983 nm. In other words, the incident light for a wavelength of 983 nm appears to be highly concentrated in Si nanowires compared to other wavelengths, suggesting that the amount of light remaining is larger accordingly just before reaching the metal back reflector. Given the almost stationary extinction characteristics of Ag for long wavelengths, the optical loss is overall governed by the light intensity, which can explain why each absorption peak observed in a SiNW array accompanies the optical loss peak for long wavelengths.

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A783

Fig. 3. (a) Light absorptances of SiNW arrays (with a lattice constant of 500 nm and diameter of 300 nm) are compared to the absorptance (loss) by a Ag BR, where the blue and black lines correspond to the light absorption by Si with and without the Ag BR, respectively. The red line shows the absorption loss by the Ag BR. (b) The electric field distributions along the probe line ‘AB’ inside the SiNW are compared for the three wavelengths, (i) 833 nm, (ii) 983 nm, and (iii) 1000 nm. The insets show the corresponding two-dimensional plots of electric field for the probe rectangle ‘C’.

3.3 SiNW arrays integrated with a thin NCG layer The tradeoff between the optical loss (by the Ag BR) and the absorption enhancement (in SiNWs) was also investigated using the SiNW arrays integrated with a thin NCG layer, as shown in Fig. 1(c). Subwavelength nanoholes patterned in a thin metal film is known to cause various intriguing optical features, such as strongly enhanced light transmission and wavelength filtering via surface plasmonic resonances [29]. Thin-film solar cells using the

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A784

Ag-coated nanoholes as a photonic substrate have been reported to obtain higher external quantum efficiencies compared to those employing a back contact with random roughness [30]. Given these previous reports, we attempted to resolve the trade-off relation between the light absorption in SiNWs by surface plasmon resonances (increasing η UE ) and the optical loss in Ag NCG by absorption (decreasing η UE ) since the optical loss peaks were welldeveloped around the resonance wavelengths in our work.

Fig. 4. (a) The ultimate efficiencies calculated using three different geometrical parameters are shown as a function of the embedded depth of Ag nanoholes. For reference, each dashed line (color-matched) denotes the ultimate efficiency value of bare SiNW arrays without a Ag BR. (b) Optical losses by the Ag BR are also shown for comparison.

Three different geometric conditions, i) a = 100 nm and d = 60 nm, ii) a = 150 nm and d = 100 nm, iii) a = 500 nm and d = 300 nm, are compared in Fig. 4 in order to analyze the effects on the ultimate efficiency (Fig. 4(a)) and on the optical losses (Fig. 4(b)) by the Ag BR as a function of the embedded depth of Ag nanoholes. In bare SiNW arrays without a Ag BR, the ultimate efficiencies were calculated to be 8.5%, 15.2%, and 22.7% for wire diameters of 60, 100, and 300 nm, respectively, (see dashed lines in Fig. 4(a)). These ultimate efficiency values are in accordance with those from Ref [16], who reported that η UE in SiNW arrays increases with increasing lattice constants without changing the filling ratio. To analyze the influence of the Ag NCG layer on the photovoltaic performances, the embedded depths of the nanoholes were varied from 10 to 50 nm in the Ag BR. After integrating the Ag BR with NCG, the ultimate efficiencies were found to have slightly increased by 3~6%, as shown in Fig. 4(a). The highest η UE was obtained at the embedded depth of 20 nm. More specifically, the SiNW array (a = 100 nm and d = 60 nm) integrated with a 20-nm embedded depth NCG layer exhibited a η UE of 13.3%, which is ~43% higher than the ultimate efficiency in a bare NW array and ~7% higher than that in a NW array with a flat Ag BR. Similar results were also observed for the SiNW arrays in the other two geometric conditions. In clear contrast to the ultimate efficiency behavior, the optical energy loss by the Ag BR was notably affected by the NCG depths, especially for smaller NW diameter values and lattice constants (Fig. 4(b)). The optical loss by the NCG layers was observed to dramatically increase upon initiation of embedding the nanoholes. Highest energy loss, η Loss in AgBR reaches ~14.5% for the SiNW array ( a = 100 nm, d = 60 nm) solar cells at the 20-nm embedded depth of a NCG layer. This energy loss value is then gradually decreased as the depths are deeper. Note that the SiNW array with a larger lattice constant recorded the highest η UE value with the minimum amount (1 are separated into the first zone (denoted ‘II’ in Fig. 5(d)) – centered on 700 nm – and the second zone (denoted ‘III’ in Fig. 5(d)), which ranged over 800−1000 nm. The region showing the AE 700 nm where constructive light interferences are feasible. 3) The optimal value of embedded depth was found to be ~20 nm, at which the light absorption by SiNWs was further enhanced by ~5%. 4. Conclusions

This work theoretically investigated the optical properties of SiNW array solar cells using a FEM method while specifically comparing the flat and NCG Ag BR contacts. Two main routes for optical loss were found to coexist in such SiNW geometries: 1) near-field absorption by localized surface plasmon, and 2) guided mode coupling of the Si nanowire to the Ag BR. Optical losses in the SiNW array solar cells mostly occurred at long wavelengths, which were remarkably larger than those in thin-film solar cells. Highest energy loss in AgBR was almost ~14.5% for the SiNW array ( a = 100 nm, d = 60 nm) solar cells, but for the SiNW array with a = 500 nm, d = 300 nm, the energy loss was only ~5%. Despite the optical losses by the Ag BR, the highly enhanced near-field caused by the localized SPR at the bottom of the SiNW was helpful in efficiently promoting the ultimate efficiencies of the SiNW solar cells after integrating them with a NCG Ag BR, where the optimal embedded wire depth was ~20 nm. Acknowledgments

This work was supported by the New & Renewable Energy R&D Program (No: 2009T100100614)) and Human Resources Development (No. 20104010100620) of the Korea Institute of Energy Technology Evaluation and Planning(KETEP) grant funded by the Korea government Ministry of Knowledge Economy. This work was also supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST) (No 2011-0028604).

#173008 - $15.00 USD (C) 2012 OSA

Received 20 Jul 2012; revised 30 Aug 2012; accepted 30 Aug 2012; published 7 Sep 2012 10 September 2012 / Vol. 20, No. S5 / OPTICS EXPRESS A787