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

Broadband short-range surface plasmon structures for absorption enhancement in organic photovoltaics

Open Access Open Access

Abstract

We theoretically demonstrate a polarization-independent nanopatterned ultra-thin metallic structure supporting short-range surface plasmon polariton (SRSPP) modes to improve the performance of organic solar cells. The physical mechanism and the mode distribution of the SRSPP excited in the cell device were analyzed, and reveal that the SRSPP-assisted broadband absorption enhancement peak could be tuned by tailoring the parameters of the nanopatterned metallic structure. Three-dimensional finite-difference time domain calculations show that this plasmonic structure can enhance the optical absorption of polymer-based photovoltaics by 39% to 112%, depending on the nature of the active layer (corresponding to an enhancement in short-circuit current density by 47% to 130%). These results are promising for the design of organic photovoltaics with enhanced performance.

©2010 Optical Society of America

Photovoltaic devices that convert solar energy into electrical energy have the potential to provide a virtually unlimited source of energy that is renewable and environmentally benign. However, to be competitive with fossil-fuel technologies, the cost of current photovoltaic technologies needs to decrease substantially. At present, the solar cell market is mainly based on crystalline and polycrystalline silicon with a thickness of approximately several hundred microns, resulting in high cost for materials and processing. Therefore, there is great interest in thin-film solar cells [1], with film thicknesses of a few hundred nanometers, which can be deposited on different substrates like glass and plastics. To date, thin-film solar cells have been fabricated using various active materials, including amorphous silicon [2], GaAs [3], CuInxGa1-xSe2 and CdTe [4,5], as well as organic semiconductors [69]. Compared with their inorganic counterparts, organic photovoltaics (OPVs) based on conjugated polymer and fullerene composites can be fabricated over large areas by means of low-cost ink-jet printing and coating technologies. The simultaneous patterning of the active materials on lightweight flexible substrates raises the prospect of organic photovoltaics potentially as inexpensive as paint [1012]. However, the low charge-carrier mobility and small exciton diffusion length of most molecular and polymeric materials [13,14] limit the thickness of the active layers (30-150 nm) [1519] in OPVs, and lead to poor solar light absorption. This in turn results in insufficient carrier generation and low power conversion efficiency.

To overcome this thickness limitation, light trapping strategies may be explored in the design of OPVs to achieve high optical absorption. The traditional light trapping strategy in bulk solar cells typically employs random surface textures at the micrometer level [2022], which are larger than the active layer of OPVs and no longer suitable for thin film photovoltaics. Consequently, researchers proposed novel plasmonic nanostructures to achieve effective light trapping for thin-film solar cells [23]. Surface plasmon polaritons (SPPs) are collective oscillations of free electrons at the boundary of a metal and a nonconducting dielectric or semiconductor material [2426]. By properly engineering novel plasmonic nanostructures, light can be concentrated into the thin active layer, thereby increasing its absorption. For device optimization, a broad absorption enhancement is desirable to improve the performance of photovoltaic devices. In recent years, researchers employed various nanoplasmonic structures to improve the performance of solar cells, including nanoparticles [2735], and periodic nanostructures [3640], which have been described in a recent review [23]. In the present work, we propose a novel architecture based on an ultra-thin nanopatterned metallic structure to improve the performance of very thin-film organic solar cells. Because of the excitation of short-range SPP modes inside the absorption layer, a broad absorption enhancement is achieved. The design principles and physical mechanisms were analyzed numerically and analytically. The improved device performance was evaluated systematically through calculations of the short circuit current density based on the enhanced absorption in the active layer. To date, the best reported power conversion efficiency for polymer-based OPVs is approximately 7% [9,15,16]. We believe that introducing a metal electrode patterned with a plasmonic nanostructure is a promising approach to surpass this experimentally achieved record on the power conversion efficiency.

Figure 1 shows a schematic diagram of the proposed organic solar cell. The structure consists of a glass substrate, on which is deposited a 20nm thick indium tin oxide (ITO) layer that serves as a transparent anode, a 10nm thick highly conductive hole transport layer, poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS), and a 30nm thick active layer. The active layer used in our design is an organic bulk heterojunction system consisting of the electron-donor material regioregular poly(3-hexylthiophene) (P3HT) and the electron-acceptor fullerene derivative [6,6]-phenyl-C71 butyric acid methyl ester (PC70BM). This is followed by a 20nm nanopatterned silver structure with a periodic subwavelength hole array that is placed immediately above the active layer. Figure 1 shows the subwavelength holes with numerically tunable diameters (D) and periods (P), which are square-symmetrically etched on the Ag layer and assumed to be air-filled.

 figure: Fig. 1

Fig. 1 A schematic diagram of the proposed plasmonic organic solar cell.

Download Full Size | PDF

In order to investigate optical absorption enhancement in this solar cell system, we perform full-field electromagnetic simulation using the three dimensional (3D) finite-difference time domain (FDTD) method. First, to demonstrate the reliability of our modeling, we performed a 3D simulation of the optical absorbance of this OPV structure. Our calculation of the optical absorbance of a P3HT:PC70BM bulk heterojunction composite film agrees well with previously reported experimental results for an identical structure [9,41], (see Fig. 6 in Appendix A), validating the predictive capability of our simulation for reproducing experimental results.

 figure: Fig. 6

Fig. 6 Comparison between the simulated and measured absorbance [see the red dots, reproduced from Fig. 2(b) in Ref. [9]. of a 150nm thick P3HT:PC70BM layer.

Download Full Size | PDF

Now we employ the modeling to investigate optical absorption enhancement in the proposed nanostructured OPV. The device is illuminated through the transparent anode, as shown in Fig. 1. The spectrally dispersive dielectric properties of the ITO, PEDOT:PSS [42], P3HT:PC70BM (1:0.8 in weight) [40], Ag and glass (SiO2) [43] are used in the calculation. Periodic boundary conditions are employed and polarized incident light (electric component parallel to X-axis) is assumed. First, the optical absorption of the active layer is determined by calculating the difference between the optical power incident on and transmitted through the layer. The absorption spectrum [A(λ)] is obtained by calculating the optical absorption at each single wavelength across a broad spectral range from 400nm to 900nm [9,41]. Figure 2(a) shows the absorption spectra of devices with a flat metallic structure above the active layer (dotted line), and a subwavelength metallic nanohole array (D = 150nm and P = 300nm, see the solid line). Figure 2(a) also shows the absorption enhancement spectrum (dashed line), which is given by the ratio of the absorption spectra using a nanopatterned and a flat metal film. The results show a large absorption enhancement, with a factor of 6 observed at a wavelength of 725nm. This enhancement is consistent with previous observations on nanopatterned ultra-thin metal films [44] that less light is transmitted through an ultra-thin nanopatterned metal film than through a simple thin metal film. This absorption enhancement was attributed to the excitation of short-range SPP modes, although no field distribution analysis was provided to support this prediction.

 figure: Fig. 2

Fig. 2 Calculations on the device with the nanopatterned metallic structure of diameter D = 150nm, and period P = 300nm. (a) Solid line - the absorption spectrum with a nanopatterned metallic structure; dotted line - the absorption spectrum with a flat metallic structure; and dashed line - the absorption enhancement spectrum. (b) and (c) are time averaged magnetic intensity (|Hy|2) distributions at the SPP resonance wavelength. (b) is the intensity distribution in the x-y plane, and (c) is the intensity distribution in x-z plane. The spatial mode profile is plotted in the right panel of (c).

Download Full Size | PDF

To investigate the physics of the absorption enhancement at the peak wavelength of 725nm, the time-averaged magnetic field intensity (|Hy|2) distribution at the P3HT:PC70BM/Ag interface was calculated and is plotted in Fig. 2(b) and 2(c). Figure 2(b) shows the intensity distribution in the X-Y plane, which demonstrates that the SPP mode is mainly confined in the space between holes. In the X-Z plane shown in Fig. 2(c), the SPP mode is confined in the active layer adjacent to organic/Ag interface. To visualize the mode distribution more clearly, the spatial mode profile is plotted in the right panel in Fig. 2(c). However, based on the intensity distribution shown in Fig. 2(b) and 2(c), it is still not clear whether the SPP modes are short range SPPs, as previously predicted [44].

It is well known that for an optically opaque metal film, there is only one type of SPP mode, the single-interface surface plasmon polariton [24]. For thinner metal films (usually less than 100 nm), the two single-interface SPPs interact with each other and lead to two coupled SPPs, the long-range and short-range surface plasmon polariton [45]. The long-range mode has an asymmetric charge distribution between the top and bottom surfaces with the electric field predominantly normal to the surface inside the metal. Conversely, the short-range surface plasmon polariton (SRSPP) has a charge distribution which is symmetric between the top and bottom surfaces with the electric field essentially parallel to the surface [46]. Consequently, we can determine the physical origin of the SPP modes by calculating their surface charge distributions. Figure 3(a) and 3(b) show the electric field distributions for the device discussed in Fig. 2 (t = 20nm, D = 150nm and P = 300nm). The time-averaged electric field strength is shown in Fig. 3(a) with the instantaneous EXZ vector and surface charge distributions in the X-Z plane. One can see that the spatial variation of the surface charge distribution on the top surface is clearly in phase with that on the bottom surface (symmetric distribution). The EZ vector distributions at the top (P3HT:PC70BM/Ag) and bottom (Ag/Air) interfaces are displayed in Fig. 3(b). The observed antisymmetric EZ field pattern also corresponds to a symmetric surface charge distribution, thus demonstrating that the SPP modes excited in the device shown in Fig. 2 are SRSPP modes.

 figure: Fig. 3

Fig. 3 (a) and (b) are electric field distributions at the SRSPP resonance wavelength at 725nm. (a) Time-averaged (color-scale) and instantaneous (arrows) electric field strengths and surface charge distribution in x-z plane, and (b) Instantaneous EZ vector distribution at the top and bottom surfaces of the Ag nanostructure. (c) Map of the absorption enhancement versus wavelength and metallic nanostructure thickness. The solid arrow corresponds to the resonance wavelength of the single-interface SPP Bloch mode. The dashed line corresponds to the analytical solutions of the SRSPP Bloch modes.

Download Full Size | PDF

The next issue that arises after confining the mode in the absorption layer is to study the coupling mechanism. It is known that the momentum mismatch between SPPs and free-space light can be bridged by Bragg vectors inherent in the periodic nanostructures [24,47]. The nanopatterned metallic structure in the proposed OPV structure enables incident light to couple to SPP Bloch modes. The Bragg grating vectors can be expressed as Gij = i Gx + j Gy, where i and j are integers. |Gx| = |Gy| = 2π/P are the reciprocal lattice vectors of the square array. For normal incidence, the in-plane wave-vector is zero. Consequently, the in-plane wave-vector can be expressed by:

kinplane=Gij=iGx+jGy.
For an optically opaque metal film, only one SPP mode exists at the interface, e.g. the single-interface SPP [24]. Its dispersion relation can be described as follow:
|kSPP|=ωcεdεmεd+εm,
where εd and εm are the relative permittivities of the dielectric and the metal. The dispersion relation of the single-interface SPP Bloch mode generated under the Bragg coupling condition can be obtained approximately by combining Eqs. (1) and (2) when kin-plan = kspp [47,48]. In addition, the dispersion relations for long-range and short-range SPPs can be described by the following equation [45,49]:

tanh(S2t)(εd1εd2S22+εm2S1S3)+[εmS2(εd1S3+εd2S1)]=0.

Here S1, S2, and S3 are defined as S12 = kSPP2d1k02, S22 = kSPP2mk02, S32 = kSPP2d2k02, k0 = ω/c, and t is the thickness of the metal film. In our design εd1, εd2, and εm are the dielectric constants of P3HT:PC70BM, air (εd2 = 1), and Ag, respectively. Under the Bragg coupling condition, the dispersion relation of the SRSPP Bloch mode can be obtained by combining Eqs. (1) and (3). The numerical map of the absorption enhancement versus wavelength and the metallic nanostructure thickness is plotted in Fig. 3(c). The analytical solutions for the SPP modes excited by the lowest order Bragg vectors {(i,j) = (1,0),(−1,0),(0,1),(0,-1)} are also plotted in this figure to verify the generation of the SPP Bloch modes. The black dashed line represents the SRSPP Bloch mode excited under the Bragg coupling condition as a function of metal film thickness (t). The excitation frequency of the single-interface SPP Bloch mode is calculated by combining Eqs. (1) and (2) (indicated by the red solid arrow). As shown in Fig. 3(c), these analytical predictions agree well with the numerical simulation of the absorption enhancement obtained by the FDTD modeling. When the thickness decreases from 100nm to 10nm, one can see that the SPP Bloch mode transits from single-interface SPP to SRSPP gradually, with an obvious red-shift of the resonant wavelength. More importantly, the bandwidth of the absorption enhancement becomes wider when the mode enters the SRSPP region. This is because the complex wave-vectors obtained from Eq. (3) have larger imaginary parts than that from Eq. (2). As shown in Fig. 3(c), the width of the enhancement spectrum at the thickness of 20nm is approximately twice that at 100nm. Consequently, these SRSPP Bloch modes on a thinner metallic structure can provide broader band absorption.

Tuning the SRSPP resonant wavelength may be accomplished by tailoring the nanopatterned metallic structure. The absorption enhancement band should be tuned spectrally in order to achieve more effective utilization of the solar energy (see Fig. 7 in Appendix B). According to Eqs. (1) and (3), the SRSPP resonant wavelength is strongly related to the period (P) of the hole array. For simplicity, here we set the metallic nanostructure thickness to 20nm, and tune the period with a constant period/diameter ratio of 2. The map of the absorption enhancement versus wavelength and period for nanostructured P3HT:PC70BM cells is plotted in Fig. 4(a) . The analytical solution of the SRSPP Bloch mode is also plotted in this figure (see the black dashed line), correlating well with the absorption peak position, and verifying that the absorption enhancement results from SRSPPs. One can see that the SRSPP resonant peak exhibits a red-shift as the period is increased. Importantly, this nanopatterned ultra-thin metallic structure design is quite general and can be employed for different organic active layers. Figure 4(b) shows similar results for another polymer, a composite blend of poly[2,6-(4,4-bis-(2-ethylhexyl)-4H- cyclopenta [2,1-b;3,4-b’]dithiophene)-alt-4,7-(2,1,3-benzothiadiazole)] (PCPDTBT) and [6,6]-phenyl-C61 butyric acid methyl ester (PCBM). The calculation for this solar cell was performed by simply replacing the P3HT:PC70BM layer with a PCPDTBT:PCBM layer (1:3 in weight) [9,40]. Figure 4(b) shows the map of the absorption enhancement versus wavelength and period for the nanostructured PCPDTBT:PCBM cell. The analytical solution of the SRSPP Bloch mode is also plotted in this figure (see the black dashed line), showing good agreement with the absorption peak position. Based on the results shown in Fig. 4, one can see that the enhancement spectra originating from the SRSPP Bloch modes may be tuned over a very broad range of wavelengths, which is highly desirable for enhancing OPV performance.

 figure: Fig. 7

Fig. 7 The photon flux density of the solar spectrum.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Maps of the absorption enhancement versus wavelength and period for two organic solar cells, i.e. a P3HT:PC70BM cell (a), and a PCPDTBT:PCBM cell (b). The dashed lines correspond to the analytical solutions of the SRSPP Bloch modes.

Download Full Size | PDF

Next, we investigate the plasmon-assisted absorption enhancement and the resulting performance improvement of these two OPV devices under solar irradiance for the standard air mass (AM1.5). For unpolarized incident sunlight, the square-symmetrically distributed nanostructures exhibit polarization-independence. Incident light with electric field components parallel to X-axis or Y-axis will excite the same SRSPP Bloch modes. According to previous reports [44], the resonant frequency shifts only slightly when the electric field is polarized along the diagonal of the square-symmetric array. For this polarization-independent solar cell design, the optical absorption is described by the expression A=λminλmaxA(λ)S(λ)dλ, where S(λ) is the solar irradiance spectrum for the standard air mass (AM1.5). The absorption enhancement (relative to a reference structure with a flat metallic layer) is given by (A/Aref −1). It is also useful to define the absorbed photon flux Фabs(λ) (number of photons absorbed per unit time and unit area) at a single wavelength as Фabs(λ) = A(λ)S(λ)/Ephoton(λ), where Ephoton(λ) = hc/λ denotes the energy of a single photon, and h is Planck’s constant. In the following calculations, we assume for convenience that every absorbed photon generates one exciton, which dissociates into holes and electrons at the donor/acceptor interface, and that all photo-induced charge carriers are collected by the electrodes [37,40]. Under these ideal conditions, the short circuit current density JSC of the prototype device can be calculated using JSC=eλminλmaxΦabs(λ)dλ, where e is the charge of an electron. As a point of reference, the JSC for the P3HT:PC70BM device with a flat metallic layer is calculated to be JSC-ref = 6.07 (mA/cm2). In this case, the enhancement of JSC is given by (JSC/JSC-ref −1). To demonstrate the achievable enhancement, the short circuit current density of the P3HT:PC70BM cell is calculated as a function of the period, P, of the nanopatterned metallic structure [shown in Fig. 5(a) ]. The peak value of JSC is found to be 8.9 (mA/cm2) at an optimal period P of 260nm. The resulting enhancement in JSC is approximately 47% (corresponding to a 39% enhancement in optical absorption). Large JSC enhancements (over 40%) can be achieved over a broad range of periods (200nm to 400nm in this simulation), indicating a good fabrication tolerance for practical devices that can be readily realized using current nanofabrication procedures. The inset of Fig. 5(a) shows the absorbed photon spectra with a nanopatterned metallic nanostructure (solid line, P = 260nm) and a flat Ag metallic structure (dotted line). The photon absorption at wavelengths near 680nm is increased greatly relative to the reference with a flat metallic layer. With this metallic nanopatterned design, light at shorter wavelengths is effectively absorbed by the cell, while the weakly absorbed red and infrared light is trapped at the back surface leading to increased absorption. This provides greater enhancement than the front surface design, which can suppress absorption at short wavelengths because of destructive interference [23,34].

 figure: Fig. 5

Fig. 5 (a) The short-circuit current density (JSC) and its corresponding enhancement versus period of the nanopatterned metallic structure for the P3HT:PC70BM cell. The inset shows the absorbed photons spectra of the device with a nanopatterned metallic structure (solid line, P = 260nm) and a flat metallic surface (dotted line). (b) JSC and its corresponding enhancement versus period of the nanopatterned structure for the PCPDTBT:PCBM cell. The inset shows the absorbed photons spectra of the device with nanopatterned metallic structure (solid line, P = 320nm) and flat metallic surface (dotted line).

Download Full Size | PDF

Similarly, Fig. 5(b) shows the JSC enhancement for a PCPDTBT:PCBM solar cell. The reference short circuit current density for the PCPDTBT:PCBM cell, JSC-ref, is calculated to be 3.62 (mA/cm2). As shown in Fig. 5(b), the JSC of the plasmonic-assisted device is dramatically improved to 8.36 (mA/cm2) with an impressive enhancement factor of ~130% in Jsc at the optimal period of 320nm (corresponding to an optical absorption enhancement of 112%). The large enhancement is achieved because the absorption coefficient of the PCPDTBT:PCBM polymer is relatively small over the entire solar spectrum [9,41] (with a small absorption peak at 750nm), resulting in a small value of JSC-ref. If the SRSPP modes are excited around the intrinsic absorption peak at 750nm, the short circuit current density enhancement could therefore be improved significantly. The inset of Fig. 5(b) shows the absorbed photon spectra with a nanopatterned metallic structure (solid line, P = 320nm) and a flat Ag back surface (dotted line). One can see that the photon absorption increases significantly in the spectral region around 750nm, also the resonant wavelength of the SRSPP. The enhancement occurs over a very broad spectral range, resulting in the very large improvement in device performance. The results in Fig. 5 demonstrate that the maximum potential JSC of 8.9 (mA/cm2) and 8.36 (mA/cm2) can be achieved for the two nanostructured OPVs with 30nm active layers, under the assumption of ideal charge collection (unity internal quantum efficiency). One can show from similar calculations that considerable SRSPP enhanced performance occurs for thicker OPV active layers as well (see Fig. 8 in Appendix C), although the trade-off between the actual collection efficiency of photo-generated charge carriers and the optical absorption needs to be accurately taken into account. Consequently, it is important to investigate the optimal thickness of the active layer to achieve the maximum conversion efficiency of the OPV at the device level [18]. Such a study would require detailed information of the intrinsic properties of the active materials, and an understanding of the relation between the internal quantum efficiency and the active layer thickness, and should be the subject of a future investigation, together with an experimental characterization of SPP enhanced organic OPVs.

 figure: Fig. 8

Fig. 8 The relation between the JSC and the active layer thickness of the P3HT:PC70BM device. The red curve is the JSC of the device with nanostructured back reflector; the black curve is the JSC of the device with a 20nm flat back reflector. Inset: The relation between the JSC enhancement factor (the ratio of the JSC-nano and JSC-ref) and the active layer thickness.

Download Full Size | PDF

It should be noted that an optically opaque flat metal film, rather than the ultra thin metal film considered here, is normally employed as the back reflector for conventional OPVs. Our modeling results (not shown here) revealed only a small change in the enhancement in absorbed photon density (< 10%) when the thickness of the flat back reflector increases from 20nm to 100nm. In addition, in terms of the practical device application, another issue should also be noted: for polymer OPVs, although 30-50nm thick active layers employed in our modeling were used in certain devices [7,19], the thickness is typically 100-150nm. In this case, optical waveguide modes could be introduced to play important roles in absorption enhancement. Previous studies [36] and [37] have explored the roles of these optical modes.

The SRSPP-assisted absorption enhancement results presented in this work for polymer based OPVs compare favorably (P3HT:PC70BM) or in case of (PCPDTBT:PCBM) surpass those reported previously for OPVs based on molecular systems [39] where a theoretical enhancement of 50% in optical absorption was reported for one dimensional metallic gratings placed on 15nm thick organic heterojunction. For comparison, our modeling results for a 15nm thick P3HT:PC70BM layer yields an 83% enhancement in optical absorption, which corresponds to an enhancement of 94% in Jsc (see Fig. 8 in Appendix C). Similarly, the absorption enhancement for a 15nm film of PCPDTBT:PCBM would be on the order of 200%. This indicates that a larger overlap between the enhanced absorption spectrum and the solar spectrum is achieved with the proposed ultra-thin nanopatterns.

In conclusion, an ultra-thin nanopatterned metallic structure placed next to the organic active layer is proposed for improved OPV devices. Due to the excitation of SRSPP Bloch modes, a large broadband absorption enhancement is achieved, and may be spectrally tuned by tailoring the geometric parameters of the plasmonic nanostructure. Using this nanostructure, calculations show that the short circuit current densities can be enhanced by 47% for an OPV based on P3HT:PC70BM and by 130% for one based on PCPDTBT:PCBM. It is demonstrated that large enhancements can be achieved over a broad range of periods, providing very good fabrication tolerance for practical devices. The design principle of the nanopatterned metallic nanostructure proposed in this article is quite general and can be employed to systematically engineer and optimize the parameters for improved photovoltaic devices.

Appendix A. Demonstration of the simulation reliability

Here we perform calculations, which employ experimental parameters identical to those reported in an experimental study of a P3HT:PC70BM bulk heterojunction composite film [9]. Calculated results are compared with the measured results to demonstrate the reliability of our simulations. The reported [9] measured absorbance of a ~150nm thick P3HT:PC70BM layer is shown by the red circles in Fig. 6. The solid curve in Fig. 6 shows the calculated optical absorbance of this 150nm thick P3HT:PC70BM layer, employing the experimental parameters reported in Ref. [9]. for this simulation. One can see that the simulation curve agrees reasonably well with the measured results, providing evidence of the predictive capability of our simulation and its ability to reproduce experimental results.

Appendix B. Solar photon flux spectrum

Appendix C. Estimation of JSC for different active layer thicknesses

Under the assumptions of unity internal quantum efficiency and 100% charge collection, the maximum calculated JSC of the P3HT:PC70BM device is estimated with different active layer thicknesses, as shown in Fig. 8. The parameters of the metallic nanostructure, which were optimized for a 30nm thick active layer, were kept constant as the active layer thickness was varied. One can see that considerable enhancement in JSC can be achieved with the nanopatterned metallic structures over a range of active layer thicknesses. While, the estimated enhancement is smaller for the thickest active layers, SRSPP absorption enhancement should permit effective device operation with thinner active layers. However, it must be emphasized that while the above simulation provides qualitative trends, the assumption of unity internal quantum efficiency is not valid over the range of thicknesses considered and a more rigorous calculation is needed to obtain quantitative results.

References and links

1. M. A. Green, “Third generation photovoltaics: Solar cells for 2020 and beyond,” Physica E 14(1-2), 65–70 (2002). [CrossRef]  

2. M. A. Green, “Recent developments in photovoltaics,” Sol. Energy 76(1-3), 3–8 (2004). [CrossRef]  

3. G. J. Bauhuis, P. Mulder, E. J. Haverkamp, J. C. C. M. Huijben, and J. J. Schermer, “26.1% thin-film GaAs solar cell using epitaxial lift-off,” Sol. Energy Mater. Sol. Cells 93(9), 1488–1491 (2009). [CrossRef]  

4. M. A. Contreras, B. Egaas, K. Ramanathan, J. Hiltner, A. Swartzlander, F. Hasoon, and R. Noufi, “Progress toward 20% efficiency in Cu(In,Ca)Se-2 polycrystalline thin-film solar cells,” Prog. Photovoltaics 7(4), 311–316 (1999). [CrossRef]  

5. A. Romeo, A. Terheggen, D. Abou-Ras, D. L. Bätzner, F.-J. Haug, M. Kälin, D. Rudmann, and A. N. Tiwari, “Development of thin-film Cu(In,Ga)Se2 and CdTe solar cells,” Prog. Photovoltaics 12(23), 93–111 (2004). [CrossRef]  

6. P. Peumans, S. Uchida, and S. R. Forrest, “Efficient bulk heterojunction photovoltaic cells using small-molecular-weight organic thin films,” Nature 425(6954), 158–162 (2003). [CrossRef]   [PubMed]  

7. G. Li, V. Shrotriya, J. Huang, Y. Yao, T. Moriarty, K. Emery, and Y. Yang, “High-efficiency solution processable polymer photovoltaic cells by self-organization of polymer blends,” Nat. Mater. 4(11), 864–868 (2005). [CrossRef]  

8. M. C. Scharber, D. Mühlbacher, M. Koppe, P. Denk, C. Waldauf, A. J. Heeger, and C. J. Brabec, “Design Rules for Donors in Bulk-Heterojunction Solar Cells - Towards 10% Energy-Conversion Efficiency,” Adv. Mater. (Deerfield Beach Fla.) 18(6), 789–794 (2006). [CrossRef]  

9. J. Y. Kim, K. Lee, N. E. Coates, D. Moses, T.-Q. Nguyen, M. Dante, and A. J. Heeger, “Efficient tandem polymer solar cells fabricated by all-solution processing,” Science 317(5835), 222–225 (2007). [CrossRef]   [PubMed]  

10. N. S. Sariciftci, L. Smilowitz, A. J. Heeger, and F. Wudl, “Photoinduced electron transfer from a conducting polymer to buckminsterfullerene,” Science 258(5087), 1474–1476 (1992). [CrossRef]   [PubMed]  

11. G. Yu, J. Gao, J. C. Hummelen, F. Wudl, and A. J. Heeger, “Polymer Photovoltaic Cells: Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions,” Science 270(5243), 1789–1791 (1995). [CrossRef]  

12. H. Hoppe and N. S. Sariciftci, “Organic solar cells: an overview,” J. Mater. Res. 19(7), 1924–1945 (2004). [CrossRef]  

13. D. E. Markov, C. Tanase, P. W. M. Blom, and J. Wildeman, “Simultaneous enhancement of charge transport and exciton diffusion in poly(p-phenylene vinylene) derivatives,” Phys. Rev. B 72(4), 045217 (2005). [CrossRef]  

14. P. E. Shaw, A. Ruseckas, and I. D. W. Samuel, “Exciton Diffusion Measurements in Poly(3-hexylthiophene,” Adv. Mater. (Deerfield Beach Fla.) 20(18), 3516–3520 (2008). [CrossRef]  

15. S. Sista, M.-H. Park, Z. Hong, Y. Wu, J. Hou, W. L. Kwan, G. Li, and Y. Yang, “Highly efficient tandem polymer photovoltaic cells,” Adv. Mater. (Deerfield Beach Fla.) 22(3), 380–383 (2010). [CrossRef]  

16. S. H. Park, A. Roy, S. Beaupre, S. Cho, N. Coates, J. S. Moon, D. Moses, M. Leclerc, K. Lee, and A. J. Heeger, “Bulk heterojunction solar cells with internal quantum efficiency approaching 100%,” Nat. Photonics 3(5), 297–302 (2009). [CrossRef]  

17. M.-H. Chen, J. Hou, Z. Hong, G. Yang, S. Sista, L.-M. Chen, and Y. Yang, “Efficient Polymer Solar Cells with Thin Active Layers Based on Alternating Polyfluorene Copolymer/Fullerene Bulk Heterojunctions,” Adv. Mater. (Deerfield Beach Fla.) 21(42), 4238–4242 (2009). [CrossRef]  

18. G. Li, V. Shrotriya, Y. Yao, and Y. Yang, “Investigation of annealing effects and film thickness dependence of polymer solar cells based on poly(3-hexylthiophene),” J. Appl. Phys. 98(4), 043704 (2005). [CrossRef]  

19. F. Monestier, J. Simon, P. Torchio, L. Escoubas, F. Flory, S. Bailly, R. Debettignies, S. Guillerez, and C. Defranoux, “Modeling the short-circuit current density of polymer solar cells based on P3HT:PCBM blend,” Sol. Energy Mater. Sol. Cells 91(5), 405–410 (2007). [CrossRef]  

20. M. A. Green, Solar Cells: Operating Principles, Technology and System Applications (Univ. New South Wales, 1998).

21. V. Y. Yerokhov, R. Hezel, M. Lipinski, R. Ciach, H. Nagel, A. Mylyanych, and P. Panek, “Cost-effective methods of texturing for silicon solar cells,” Sol. Energy Mater. Sol. Cells 72(1-4), 291–298 (2002). [CrossRef]  

22. H. Sai, H. Fujiwara, and M. Kondo, “Back surface reflectors with periodic textures fabricated by self-ordering process for light trapping in thin-film microcrystalline silicon solar cells,” Sol. Energy Mater. Sol. Cells 93(6-7), 1087–1090 (2009). [CrossRef]  

23. H. A. Atwater and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9(3), 205–213 (2010). [CrossRef]   [PubMed]  

24. H. Raether, Surface Plasmons on Smooth and Rough Surfaces and on Gratings (Springer: Berlin, 1988).

25. E. Ozbay, “Plasmonics: merging photonics and electronics at nanoscale dimensions,” Science 311(5758), 189–193 (2006). [CrossRef]   [PubMed]  

26. V. M. Shalaev, and S. Kawata, Nanophotonics with surface plasmons (Elsevier, 2007).

27. M. Westphalen, U. Kreibig, J. Rostalski, H. Luth, and D. Meissner, “Metal cluster enhanced organic solar cells,” Sol. Energy Mater. Sol. Cells 61(1), 97–105 (2000). [CrossRef]  

28. B. P. Rand, P. Peumans, and S. R. Forrest, “Long-range absorption enhancement in organic tandem thin-film solar cells containing silver nanoclusters,” J. Appl. Phys. 96(12), 7519 (2004). [CrossRef]  

29. K. Tvingstedt, N.-K. Persson, O. Inganäs, A. Rahachou, and I. V. Zozoulenko, “Surface plasmon increase absorption in polymer photovoltaic cells,” Appl. Phys. Lett. 91(11), 113514 (2007). [CrossRef]  

30. A. J. Morfa, K. L. Rowlen, T. H. Reilly III, M. J. Romero, and J. van de Lagemaat, “Plasmon-enhanced solar energy conversion in organic bulk heterojunction photovoltaics,” Appl. Phys. Lett. 92(1), 013504 (2008). [CrossRef]  

31. D. S. Derkacs, H. Lim, P. Matheu, W. Mar, and E. T. Yu, “Improved performance of amorphous silicon solar cells via scattering from surface plasmon polaritons in nearby metallic nanoparticles,” Appl. Phys. Lett. 89(9), 093103 (2006). [CrossRef]  

32. S. Pillai, K. R. Catchpole, T. Trupke, and M. A. Green, “Surface plasmon enhanced silicon solar cells,” J. Appl. Phys. 101(9), 093105 (2007). [CrossRef]  

33. K. R. Catchpole and A. Polman, “Plasmonic solar cells,” Opt. Express 16(26), 21793–21800 (2008). [CrossRef]   [PubMed]  

34. K. Nakayama, K. Tanabe, and H. A. Atwater, “Plasmonic nanoparticle enhanced light absorption in GaAs solar cells,” Appl. Phys. Lett. 93(12), 121904 (2008). [CrossRef]  

35. C. Hägglund, M. Zäch, G. Petersson, and B. Kasemo, “Electromagnetic coupling of light into a silicon solar cell by nanodisk plasmons,” Appl. Phys. Lett. 92(5), 053110 (2008). [CrossRef]  

36. 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). [CrossRef]  

37. R. A. Pala, J. White, E. Barnard, J. Liu, and M. L. Brongersma, “Design of Plasmonic Thin-Film Solar Cells with Broadband Absorption Enhancements,” Adv. Mater. (Deerfield Beach Fla.) 21(34), 3504–3509 (2009). [CrossRef]  

38. W. Bai, Q. Gan, F. Bartoli, J. Zhang, L. Cai, Y. Huang, and G. Song, “Design of plasmonic back structures for efficiency enhancement of thin-film amorphous Si solar cells,” Opt. Lett. 34(23), 3725–3727 (2009). [CrossRef]   [PubMed]  

39. C. Min, J. Li, G. Veronis, J.-Y. Lee, S. Fan, and P. Peumans, “Enhancement of optical absorption in thin-film organic solar cells through the excitation of plasmonic modes in metallic gratings,” Appl. Phys. Lett. 96(13), 133302 (2010). [CrossRef]  

40. 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). [CrossRef]   [PubMed]  

41. G. Dennler, K. Forberich, T. Ameri, C. Waldauf, P. Denk, C. J. Brabec, K. Hingerl, and A. J. Heeger, “Design of efficient organic tandem cells: On the interplay between molecular absorption and layer sequence,” J. Appl. Phys. 102(12), 123109 (2007). [CrossRef]  

42. H. Hoppe, N. S. Sariciftci, and D. Meissner, “Optical Constants of Conjugated Polymer/Fullerene Based Bulk-Heterojunction Organic Solar Cells,” Mol. Cryst. Liq. Cryst. (Phila. Pa.) 385(1), 113 (2002). [CrossRef]  

43. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press: Orlando, FL, 1985).

44. J. Braun, B. Gompf, G. Kobiela, and M. Dressel, “How holes can obscure the view: suppressed transmission through an ultrathin metal film by a subwavelength hole array,” Phys. Rev. Lett. 103(20), 203901 (2009). [CrossRef]  

45. J. J. Burke, G. I. Stegeman, and T. Tamir, “Surface-polariton-like waves guided by thin, lossy metal films,” Phys. Rev. B Condens. Matter 33(8), 5186–5201 (1986). [CrossRef]   [PubMed]  

46. Z. Chen, I. R. Hooper, and J. R. Sambles, “Strongly coupled surface plasmons on thin shallow metallic gratings,” Phys. Rev. B 77(16), 161405 (2008). [CrossRef]  

47. A. V. Zayats, I. I. Smolyaninov, and A. A. Maradudin, “Nano-optics of surface plasmon polaritons,” Phys. Rep. 408(3-4), 131–314 (2005). [CrossRef]  

48. C. Genet and T. W. Ebbesen, “Light in tiny holes,” Nature 445(7123), 39–46 (2007). [CrossRef]   [PubMed]  

49. F. Yang, J. R. Sambles, and G. W. Bradberry, “Long-range surface modes supported by thin film,” Phys. Rev. B 44(11), 5855–5872 (1991). [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 (8)

Fig. 1
Fig. 1 A schematic diagram of the proposed plasmonic organic solar cell.
Fig. 6
Fig. 6 Comparison between the simulated and measured absorbance [see the red dots, reproduced from Fig. 2(b) in Ref. [9]. of a 150nm thick P3HT:PC70BM layer.
Fig. 2
Fig. 2 Calculations on the device with the nanopatterned metallic structure of diameter D = 150nm, and period P = 300nm. (a) Solid line - the absorption spectrum with a nanopatterned metallic structure; dotted line - the absorption spectrum with a flat metallic structure; and dashed line - the absorption enhancement spectrum. (b) and (c) are time averaged magnetic intensity (|Hy|2) distributions at the SPP resonance wavelength. (b) is the intensity distribution in the x-y plane, and (c) is the intensity distribution in x-z plane. The spatial mode profile is plotted in the right panel of (c).
Fig. 3
Fig. 3 (a) and (b) are electric field distributions at the SRSPP resonance wavelength at 725nm. (a) Time-averaged (color-scale) and instantaneous (arrows) electric field strengths and surface charge distribution in x-z plane, and (b) Instantaneous EZ vector distribution at the top and bottom surfaces of the Ag nanostructure. (c) Map of the absorption enhancement versus wavelength and metallic nanostructure thickness. The solid arrow corresponds to the resonance wavelength of the single-interface SPP Bloch mode. The dashed line corresponds to the analytical solutions of the SRSPP Bloch modes.
Fig. 7
Fig. 7 The photon flux density of the solar spectrum.
Fig. 4
Fig. 4 Maps of the absorption enhancement versus wavelength and period for two organic solar cells, i.e. a P3HT:PC70BM cell (a), and a PCPDTBT:PCBM cell (b). The dashed lines correspond to the analytical solutions of the SRSPP Bloch modes.
Fig. 5
Fig. 5 (a) The short-circuit current density (JSC) and its corresponding enhancement versus period of the nanopatterned metallic structure for the P3HT:PC70BM cell. The inset shows the absorbed photons spectra of the device with a nanopatterned metallic structure (solid line, P = 260nm) and a flat metallic surface (dotted line). (b) JSC and its corresponding enhancement versus period of the nanopatterned structure for the PCPDTBT:PCBM cell. The inset shows the absorbed photons spectra of the device with nanopatterned metallic structure (solid line, P = 320nm) and flat metallic surface (dotted line).
Fig. 8
Fig. 8 The relation between the JSC and the active layer thickness of the P3HT:PC70BM device. The red curve is the JSC of the device with nanostructured back reflector; the black curve is the JSC of the device with a 20nm flat back reflector. Inset: The relation between the JSC enhancement factor (the ratio of the JSC-nano and JSC-ref ) and the active layer thickness.

Equations (3)

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

k i n p l a n e = G i j = i G x + j G y .
| k S P P | = ω c ε d ε m ε d + ε m ,
tanh ( S 2 t ) ( ε d 1 ε d 2 S 2 2 + ε m 2 S 1 S 3 ) + [ ε m S 2 ( ε d 1 S 3 + ε d 2 S 1 ) ] = 0.
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.