• 沒有找到結果。

Surface Plasmonic Effects of Metallic Nanoparticles on the Performance of Polymer Bulk Heterojunction Solar Cells

N/A
N/A
Protected

Academic year: 2021

Share "Surface Plasmonic Effects of Metallic Nanoparticles on the Performance of Polymer Bulk Heterojunction Solar Cells"

Copied!
9
0
0

加載中.... (立即查看全文)

全文

(1)

January 13, 2011

C 2011 American Chemical Society

Surface Plasmonic Effects of Metallic

Nanoparticles on the Performance

of Polymer Bulk Heterojunction

Solar Cells

Jyh-Lih Wu,†Fang-Chung Chen,‡,* Yu-Sheng Hsiao,§Fan-Ching Chien,§Peilin Chen,§Chun-Hong Kuo,^ Michael H. Huang,^and Chain-Shu Hsu)

Department of Photonics and Institute of Electro-optical Engineering, National Chiao Tung University, Hsinchu 30010, Taiwan,Department of Photonics and

Display Institute, National Chiao Tung University, Hsinchu 30010, Taiwan,§Research Center for Applied Sciences, Academia Sinica, Taipei 11529, Taiwan, ^Department of Chemistry, National Tsing Hua University, Hsinchu 3001, Taiwan, and )Department of Applied Chemistry, National Chiao Tung University,

Hsinchu 30010, Taiwan

O

rganic photovoltaic devices (OPVs) are promising alternative tools for harnessing renewable energy be-cause of their properties including light-weight, low-cost, low-temperature fabrica-tion, semi-transparency, and mechanical flexibility. To date, the most representative high-performance OPVs reported in the literature have been fabricated using the bulk heterojunction (BHJ) concept,1-6 where a light-absorbing polymer (donor) and a solu-ble fullerene (acceptor) form a three-dimen-sional matrix possessing a large-area phase-separated interface for efficient exciton dissociation.7,8The resultant power conver-sion efficiencies (PCEs) of these polymer/ful-lerene BHJ OPV devices have reached as high as 6-8%,1-4opening up the possibility for

their practical use asflexible, low-cost, renew-able energy systems. As a result, increasing effort is being exerted in the quest for higher-performance OPVs.

The overall efficiency of an OPV device is governed by its internal quantum efficiency and absorption efficiency.9,10 The former, which is influenced by the diffusion and dissociation of photogenerated excitons and charge collection at the electrode con-tacts, can approach 100%.3 Therefore, ab-sorption efficiency of incoming light in these devices remains one of the major limitations toward realizing high external quantum efficiencies and PCEs. Typically, the optimum thickness of the active layer for an OPV device is on the order of 100-200 nm, or possibly less; such a thin layer can lead to low absorption of light. Therefore, increasing the thickness of the active layer is one possible approach toward

more efficient light absorption. A thicker layer, however, inevitably increases the de-vice resistance, due to the low carrier mo-bilities and short exciton diffusion lengths of organic materials.11,12This situation

im-poses a trade-off between light absorption and charge transport efficiencies in OPV devices, motivating the development of a variety of light-trapping techniques.

Efficient light trapping in the active layer enhances photon absorption without the need for a thickfilm. The introduction of an optical spacer that spatially redistributes the light intensity in the device can result in enhanced light absorption.13-16Changes in

*Address correspondence to fcchen@mail.nctu.edu.tw.

Received for review September 6, 2010 and accepted January 7, 2011. Published online

10.1021/nn102295p

ABSTRACT We have systematically explored how plasmonic effects influence the characteristics of polymer photovoltaic devices (OPVs) incorporating a blend of poly(3-hexylthiophene) (P3HT) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM). We blended gold nanoparticles (Au NPs) into the

anodic buffer layer to trigger localized surface plasmon resonance (LSPR), which enhanced the performance of the OPVs without dramatically sacrificing their electrical properties. Steady state photoluminescence (PL) measurements revealed a significant increase in fluorescence intensity, which we attribute to the increased light absorption in P3HT induced by the LSPR. As a result, the rate of generation of excitons was enhanced significantly. Furthermore, dynamic PL measurements revealed that the LSPR notably reduced the lifetime of photogenerated excitons in the active blend, suggesting that interplay between the surface plasmons and excitons facilitated the charge transfer process. This phenomenon reduced the recombination level of geminate excitons and, thereby, increased the probability of exciton dissociation. Accordingly, both the photocurrents andfill factors of the OPV devices were enhanced significantly. The primary origin of this improved performance was local enhancement of the electromagneticfield surrounding the Au NPs. The power conversion efficiency of the OPV device incorporating the Au NPs improved to 4.24% from a value of 3.57% for the device fabricated without Au NPs.

KEYWORDS: polymer photovoltaics • gold nanoparticles • surface plasmon• photoluminescence • exciton lifetime

(2)

WU ET AL. VOL. 5 NO. 2 959–967 2011 960 the device symmetry and the implementation of

peri-odic nanostructures can also increase the optical path length in the active layer.17-20Moreover, microcavity and photonic crystal effects have been proposed to effectively trap light in OPV devices.21-23

Recently, the exploitation of surface plasmon reso-nance (SPR) effects, based on advantageous optical properties such as light concentration and/or scatter-ing, has attracted much attention as a means for increasing the photocurrents of OPVs.24-32 Two pri-mary schemes are commonly employed for excitation of the SPR, which is the coherent collective oscillation of conduction electrons surrounding the metallic sur-faces. In one, surface plasmon polaritons (SPPs) propa-gating along the metal-dielectric interface are triggered by incorporating metallic nanostructures, such as periodic arrays or gratings.24-26In the other, surface plasmons are localized by noble metallic nanoparticles (NPs);such as Cu, Ag, Pt, and Au; resulting in localized surface plasmon resonance (LSPR).27-32The excitation of LSPR can be achieved

when the frequency of the incident light matches its resonance peak, resulting in unique optical proper-ties;selective light extinction as well as local en-hancement of electromagneticfield near the surface of metallic NPs. The resonance peak of LSPR depends strongly on the size, shape, and the dielectric environ-ment of the metallic NPs. Although both approaches have been employed to increase the photocurrents of OPV devices, a comprehensive understanding of such plasmonic effects in OPVs remains rare. In this study, we systematically explored how plasmonic effects influence the device characteristics of OPVs incorpor-ating a blend of poly(3-hexylthiophene) (P3HT) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM).

Ex-citation of the LSPR;triggered by adding Au NPs into the anodic buffer layer;significantly enhanced the overall PCEs of the OPV devices. Electrical characteriza-tion results revealed that the presence of the metallic NPs had a negligible effect on the charge transport process, suggesting that the electrical properties were not sacrificed. Moreover, steady state and dyna-mic photoluminescence (PL) measurements provided strong evidence that the LSPR induced by the Au NPs not only increased the degree of light absorption but also enhanced the degree of exciton dissociation. As a result, the photocurrent and overall device efficiency were both improved considerably after exploiting the optical effects of the LSPR. Compared with previous approaches toward plasmonic-enhanced OPV devices, which involved the nanofabrication of plasmoic nanostructures,24-26the method we report herein is simpler (solution processable) and also suitable for the applications to the large-area processes.

RESULTS AND DISCUSSION

Photovoltaic Characteristics. We used solution-proces-sable Au NPs to trigger the LSPR; the average particle size was ca. 45( 5 nm, estimated from a scanning electron microscopy (SEM) image (inset of Figure 1). The extinction spectrum of the Au NPs, determined using UV-vis spectroscopy, is displayed in Figure 1. The resonance of the Au NPs in solution was located at ca. 550 nm. The plasmonic resonance regime of the Au NPs was close to the absorption peak of the P3HT/PCBM blends, thereby suggesting enhanced light-harvesting efficiency. Figure 2a displays the architecture of our plasmonic-enhanced OPV devices. The presence of Au NPs embedded in the anodic buffer layer was evident from the SEM images recorded for the modified (Figure 2c) and pristine (Figure 2b) poly(3,4-ethylene-dioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) layers. Notably, the distribution of the Au NPs (white dots) in the modified PEDOT:PSS layer was uniform; we observed no apparent aggregation between the Au NPs in the film prepared using this simple method.

Figure 3a displays the representative current density -voltage (J-V) characteristics, recorded under

Figure 1. UV-vis extinction spectrum of the solution-processable Au NPs. Inset: Corresponding SEM image. Average Au NP size =ca. 45 ( 5 nm.

Figure 2. (a) Device architecture of the OPV incorporating Au NPs in the PEDOT:PSS layer. (b,c) SEM images of anodic buffer layers prepared with (b) pristine PEDOT:PSS and (c) PEDOT:PSS featuring embedded Au NPs. The Au NPs appear as white dots in (c).

(3)

100 mW cm-2illumination (AM 1.5G), of OPV devices prepared using buffer layers with and without Au NPs. The reference device possessing the structure ITO/ PEDOT:PSS/P3HT:PCBM/Ca/Al exhibited character-istics comparable to those of previously reported devices,20,29with an open-circuit voltage (Voc) of 0.59 V,

a short-circuit current (Jsc) of 9.16 mA cm-2, and afill

factor (FF) of 66.06%, resulting in a PCE of 3.57%. To take advantage of the unique optical properties of the LSPR, we blended the Au NPs into the anodic buffer layer (denoted herein as the “plasmonic device”). The value of Vocfor the plasmonic device remained

at 0.59 V, suggesting the unchanged nature of the electrode-organics interface. In contrast, the values of Jsc and FF both increased, to 10.22 mA cm-2 and

70.32%, respectively. Notably, the device resistance, extracted from the J-V curves in the dark, increased

slightly from 1.79 to 1.82Ω 3 cm2after incorporating the Au NPs. This potential drawback was, however, overwhelmed by the advantageous plasmonic effects. Overall, the plasmonic device's PCE was 4.24%.

Figure 3b presents incident photon-to-electron conversion efficiency (IPCE) curves for these devices. We also compared the curve of the increase in IPCE (ΔIPCE) after incorporating Au NPs with the extinction spectrum of the Au NPs (Figure 3c). The photocurrent within the wavelength range from 450 to 650 nm increased significantly after incorporating the Au NPs. This wavelength regime coincides with the extinction range of the Au NPs, indicating that LSPR effects did indeed improve the photocurrent.

Photocurrent Behavior. To further explore the effects of LSPR in OPV devices, we determined the maximum exciton generation rate (Gmax). The devices were

biased sweeping fromþ1 to -10 V. Figure 4a reveals the dependence of the photocurrent density (Jph) on the

effective voltage (Veff), recorded under illumination at 100

mW cm-2, for our reference and plasmonic devices. Here Jph ¼ JL- JD

where JLand JDare the current densities under

illumina-tion and in the dark, respectively, and Veff ¼ Vo- Va

where Vois the voltage when Jphequals zero (i.e., JL=

JD) and Va is the applied voltage. Apparently, Jph

Figure 3. (a)J-V characteristics, recorded under illumina-tion at 100 mW cm-2(AM 1.5G), of polymer solar cells with (plasmonic device) and without (reference device) Au NPs in the PEDOT:PSS layer. (b) Corresponding IPCE curves of these OPV devices. (c) Comparison between the curve of the increase in IPCE (ΔIPCE) after incorporating Au NPs and the extinction spectrum of the Au NPs.

Figure 4. (a) Photocurrent density (Jph) plotted with respect

to effective bias (Veff) for the reference and plasmonic

devices. (b) Exciton dissociation probability [P(E,T)] plotted with respect to effective bias (Veff) for these OPV devices.

The dissociation probabilities for the reference and plas-monic devices under short-circuit conditions were 78.2 and 84.2%, respectively.

(4)

WU ET AL. VOL. 5 NO. 2 959–967 2011 962 linearly increased with the voltage at a low value of Veff

and then saturated at a sufficiently high value of Veff.

From Figure 4a, we therefore determined the values of the saturation photocurrent density (Jsat), which is

independent of the bias and temperature. Further-more, assuming that all of the photogenerated exci-tons dissociated and contributed to the current in the saturated regime due to the sufficiently high electric field, we obtained the values of Gmaxusing the

equation

Jsat ¼ qGmaxL

where q is the electronic charge and L is the thickness of the active layer.5,33The values of Gmaxfor the reference

and plasmonic devices were 3.96 1027m-3s-1(Jsat=

114 A m-2) and 4.27 1027m-3s-1(Jsat= 123 A m-2),

respectively. Thus, a noticeable enhancement in Gmax

occurred after incorporating the Au NPs into the device. Because the value of Gmaxis a measure of the maximum

number of photons absorbed,5,33such an increase

sug-gests enhanced light absorption in the active layer of the plasmonic device.

Next, we compared the exciton dissociation prob-abilities [P(E,T)], which are related to the electricfield (E) and temperature (T), for our devices. For OPVs, when the excitons are photogenerated, only a portion of them can be dissociated into free carriers. As a result, Jphcan be expressed using the equation5,33

Jph ¼ qGmaxP(E, T)L

As a result, the value of P(E,T) at any bias can be obtained from the plot of the normalized photocurrent density (Jph/Jsat) with respect to Veff.34Figure 4b reveals

that the value of P(E,T) under the short-circuit condi-tions (Va= 0 V) increased from 79.2% for the reference

device to 84.4% for the plasmonic device, indicating that excitation of the LSPR also facilitated excitons to dissociate into free carriers. Thus, excitation of the LSPR increased both the exciton generation rate and the dissociation probability, thereby enhancing the photo-current of the OPVs.

Charge Transport Behavior. We measured the values of Jscof our OPV devices at various illumination intensities

to examine how the plasmonic effects influenced the charge transport process. Figure 5 displays the depen-dence of the incident light intensity (Plight) on the

photocurrent. In general, the value of Jscof an OPV

device follows a power-law dependence with respect to Plight(i.e., Jsc Plights).5,10From Figure 5, we

deter-mined exponential factors (s) for the reference and plasmonic devices of 0.984 and 0.980, respectively. The nearly linear dependence of Jsc on the incident light

intensity suggests the absence of bimolecular recom-bination and space-limited charges in either of these devices.5,10Therefore, we suspect that the

incorpora-tion of Au NPs into the buffer layer had a negligible influence on the nature of the charge transport process in the device. More importantly, however, the result-ing plasmonic effects resulted in unique optical prop-erties that improved the performance of the OPV devices.

Steady State Photoluminescence. To explore the effect of the LSPR on the exciton generation behavior, we performed steady state PL measurements. Because absorption in the donor phase (i.e., P3HT) contributed predominately to the exciton generation in the active blend, we recorded fluorescence spectra of the pristine P3HT films. The samples were prepared by spin-coat-ing pristine P3HT films onto the PEDOT:PSS buffer layers in the presence and absence of the Au NPs. Figure 6 presents the room temperature PL spec-tra obtained using different wavelength excitation sources (λexc = 470 and 532 nm) for the reference

and plasmonic samples. The integrating PL intensity of the plasmonic sample was enhanced by ca. 12 and 20% with respect to the reference sample when we excited the sample at 470 and 532 nm, respectively; these excitation wavelengths were selected intention-ally to be located within the resonance regime of the LSPR (see Figure 1). The fluorescence process is de-pendent on the light excitation rate and the quantum yield, which is dependent on competition between the radiative and nonradiative decaying mechanisms.35 Because the resonance frequency of the Au NPs was close to the absorption band of P3HT, we attribute the enhanced PL intensity to the fact that excitation of the LSPR increased the degree of light absorption and, thereby, enhanced the light excitation rate. This hy-pothesis is consistent with our measured device char-acteristics;namely, the greater value of Gmaxfor the

plasmonic device relative to that of the reference device. Consequently, the PL measurements sug-gested that, upon excitation of the LSPR, the enhanced excitation rate increased the density of photogener-ated excitons in P3HT, thereby enhancing the fluores-cence intensity.36,37Notably, another possible factor responsible for the enhanced PL intensity was related to the modification of excitonic states of P3HT due to

Figure 5. Values ofJscof the OPVs plotted with respect to

the light intensity. Straight lines werefitted using the expressionJsc Plights; the values ofs for the reference and

plasmonic devices were 0.984 and 0.980, respectively.

(5)

the presence of Au NPs.31This therefore accounted for the occurrence of the strong interactions between plasmonic field and excitons.

Our observation that LSPR enhanced the light absorption efficiency might result from two possible factors. First, the forward scattering lengthened the optical path in the active layer, thereby enhancing the degree of light absorption, as illustrated schematically in Figure 7a. Alternatively, excitation of the LSPR resulted in local enhancement of the electromagnetic field in the vicinity of the Au NPs (see Figure 7b); enhancement factors of up to 100 can be achieved.38,39 Therefore, the resulting local field enhancement in-creased the total number of excitons created in the active layer because the energy dissipation is propor-tional to the intensity of the electromagneticfield.40,41 For plasmonic metallic NPs, however, the particle size

plays an important role in dominating the extinction behavior (e.g., the sum of light absorption and scattering). Light scattering, which is beneficial for collecting light in inorganic solar cells,42,43becomes more dominant when using larger NPs. Typically, the light scattering efficiency is lower than the light ab-sorption efficiency for Au NPs having particle sizes of less than 100 nm.42,44Because the average particle size of the Au NPs used in this study was only 45 nm, they were presumably unable to induce efficient light scat-tering. Therefore, we conclude that the LSPR effects, induced by the presence of the Au NPs, enhanced light collection in the active layer of OPVs predominately through localfield enhancement.

Dynamic Photoluminescence. We employed time-resolved PL spectroscopy to acquire insight into the coupling process between the plasmonic field and the photogenerated excitons within the photoactive blend. Through measurements of exciton lifetimes (τexciton), the dynamic PL signal allowed us to probe

how the plasmonic effects influenced the photophysi-cal processes between the donor and acceptor phases.45-47Figure 8a displays the PL intensity decay profiles, obtained using a 470 nm excitation source, of P3HT/PCBM films that had been deposited on PEDOT:PSS layers in the presence and absence of Au NPs. The PL decay data were fitted using the following

Figure 6. PL spectra of the reference and plasmonic samples recorded using excitation source wavelengths (λexc) of (a) 470 and (b) 532 nm.

Figure 7. (a) Schematic representation of light trapping through forward scattering as a result of Au NP induced LSPR. The optical path length increased because light was trapped through multiple and high-angle scattering. (b) Schematic representation of the local enhancement of the electromagneticfield. The plasmonic field decays exponentially with respect to the distance from the surface of the Au NPs.

Figure 8. (a) PL decay profiles for the P3HT/PCBM blends in the reference and plasmonic samples. Excitation source: 470 nm pulsed laser. Inset: Schematic representation of the charge transfer process in the P3HT/PCBM blend. For the reference and plasmonic samples, the values ofτexcitonwere

0.36 and 0.27 ns, respectively. (b) Cartoon depiction of the interplay between the LSPR and the excitons. The resulting interactions enhanced the rate of exciton dissociation, thereby reducing exciton recombination.

(6)

WU ET AL. VOL. 5 NO. 2 959–967 2011 964 multiexponential function:48 IPL(t) ¼ Xn i¼ 1 Aiexp -t τi   (1) where Ai is the amplitude of the ith decay, n is the

number of decays involved, andτiis the ith exponential

constant. For the reference sample, we used two exponential constants (τ1andτ2) to obtain a good fit

to the data; their values of 0.200 and 0.558 ns, respec-tively, provided a corresponding value ofτexcitonof 0.36 ns.

For the plasmonic sample, however, we required a third exponential constant (τ3) to obtain a better fit; the

values of these three exponential constants were 0.131, 0.324, and 0.015 ns, respectively, providing a resulting lower value ofτexcitonof 0.27 ns. Note that the

introduction ofτ3was necessary presumably because

of the presence of the LSPR, which interacted with photogenerated excitons.

The dramatic change in the value ofτexcitonafter

incorporating Au NPs also accounted for the pre-sence of the strong coupling between the plasmo-nic field and excitons, which has been reported previously.31,49,50 We speculated that the resulting plasmon-exciton coupling participated in the charge transfer process, thus facilitating exciton dissociation, as illustrated in Figure 8b. The photophysical process can be further interpreted by the concept of “hot excitons” which possess excess energy to overcome their initial Coulombic potential.31,49,50Therefore, the generation of hot excitons is beneficial for enhancing the probability of dissociation into free polarons.31,49,50 Previous reports have shown that the plasmonicfield

can strongly modify the dynamic properties of photo-generated excitons through the plasmon-exciton coupling.51,52This phenomenon presumably increased the amount of hot excitons for efficient dissociation.31 As a result, the enhanced degree of exciton dissocia-tion reduced the recombinadissocia-tion level of geminate excitons through radiative and/or nonradiative pro-cesses and, thereby, improved the FFs of OPVs.53,54 Note that the device resistance for the plasmonic device increased slightly relative to that of the refer-ence device; therefore, it was not responsible for the enhanced FF. Consequently, we assigned the in-creased FF observed in Figure 3a primarily to the increased value of P(E,T) resulting from the interplay between the plasmonicfield and the photogenerated excitons.

We performed exciton lifetime mapping by com-bining the time-resolved PL measurement system with confocal laser scanning microscopy to further investi-gate the plasmonic effects.45,47This technique enabled us to examine the distribution of the exciton lifetimes on any given z-axial planes (close to or away from the PEDOT:PSS layer, in which the Au NPs were embedd-ed). As a result, we could obtain the vertical evolution ofτexcitonin the P3HT/PCBMfilms. Figure 9 displays the

exciton lifetime images (15 μm  15 μm) for the reference and plasmonic samples (P3HT/PCBMfilms). The values ofτexciton, both close to or away from the

PEDOT:PSS layer, generally decreased after incorporat-ing the Au NPs. To quantify the average values and distributions of the exciton lifetimes, we constructed the corresponding statistical histograms for these

Figure 9. Exciton lifetime images for the reference (a,c) and plasmonic (b,d) samples. (a,b) Positions away from the PEDOT:PSS layer; (c,d) positions close to the PEDOT:PSS layer.

(7)

samples. For the reference sample, the average value ofτexcitondecreased from 0.48 ns close to the PEDOT:

PSS layer (Figure 10c) to 0.35 ns away from the PEDOT: PSS layer (Figure 10a), presumably because the differ-ence between the surface energies of P3HT and PCBM resulted in an inhomogeneous distribution of PCBM molecules in the P3HT matrix when the blend under-went solvent annealing.45The average values ofτexciton

decreased significantly (to 0.27 and 0.23 in panels b and d in Figure 10, respectively) after incorporating the Au NPs; notably, the average value ofτexcitonat the

position close to the PEDOT:PSS layer was even lower than that away from the PEDOT:PSS layer. Therefore, for the plasmonic sample, we observed a reverse trend in the evolution ofτexciton. This result is not surprising

because the plasmonicfield decays exponentially with respect to the distance from the surface of the metallic NPs.55,56 Therefore, stronger interplay between the plasmonic field and excitons occurred at positions close to the PEDOT:PSS layer, due to the localized enhanced nearfield in the proximity of the Au NPs. Notably, the time-resolved PL measurement data, in which LSPR changed the lifetime of the excitons in the active blend, further support the field enhancement

mechanism that we proposed above because we would not expect the scattering scheme to influence the exciton lifetime significantly.

CONCLUSIONS

Efficient light absorption in thin film solar cells is critical for their high performance;especially for those based on organic materials, where the thickness of the active layer is relatively thin. In this study, we improved the device PCEs of OPVs by incorporating Au NPs in the PEDOT:PSS buffer layer. The degree of light absorption in the plasmonic-enhanced OPV device increased significantly as a result of LSPR-induced local field enhancement. Moreover, interactions between the plasmons and the photogenerated excitons resulted in an enhanced degree of exciton dissociation, thereby reducing the level of exciton loss through geminate recombination. Without sacrificing the electrical proper-ties, the incorporation of the Au NPs allowed the unique optical properties of the LSPR to improve the perfor-mance of the OPV. We believe that the results of this study might pave the way toward higher-efficiency OPV devices, and that this approach might also be applicable to systems featuring other types of active materials.

METHODS

Device Fabrication. The OPV device was fabricated on an indium tin oxide (ITO)-coated glass substrate. After a routine cleaning process, the substrate was dried overnight in an oven

and treated with UV ozone prior to use. To prepare the composite buffer layer, a Au NP solution was blended into the PEDOT:PSS (Baytron 4071) solution; the doping concentration of Au NPs inside the PEDOT:PSS solution was ca. 2 1011cm-3.

Figure 10. Histograms corresponding to the exciton lifetime images in Figure 9. Average values ofτexciton: (a) 0.35, (b) 0.27, (c)

0.48, and (d) 0.23 ns.

(8)

WU ET AL. VOL. 5 NO. 2 959–967 2011 966

The Au NP solution was prepared using procedures described previously.57The anodic buffer layer was deposited through spin-coating onto a ITO-coated glass substrate and then ther-mally annealed at 120C for 1 h. To reduce the surface rough-ness of the composite buffer layer, an additional thinner layer of PEDOT:PSS was spin-coated as a capping layer. For comparison, the thickness of each buffer layer was set at ca. 50 nm. In a nitrogen-filled glovebox, a solution of P3HT (Rieke Metals) and PCBM (Solenne) dissolved (each at 17 mg mL-1) in 1,2-dichloro-benzene (DCB) was spin-coated on top of the buffer layer. The wet film was subjected to solvent annealing in a glass Petri dish,6followed by a thermal annealing at 110C for 15 min. The

thickness of the resulting photoactive layer was ca. 180 nm. To complete the device, a cathode (30 nm Ca and 100 nm Al) was deposited through thermal evaporation under a vacuum of ca. 6 10-6Torr. The device area, defined through a shadow mask, was 0.12 cm2. The completed device was encapsulated with a cover glass and sealed with a UV-curing epoxy resin prior to device testing.

Characterization. The J-V characteristics of the devices were measured using a Keithley 2400 source measure unit. The photocurrent response was obtained under illumination from a 150 W Thermal Oriel solar simulator (AM 1.5G). The illumina-tion intensity was calibrated using a standard Si photodiode equipped with a KG-5 filter (Hamamatsu).58The IPCE

mea-surement system (Enli Technology) comprised a quartz-tungsten-halogen (QTH) lamp as the light source, a mono-chromator, an optical chopper, a lock-in amplifier, and a cali-brated silicon photodetector. The light intensity dependence study was conducted using neutral density filters to tune the light intensity. The dimensions of the Au NPs were measured using a JEOL JSM-7000F scanning electron microscope. Absorp-tion spectra were measured using an UV/vis/near-IR spectro-meter (PerkinElmer Lambda 950). The steady state PL spectra were recorded in the transmission mode under air ambient conditions; continuous-wave (CW) lasers (470 and 532 nm) were used as the excitation sources. The fluorescence emission signal was measured using a high-resolution spectrometer (HR4000, Ocean Optics). For the time-resolved PL measurements, the samples were excited using a 470 nm pulsed laser; the dynamic signal was recorded using a time-correlated single photon counting (TCSPC) spectrometer (LDH-P-C-470, Picoquant). For mapping of exciton lifetime images, the confocal laser scanning microscope (FV300, Olympus Corporation) was equipped with a single photon avalanche diode detector (PDM series, Picoquant).

Acknowledgment. We thank the National Science Council of Taiwan (NSC 98-3114-E-009-005 and NSC 99-2221-E-009-181) and the Ministry of Education of Taiwan (through the ATU program) forfinancial support.

REFERENCES AND NOTES

1. Chen, H. Y.; Hou, J. H.; Zhang, S. Q.; Liang, Y. Y.; Yang, G. W.; Yang, Y.; Yu, L. P.; Wu, Y.; Li, G. Polymer Solar Cells with Enhanced Open-Circuit Voltage and Efficiency. Nat. Photo-nics2009, 3, 649–653.

2. Hou, J. H.; Chen, H. Y.; Zhang, S. Q.; Chen, R. I.; Yang, Y.; Wu, Y.; Li, G. Synthesis of a Low Band Gap Polymer and Its Application in Highly Efficient Polymer Solar Cells. J. Am. Chem. Soc.2009, 131, 15586–15587.

3. Park, S. H.; Roy, A.; Beaupre, S.; Cho, S.; Coates, N.; Moon, J. S.; Moses, D.; Leclerc, M.; Lee, K.; Heeger, A. J. Bulk Heterojunction Solar Cells with Internal Quantum Effi-ciency Approaching 100%. Nat. Photonics2009, 3, 297– 303.

4. Liang, Y. Y.; Xu, Z.; Xia, J. B.; Tsai, S. T.; Wu, Y.; Li, G.; Ray, C.; Yu, L. P. For the Bright Future;Bulk Heterojunction Poly-mer Solar Cells with Power Conversion Efficiency of 7.4%. Adv. Mater.2010, 22, E135–E138.

5. Mihailetchi, V. D.; Xie, H. X.; de Boer, B.; Koster, L. J. A.; Blom, P. W. M. Charge Transport and Photocurrent Generation in Poly(3-hexylthiophene):Methanofullerene Bulk-Hetero-junction Solar Cells. Adv. Funct. Mater.2006, 16, 699–708.

6. Li, G.; Shrotriya, V.; Huang, J. S.; Yao, Y.; Moriarty, T.; Emery, K.; Yang, Y. High-Efficiency Solution Processable Polymer Photovoltaic Cells by Self-Organization of Polymer Blends. Nat. Mater.2005, 4, 864–868.

7. Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F. Photo-induced Electron-Transfer from a Conducting Polymer to Buckminsterfullerene. Science1992, 258, 1474–1476. 8. Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J.

Polymer Photovoltaic Cells. Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions. Science1995, 270, 1789–1791.

9. Jo, J.; Na, S. I.; Kim, S. S.; Lee, T. W.; Chung, Y.; Kang, S. J.; Vak, D.; Kim, D. Y. Three-Dimensional Bulk Heterojunction Morphology for Achieving High Internal Quantum Effi-ciency in Polymer Solar Cells. Adv. Funct. Mater.2009, 19, 2398–2406.

10. Schilinsky, P.; Waldauf, C.; Brabec, C. J. Recombination and Loss Analysis in Polythiophene Based Bulk Heterojunction Photodetectors. Appl. Phys. Lett.2002, 81, 3885–3887. 11. Shrotriya, V.; Wu, E. H. E.; Li, G.; Yao, Y.; Yang, Y. Efficient

Light Harvesting in Multiple-Device Stacked Structure for Polymer Solar Cells. Appl. Phys. Lett.2006, 88, 064104. 12. Yakimov, A.; Forrest, S. R. High Photovoltage

Multiple-Het-erojunction Organic Solar Cells Incorporating Interfacial Metallic Nanoclusters. Appl. Phys. Lett.2002, 80, 1667–1669. 13. Chen, F. C.; Wu, J. L.; Hung, Y. Spatial Redistribution of the Optical Field Intensity in Inverted Polymer Solar Cells. Appl. Phys. Lett.2010, 96, 193304.

14. Gilot, J.; Barbu, I.; Wienk, M. M.; Janssen, R. A. J. The Use of ZnO as Optical Spacer in Polymer Solar Cells: Theoretical and Experimental Study. Appl. Phys. Lett. 2007, 91, 113520.

15. Kim, J. Y.; Kim, S. H.; Lee, H. H.; Lee, K.; Ma, W. L.; Gong, X.; Heeger, A. J. New Architecture for High-Efficiency Polymer Photovoltaic Cells Using Solution-Based Titanium Oxide as an Optical Spacer. Adv. Mater.2006, 18, 572–576. 16. Roy, A.; Park, S. H.; Cowan, S.; Tong, M. H.; Cho, S. N.; Lee, K.;

Heeger, A. J. Titanium Suboxide as an Optical Spacer in Polymer Solar Cells. Appl. Phys. Lett.2009, 95, 013302. 17. Tvingstedt, K.; Andersson, V.; Zhang, F.; Ingan€as, O. Folded

Reflective Tandem Polymer Solar Cell Doubles Efficiency. Appl. Phys. Lett.2007, 91, 123514.

18. Zhou, Y. H.; Zhang, F. L.; Tvingstedt, K.; Tian, W. J.; Ingan€as, O. Multifolded Polymer Solar Cells on Flexible Substrates. Appl. Phys. Lett.2008, 93, 033302.

19. Cocoyer, C.; Rocha, L.; Sicot, L.; Geffroy, B.; de Bettignies, R.; Sentein, C.; Fiorini-Debuisschert, C.; Raimond, P. Imple-mentation of Submicrometric Periodic Surface Structures toward Improvement of Organic-Solar-Cell Performances. Appl. Phys. Lett.2006, 88, 133108.

20. Na, S. I.; Kim, S. S.; Jo, J.; Oh, S. H.; Kim, J.; Kim, D. Y. Efficient Polymer Solar Cells with Surface Relief Gratings Fabri-cated by Simple Soft Lithography. Adv. Funct. Mater. 2008, 18, 3956–3963.

21. Ko, D. H.; Tumbleston, J. R.; Zhang, L.; Williams, S.; DeSimone, J. M.; Lopez, R.; Samulski, E. T. Photonic Crystal Geometry for Organic Solar Cells. Nano Lett. 2009, 9, 2742–2746.

22. Tumbleston, J. R.; Ko, D. H.; Samulski, E. T.; Lopez, R. Absorption and Quasiguided Mode Analysis of Organic Solar Cells with Photonic Crystal Photoactive Layers. Opt. Express2009, 17, 7670–7681.

23. Long, Y. B. Improving Optical Performance of Inverted Organic Solar Cells by Microcavity Effect. Appl. Phys. Lett. 2009, 95, 193301.

24. Kang, M. G.; Xu, T.; Park, H. J.; Luo, X.; Guo, L. J. Efficiency Enhancement of Organic Solar Cells Using Transparent Plasmonic Ag Nanowire Electrodes. Adv. Mater.2010, 22, 4378–4383.

25. Lindquist, N. C.; Luhman, W. A.; Oh, S. H.; Holmes, R. J. Plasmonic Nanocavity Arrays for Enhanced Efficiency in Organic Photovoltaic Cells. Appl. Phys. Lett. 2008, 93, 123308.

26. Tvingstedt, K.; Persson, N. K.; Ingan€as, O.; Rahachou, A.; Zozoulenko, I. V. Surface Plasmon Increase Absorption in

(9)

Polymer Photovoltaic Cells. Appl. Phys. Lett.2007, 91, 123514.

27. Kim, S. S.; Na, S. I.; Jo, J.; Kim, D. Y.; Nah, Y. C. Plasmon Enhanced Performance of Organic Solar Cells Using Elec-trodeposited Ag Nanoparticles. Appl. Phys. Lett.2008, 93, 073307.

28. Chang, Y. C.; Chou, F. Y.; Yeh, P. H.; Chen, H. W.; Chang, S. H.; Lan, Y. C.; Guo, T. F.; Tsai, T. C.; Lee, C. T. Effects of Surface Plasmon Resonant Scattering on the Power Conversion Efficiency of Organic Thin-Film Solar Cells. J. Vac. Sci. Technol., B2007, 25, 1899–1902.

29. Chen, F. C.; Wu, J. L.; Lee, C. L.; Hong, Y.; Kuo, C. H.; Huang, M. H. Plasmonic-Enhanced Polymer Photovoltaic Devices Incorporating Solution-Processable Metal Nanoparticles. Appl. Phys. Lett.2009, 95, 013305.

30. Morfa, A. J.; Rowlen, K. L.; Reilly, T. H.; Romero, M. J.; van de Lagemaat, J. Plasmon-Enhanced Solar Energy Conversion in Organic Bulk Heterojunction Photovoltaics. Appl. Phys. Lett.2008, 92, 013504.

31. Lee, J. H.; Park, J. H.; Kim, J. S.; Lee, D. Y.; Cho, K. High Efficiency Polymer Solar Cells with Wet Deposited Plas-monic Gold Nanodots. Org. Electron.2009, 10, 416–420. 32. Kulkarni, A. P.; Noone, K. M.; Munechika, K.; Guyer, S. R.;

Ginger, D. S. Plasmon-Enhanced Charge Carrier Genera-tion in Organic Photovoltaic Films Using Silver Nano-prisms. Nano Lett.2010, 10, 1501–1505.

33. Mihailetchi, V. D.; Koster, L. J. A.; Hummelen, J. C.; Blom, P. W. M. Photocurrent Generation in Polymer-Fullerene Bulk Heterojunctions. Phys. Rev. Lett.2004, 93, 216601. 34. Shrotriya, V.; Yao, Y.; Li, G.; Yang, Y. Effect of

Self-Organiza-tion in Polymer/Fullerene Bulk HeterojuncSelf-Organiza-tions on Solar Cell Performance. Appl. Phys. Lett.2006, 89, 063505. 35. Anger, P.; Bharadwaj, P.; Novotny, L. Enhancement and

Quenching of Single-Molecule Fluorescence. Phys. Rev. Lett.2006, 96, 113002.

36. Mahmoud, M. A.; Poncheri, A. J.; Phillips, R. L.; El-Sayed, M. A. Plasmonic Field Enhancement of the Exci-ton-Exciton Annihilation Process in a Poly(p-pheny-leneethynylene) Fluorescent Polymer by Ag Nanocubes. J. Am. Chem. Soc.2010, 132, 2633–2641.

37. Konda, R. B.; Mundle, R.; Mustafa, H.; Bamiduro, O.; Pradhana, A. K.; Roy, U. N.; Cui, Y.; Burger, A. Surface Plasmon Excitation via Au Nanoparticles in n-CdSe/p-Si Heterojunction Diodes. Appl. Phys. Lett.2007, 91, 191111. 38. Maier, S. A.; Atwater, H. A. Plasmonics: Localization and Guiding of Electromagnetic Energy in Metal/Dielectric Structures. J. Appl. Phys.2005, 98, 011101.

39. Rand, B. P.; Peumans, P.; Forrest, S. R. Long-Range Absorp-tion Enhancement in Organic Tandem Thin-Film Solar Cells Containing Silver Nanoclusters. J. Appl. Phys.2004, 96, 7519–7526.

40. Monestier, F.; Simon, J. J.; Torchio, P.; Escoubas, L.; Florya, F.; Bailly, S.; de Bettignies, R.; Guillerez, S.; Defranoux, C. Modeling the Short-Circuit Current Density of Polymer Solar Cells Based on P3HT:PCBM Blend. Sol. Energy Mater. Sol. Cells2007, 91, 405–410.

41. Pettersson, L. A. A.; Roman, L. S.; Ingan€as, O. Modeling Photocurrent Action Spectra of Photovoltaic Devices Based on Organic Thin Films. J. Appl. Phys.1999, 86, 487–496. 42. Derkacs, D.; Lim, S. H.; Matheu, P.; Mar, W.; Yu, E. T.

Improved Performance of Amorphous Silicon Solar Cells via Scattering from Surface Plasmon Polaritons in Nearby Metallic Nanoparticles. Appl. Phys. Lett.2006, 89, 093103. 43. Derkacs, D.; Chen, W. V.; Matheu, P. M.; Lim, S. H.; Yu, P. K. L.; Yu, E. T. Nanoparticle-Induced Light Scattering for Im-proved Performance of Quantum-Well Solar Cells. Appl. Phys. Lett.2008, 93, 091107.

44. van Dijk, M. A.; Tchebotareva, A. L.; Orrit, M.; Lippitz, M.; Berciaud, S.; Lasne, D.; Cognet, L.; Lounis, B. Absorption and Scattering Microscopy of Single Metal Nanoparticles. Phys. Chem. Chem. Phys.2006, 8, 3486–3495.

45. Huang, J. H.; Chien, F. C.; Chen, P. L.; Ho, K. C.; Chu, C. W. Monitoring the 3D Nanostructures of Bulk Heterojunction Polymer Solar Cells Using Confocal Lifetime Imaging. Anal. Chem.2010, 82, 1669–1673.

46. Huang, J. H.; Yang, C. Y.; Ho, Z. Y.; Kekuda, D.; Wu, M. C.; Chien, F. C.; Chen, P. L.; Chu, C. W.; Ho, K. C. Annealing Effect of Polymer Bulk Heterojunction Solar Cells Based on Polyfluorene and Fullerene Blend. Org. Electron. 2009, 10, 27–33.

47. Huang, J. H.; Li, K. C.; Chien, F. C.; Hsiao, Y. S.; Kekuda, D.; Chen, P.; Lin, H. C.; Ho, K. C.; Chu, C. W. Corrrelation between Exciton Lifetime Distribution and Morphology of Bulk Heterojunction Films after Solvent Annealing

J. Phys. Chem. C2010, 114, 9062–9069.

48. Jones, M.; Nedeljkovic, J.; Ellingson, R. J.; Nozik, A. J.; Rumbles, G. Photoenhancement of Luminescence in Col-loidal CdSe Quantum Dot Solutions. J. Phys. Chem. B2003, 107, 11346–11352.

49. Zhu, X. Y.; Yang, Q.; Muntwiler, M. Charge-Transfer Ex-citons at Organic Semiconductor Surfaces and Interfaces. Acc. Chem. Res.2009, 42, 1779–1787.

50. Ohkita, H.; Cook, S.; Astuti, Y.; Duffy, W.; Tierney, S.; Zhang, W.; Heeney, M.; McCulloch, I.; Nelson, J.; Bradley, D. D. C.; et al. Charge Carrier Formation in Polythiophene/Full-erene Blend Films Studied by Transient Absorption Spec-troscopy. J. Am. Chem. Soc.2008, 130, 3030–3042. 51. Fofang, N. T.; Park, T. H.; Neumann, O.; Mirin, N. A.;

Nordlander, P.; Halas, N. J. Plexcitonic Nanoparticles: Plasmon-Exciton Coupling in Nanoshell-J-Aggregate Complexes. Nano Lett.2008, 8, 3481–3487.

52. Bellessa, J.; Bonnand, C.; Plenet, J. C. Strong Coupling between Surface Plasmons and Excitons in an Organic Semiconductor. Phys. Rev. Lett.2004, 93, 036404. 53. Moet, D. J. D.; Lenes, M.; Morana, M.; Azimi, H.; Brabec, C. J.;

Blom, P. W. M. Enhanced Dissociation of Charge-Transfer States in Narrow Band Gap Polymer:Fullerene Solar Cells Processed with 1,8-Octanedithiol. Appl. Phys. Lett.2010, 96, 213506.

54. Mandoc, M. M.; Veurman, W.; Koster, L. J. A.; de Boer, B.; Blom, P. W. M. Origin of the Reduced Fill Factor and Photocurrent in MDMO-PPV:PCNEPV All-Polymer Solar Cells. Adv. Funct. Mater.2007, 17, 2167–2173.

55. Lal, S.; Link, S.; Halas, N. J. Nano-Optics from Sensing to Waveguiding. Nat. Photonics2007, 1, 641–648. 56. Barnes, W. L.; Dereux, A.; Ebbesen, T. W. Surface Plasmon

Subwavelength Optics. Nature2003, 424, 824–830. 57. Chang, C. C.; Wu, H. L.; Kuo, C. H.; Huang, M. H.

Hydro-thermal Synthesis of Monodispersed Octahedral Gold Nanocrystals with Five Different Size Ranges and Their Self-Assembled Structures. Chem. Mater.2008, 20, 7570– 7574.

58. Shrotriya, V.; Li, G.; Yao, Y.; Moriarty, T.; Emery, K.; Yang, Y. Accurate Measurement and Characterization of Organic Solar Cells. Adv. Funct. Mater.2006, 16, 2016–2023.

數據

Figure 1. UV -vis extinction spectrum of the solution- solution-processable Au NPs. Inset: Corresponding SEM image
Figure 3. (a) J-V characteristics, recorded under illumina- illumina-tion at 100 mW cm -2 (AM 1.5G), of polymer solar cells with (plasmonic device) and without (reference device) Au NPs in the PEDOT:PSS layer
Figure 5. Values of J sc of the OPVs plotted with respect to
Figure 6. PL spectra of the reference and plasmonic samples recorded using excitation source wavelengths ( λ exc ) of (a) 470 and (b) 532 nm.
+3

參考文獻

相關文件

In BHJ solar cells using P3HT:PCBM, adjustment of surface energy and work function of ITO may lead to a tuneable morphology for the active layer and hole injection barrier

In order to use the solar rays more efficient and improve the conversion efficiency of solar cell, it is necessary to use antireflection layer to reduce the losses of

Gratzel, “Low cost photovoltaic modules based on dye sensitized nano-crystalline titanium dioxide and carbon power,” Solar Energy Materials and Solar Cells, Vol.

Zhang, “A flexible new technique for camera calibration,” IEEE Tran- scations on Pattern Analysis and Machine Intelligence,

Kim, “Improvement on surface texturing of single crystalline silicon for solar cells by saw-damage etching using an acidic solution,” Solar Energy Materials and

This study first surveys the thin film solar cell application of new components and thin film solar photovoltaic characteristics of the current situation in the

The aim of this study is to investigate the actual performance of fifth grade children in diagnostic teaching and their changes in scores on fraction word problems test.. Through

Over the last decade, many research has gone into the development of more efficient alternate energy technologies, especially solar photovoltaic cells, are becoming