You are on page 1of 17

119

CHAPTER 4 EFFECTS OF ACTIVE LAYER THICKNESS AND THERMAL ANNEALING ON P3HT-PCBM PHOTOVOLTAIC CELLS

4.1 INTRODUCTION
With a steady improvement in energy conversion efficiency during the past decades, organic photovoltaics (OPV) has evolved into a promising technology for renewable energy made possible by the first report of planar donor-acceptor heterojunction
[1]

where efficient exciton dissociation takes place. The device

efficiency got another boost with the introduction of bulk heterojunction OPV (BHJOPV)
[2]

comprising blends of donor and acceptor materials that phase-separate into

nanoscale interpenetrating networks to dramatically increase the donor-acceptor interface per unit volume. The most promising BHJ-OPV devices to date consist of conjugated polymers, such as poly(3-hexylthiophene) (P3HT), blended with soluble fullerene derivatives, such as [6,6]-phenyl C61 butyric acid methyl ester (PCBM). Both thermal annealing [3, 4] and solvent-vapor annealing [5, 6] have been performed to optimize phase separation between P3HT and PCBM and to induce crystallization in the P3HT domain with an objective of improving device efficiency. In addition,

120

thermal annealing of completed devices has been reported to improve charge collection efficiency by promoting active layer/electrode interaction
[4]

. The active

layer thickness has been limited largely to 100-200 nm, although a thicker active layer should absorb more incident light to be more efficient in power conversion. Thick-film polymer BHJ-OPV devices, however, have been found to be less efficient than thin-film devices because of inadequate mobility for supporting charge transport to electrodes for collection.[7,
8]

There have been no systematic studies of how

chemical purity of conjugated polymers affects device efficiency. Using a copper phthalocyanine:C60 active layer, Hiramoto et al. have reported the effects of chemical purity via repeated sublimations and of active layer thickness up to microns on BHJOPV device efficiency.[9] In this study we have examined the thickness-dependent device characteristics of BHJ-OPV with a P3HT:PCBM layer thickness from 130 to 1200 nm. In particular, it will be demonstrated that thick-film devices can be made as efficient as their thinfilm counterparts through thermal annealing. Furthermore, the observed OPV device characteristics can be accounted for by vertical phase separation in the active layer and Li+ diffusion from the Al/LiF cathode into the neighboring active layer.

4.2 EXPERIMENTAL
Device Fabrication

121

Pre-patterned indium tin oxide (ITO) substrates (1.5 1.5 inch) with surface resistance of 15 / square were cleaned and treated with oxygen plasma prior to the deposition of a thin MoO3 layer by thermal sublimation at 0.03 nm/s. Regioregular P3HT (Rieke Metal) and PCBM (Nano-C) at a 1:1 mass ratio were dissolved in odichlorobenzene by stirring at 50 oC under nitrogen atmosphere overnight. The active layers were spin-coated from 0.2 g of the solution onto the MoO3-coated ITO substrates at 300 rpm for 10 s plus 600 rpm for 60 s, and kept in a covered petri dish for solvent vapor annealing for 3 h to create bulk heterojunctions [5]. The solution concentration was varied to adjust the film thickness measured with an optical interferometer (Zygo NewView 5000). The devices were completed by successive deposition of LiF and Al at 0.01 and 1 nm/s, respectively, through a shadow mask to define an active area of 0.1 cm2. The finished device has the structure: ITO/MoO3(3 nm)/P3HT:PCBM(x nm)/LiF(0.8 nm)/Al(100 nm), where x is varied from 130 to 1200 nm. Incremental thermal annealing was conducted by placing the tested device on a hotplate set at 110oC for a pre-determined period of time before quenching to room temperature on an aluminum block for the next round of measurement. Spin coating and thermal annealing were performed in a nitrogen-filled glove box with a oxygen level less than 1ppm, and the thermal evaporation was performed at a base pressure less than 4 106 torr. Device Characterization

122

Devices were characterized in ambient atmosphere without encapsulation. The current density (J) voltage (V) characteristics were recorded with a Keithley 2400 under the illumination of white light at 100 mW/cm2. The white light was generated using an Oriel xenon lamp solar simulator equipped with an AM 1.5 G color filter, and the intensity is calibrated with a silicon diode traceable to National Renewable Energy Laboratory. Defined as the ratio of maximum attainable electrical power to the incident light power (Pin), the power conversion efficiency () was calculated as JSCVOCFF/Pin, where JSC, VOC and FF are short circuit current density, open circuit voltage, and fill factor, respectively. The reported values are accompanied by an error of 5%. For the measurement of spectral response, a beam of white light is passed through a monochromator before being focused onto the devices with photocurrent recorded by the Keithley 2400. The monochromatic photocurrent was recorded at a step size of 10 nm. The external quantum efficiency (EQE) is calculated as the ratio of collected electrons to incident photon with a typical error of 5%. For the measurement of spectral response under applied field, the dark current under the same applied field was subtracted from the recorded photocurrent.

4.3 RESULTS AND DISCUSSION


Depicted in Figure 4.1 is the schematic diagram of device architecture together with the chemical structures of P3HT and PCBM. Recently, there has been a growing interest in adding a metal oxide layer on the ITO anode, such as MoO3,
[10]

to improve hole-collection efficiency. The bulk heterojunction between P3HT and

123

PCBM were created by solvent vapor annealing. [5] To facilitate electron collection, a thin LiF layer was inserted between the active layer and the Al cathode. [11]

e
Al (100 nm) LiF (0.8 nm) P3HT:PCBM (x nm) MoO3 (3 nm) ITO Glass

h+ P3HT PCBM

Figure 4.1. Schematic diagram of the BHJ-OPV device architecture and chemical structures of P3HT and PCBM.
15 J, mA/cm
Unannealed Annealed
2

10 5 0 5 10 0.5 1.0 V, V

1.0

0.5

Figure 4.2. J-V characteristics under 100 mW/cm2 light illumination before and after thermal annealing at 110oC for 20 min of 130-nm BHJ-OPV devices without the LiF layer to improve electron collection-efficiency over Al cathode alone.

124

As the baseline experiment, a BHJ-OPV device with a 130-nm active layer but without a LiF layer was fabricated and characterized to yield the J-V relationship as displayed in Figure 4.2. It was found that = 1.3 and 0.9% before and after thermal annealing at 110oC for 20 min, respectively, both inferior to = 2.9 and 1.1% from devices with a LiF layer processed under the same conditions as to be shown below. Therefore, a LiF layer was consistently adopted in all the BHJ-OPV devices in the present study.

130 nm 220 nm 370 nm 540 nm 830 nm 1200 nm

20 J, mA/cm 15 10 5 0 5 10 0.5

1.0

0.5

1.0 V, V

Figure 4.3: J-V characteristics under 100 mW/cm2 light illumination of BHJ-OPV devices with varying active layer thicknesses before thermal annealing.

A series of OPV devices with an active layer thickness varied from 130 to 1200 nm were fabricated and characterized. Their J-V curves before thermal annealing are plotted in Figure 4.3, and the relevant performance parameters are compiled in Table 4.1. Note that all the devices exhibit a weak dependence of photocurrent on the

125

applied field with virtually the same FF values around 0.62. As the film thickness increases from 130 to 1200 nm, JSC decreases sharply from 8.4 to 1.3 mA/cm2 accompanied by a VOC decreasing from 0.58 to 0.50 V. The weak dependence of photocurrent on applied field for thick-film device precludes recombination as a possible cause for the low JSC. [12]

5 4

0.8

(a)

Absorbance

EQE

3 2 1 0 400

130 nm 220 nm 370 nm 540 nm 830 nm 1200 nm

(b)
0.6 0.4 0.2

130 nm 220 nm 370 nm 540 nm 830 nm 1200 nm 1200 nm

500 600 700 Wavelength, nm

800

400

500 600 700 Wavelength, nm

800

Figure 4.4: (a) UV-vis absorption spectra of P3HT:PCBM blend films with varying active layer thicknesses before thermal annealing, and (b) Spectral responses from BHJ-OPV devices with varying active layer thicknesses before thermal annealing; the dotted curve represents the spectral response from a 1200-nm device under illumination through the semitransparent cathode.

The absorption spectra of P3HT:PCBM films with varied thicknesses are presented in Figure 4.4a. All the films show a main peak at 515 nm with two shoulders at 560 and 610 nm, which originates from the highly crystalline P3HT domains with improved stacking. [4, 5] As the film thickness reaches 830 nm, essentially all incident light between 350 and 610 nm is absorbed in one optical path. The spectral responses

126

of the devices under short circuit conditions are presented in Figure 4.4b. The EQE spectrum of the 130-nm device tracks the P3HT:PCBM absorption spectrum quite well as expected.

Table 4.1: Performance parameters of BHJ-OPV devices with varying active layer thicknesses before and after thermal annealing at 110oC for 20 min.
Unannealed JSC (mA/cm2) 8.2 7.6 6.4 3.9 1.6 1.3 VOC (V) 0.58 0.58 0.58 0.55 0.52 0.50 FF 0.61 0.61 0.62 0.61 0.65 0.64 Annealed

Thickness (nm) 130 220 370 540 830 1200

(%) 2.9 2.7 2.3 1.3 0.5 0.4

JSC (mA/cm2) 4.4 5.3 6.7 8.7 11.3 9.2

VOC (V) 0.57 0.60 0.61 0.61 0.61 0.61

FF 0.42 0.54 0.57 0.53 0.45 0.40

(%) 1.1 1.7 2.4 2.8 3.1 2.3

As the film thickness increases from 130 nm, the relative EQE around absorption peak decreases and eventually dwindles to a valley for film thickness beyond 540 nm with the maximum EQE shifted to the absorption edge. This inverse spectral response indicates that photons that are more strongly absorbed by P3HT are less likely to be converted into collected charge carriers, as observed in planar heterojunction devices with a layer thickness much greater than exciton diffusion length.
[13]

The apparent filtering effect arises from excitons created in the

neighborhood of ITO electrode have limited donor-acceptor interface within their diffusion length to dissociate efficiently. We propose that vertical phase separation in

127

a thick film yield a separated P3HT layer on ITO as inspired by numerous recent reports. [14-16] This P3HT layer absorbs most of incident photons to produce excitons that fail to diffuse sufficiently deep into the region rich in P3HT:PCBM interfaces for dissociation into charge carriers, resulting in a low photocurrent density. On the other hand, photons with wavelength in the weakly absorbing region can penetrate deeper into the active layer where more abundant P3HT:PCBM interfaces are available for efficient exciton dissociation. Meanwhile, the resultant holes and electrons are transported in spatially separated P3HT and PCBM domains, respectively, without encountering significant recombination, thereby achieving FF values in thick-film devices comparable to those in thin-film devices. Furthermore, the reduced exciton dissociation efficiency in vertically separated films is also responsible for the inferior VOC in thick-film devices.[17] Overall, decreases steadily from 2.9% for the 130-nm device to 0.4% for the 1200-nm device before thermal annealing. To confirm the presence of a separated P3HT layer on ITO as an explanation for the inverse spectral response in thick-film devices, Al(100 nm) was replaced with a semitransparent Al(2 nm)/Ag(10 nm) having a transmittance of about 70% in the 350 to 650 nm spectral region. As shown in Figure 4.4b, illumination through the semitransparent metal electrode of a 1200-nm device yielded a spectral response consistent with its absorption spectrum with a much improved JSC = 3.8 mA/cm2. The finished devices with a 1200-nm active layer were further subjected to thermal annealing at 110oC over incremental durations. The J-V characteristics as a

128

function of annealing time are plotted in Figure 4.5, and their performance parameters are summarized in Table 4.2. The JSC increases steadily upon thermal annealing, reaching the maximum value of 10.1 mA/cm2 after annealing for 40 min, accompanied by VOC rising from 0.50 to 0.62 V and FF decreasing from 0.64 to 0.40, corresponding to an improvement of from 0.4 to 2.5%. The maximum of 2.6% was achieved with thermal annealing for 90 min, beyond which the value decreases slightly to 2.2% because of the deteriorated JSC.

15 J, mA/cm2
0 min 4 min 10 min 20 min 40 min 60 min 10 90 min 120 min 5

1.0

0.5

0 5 10

0.5

1.0 V, V

Figure 4.5: J-V characteristics under 100 mW/cm2 white light illumination of a 1200nm BHJ-OPV device thermally annealed at 110oC up to 120 min. Similar to what has been reported previously, [5] thermal annealing of solvent-vaporannealed films did not significantly alter the absorption spectra, as shown in Figure 4.6a. However, the EQE around the absorption maximum is substantially improved upon thermal annealing, thus reversing the inverse spectral response (Figure 4.6b) through morphological modification of the separated P3HT layer on ITO. It has been reported that thermal annealing of polymer/fullerene bilayer films induces the

129

penetration of C60

[13]

and PCBM

[18]

domains into the underlying electron-donating

polymer layers, creating extended donor/acceptor interfaces to improve OPV device efficiency. By analogy the thermally induced penetration of PCBM domains into the separated P3HT layer in thick-film devices is expected to create P3HT/PCBM interfaces in the separated P3HT layer on ITO for efficient exciton dissociation but apparently at the expense of increased charge recombination, thus diminishing FF value with thermal annealing.
5 4
0.8

(a)
0 min 20 min

(b)
0.6

EQE

3 2 1 0 400

0.4 0.2 0

0 min 4 min 10 min 20 min 40 min 60 min 90 min 120 min

Absorbance

500 600 700 Wavelength, nm

800

400

500 600 700 Wavelength, nm

800

Figure 4.6: (a) UV-vis absorption spectra of a 1200-nm film before and after thermal annealing at 110oC for 20 min, and (b) Spectral responses of a 1200-nm BHJ-OPV device thermally annealed at 110oC up to 120 min.

Devices with active layers of other thicknesses were also subjected to thermal annealing at 110oC for 20 min, and their performance parameters are included in Table 4.1. Similar to the 1200-nm device, of 830- and 540-nm devices is improved upon thermal annealing from 0.5 and 1.3% to 3.1 and 2.8%, respectively. In contrast, the same thermal treatment of 130- and 220-nm devices leads to inferior values,

130

from 2.9 and 2.7% to 1.1 and 1.7%, respectively. The dark J-V curves (Figure 4.7) reveals significant current leakage under reverse bias for thin film devices, which could have been caused by the conducting channels created by diffusion of Li+ into the active organic layer upon thermal annealing. [19] Moreover, Li+ diffusion would also consume active organic materials to decrease JSC. In the absence of a LiF layer, however, JSC was retained at 8.2 mA/cm2 upon thermal annealing (Figure 4.2), and no insignificant current leakage was detected in the dark J-V curve (Figure 4.7).

Table 4.2: Performance parameters of the 1200-nm BHJ-OPV device after thermal annealing at 110oC up to 120 min.
Annealing time (min) 0 4 10 20 30 40 60 90 120 JSC (mA/cm2) 1.3 3.5 6.5 9.2 9.8 10.1 9.8 9.1 7.6 VOC (V) 0.50 0.58 0.60 0.61 0.62 0.62 0.63 0.63 0.62 FF 0.64 0.56 0.45 0.40 0.40 0.40 0.41 0.44 0.47

(%) 0.4 1.2 1.8 2.3 2.4 2.5 2.5 2.6 2.2

131

10 10

5 3

J, mA/cm

10 10 10 10 10

130 nm, Al 130 nm, LiF/Al 1200 nm, LiF/Al

3 5 7 9

10 1.0

0.5

0.0 V, V

0.5

1.0

Figure 4.7: J-V characteristics in dark of BHJ-OPV devices with a 130- and 1200-nm active layer thermally annealed at 110oC for 20 min. The dotted curve represents the J-V characteristics of a 130-nm device without a LiF layer.

Let us proceed to examine the spectral response from a 1200-nm device that has been thermally annealed at 110oC for 20 min, where the extent of Li+ diffusion is overshadowed by the active layer thickness. It is noted that incident photons in the 350-610 nm spectral range are absorbed to practically the full extent, viz. greater than 99% (Figure 4.6a). Nonetheless, EQE reaches its maximum value at 610 nm but rolls off toward the short wavelength regime. Light in the long wavelength region is absorbed primarily by crystalline P3HT domains, [4, 5] where more efficient charge generation is expected of the lower ionization potential compared to amorphous domains.
[20]

In contrast, charge generation is less efficient and hence the inferior

EQE in the short wavelength region, where light absorption is contribute mainly by
amorphous P3HT domains. The sensitivity of charge generation to P3HT morphology originates in the weak built-in electric field strength in thick-film

132

devices. Therefore, an applied electric field is expected to facilitate charge generation, thus significantly improving EQE across the entire light absorption region, as borne out in Figure 4.8.

0.8 0.6
0V 5 V 20 V 40 V

EQE

0.4 0.2 0 400

1200 nm
500 600 700 Wavelength, nm 800

Figure 4.8: Electric-field dependence of spectral responses under reverse bias of a 1200-nm BHJ-OPV device after thermal annealing at 110oC for 20 min.

4.4 SUMMARY
Bulk heterojunction organic photovoltaic cells comprising an active layer of P3HT:PCBM blend film with thickness from 130 to 1200 nm were fabricated and characterized before and after thermal annealing. Before thermal annealing, both short circuit current density and power conversion efficiency decrease with an increasing film thickness, resulting in an inverse spectral response for thick-film devices. After thermal annealing at 110oC for 20 min, the efficiency of the 130-nm device decreases from 2.9 to 1.1%, while that of the 1200-nm device increases from 0.4 to 2.6% without the inverse spectral response. Contrary to previous reports, we

133

have demonstrated that thick-film OPV devices can outperform thin-film devices with thermal annealing. The variations of the photovoltaic characteristics with active layer thickness and thermal annealing are understandable in terms of vertical phase separation in P3HT:PCBM films and Li+ diffusion from the Al/LiF cathode into the neighboring active layer.

134

REFERENCES:
[1]. [2]. Tang, C. W. Appl. Phys. Lett. 1986, 48, 183. Yu, G.; Gao, J.; Hummelen, J. C.; Wudle, F.; Heeger, A. J. Science 1995, 270, 1789. [3]. [4]. Padinger, F.; Rittberger, R.; Sariciftci, N. Adv. Funct. Mater. 2003, 13, 85. Ma, W. L.; Yang, C. Y.; Gong, X.; Lee, K.; Heeger, A. J. Adv. Funct. Mater. 2005, 15, 1617. [5]. Li, G.; Shrotriya, S.; Huang, J. S.; Yao, Y.; Moriarty, T.; Emery, K.; Yang, Y. Nat. Mater. 2005, 4, 864. [6]. Zhao, Y.; Xie, Z. Y.; Qu, Y.; Geng, Y. H.; Wang, L. X. Appl. Phys. Lett. 2007, 90, 043904. [7]. [8]. Somani, P. R.; Somani, S. P.; Umeno, M. Appl. Phys. Lett. 2007, 91, 173503. Kim, M. S.; Kim, B. G.; Kim, J. S. ACS. Appl. Mater. Interfaces 2009, 1, 1264. [9]. [10]. Sakai, K.; Hiramoto, M. Mol. Cryst. Liq. Cryst. 2008, 491, 284. Shrotriya, V.; Li, G.; Yao, Y.; Chu, C. W.; Yang, Y. Appl. Phys. Lett. 2006, 88, 073508. [11]. [12]. Ahlswede, E.; Hanisch, J.; Powalla, M. Appl. Phys. Lett. 2007, 90, 163504 Koster, L. J. A.; Mihailetchi, V. D.; Blom, P. W. M. Appl. Phys. Lett. 2006, 88, 052104. [13]. Drees, M.; Davis, R. M.; Heflin, J. R. Phys. Rev. B. 2004, 69, 165320.

135

[14].

Campoy-Quiles, M.; Ferenczi, T.; Agostinelli, T.; Etchegoin, P. G.; Kim, Y.; Anthopoulos, T.; Stavrinou, P. N.; Bradley, D. D. C.; Nelson, J. Nat. Mater. 2008, 7, 158.

[15].

Xu, Z.; Chen, L. M.; Yang, G. W.; Huang, C. H.; Hou, J. H.; Wu, Y.; Li, G.; Hsu, C. S.; Yang, Y. Adv. Funct. Mater. 2009, 19, 1227.

[16]. [17].

van Bavel, S. S.; Sourty, E.; de With, G.; Loos, J. Nano. Lett. 2009, 9, 507. Koster, L. J. A.; Mihailetchi, V. D.; Ramaker, R.; Blom, P. W. M. Appl. Phys. Lett. 2005, 86, 123509.

[18]. [19]. [20].

Kumar, A.; Li, G.; Hong, Z. R.; Yang, Y. Nanotechnology 2009, 20, 165202. Tachikawa, H. J. Phys. Chem. C. 2007, 111, 13087. Clarke, T. M.; Ballantyne, A. M.; Nelson, J.; Bradley, D. D. C.; Durrant, J. R. Adv. Funct. Mater. 2008, 18, 4029.

You might also like