Block Copolymer as a Nanostructuring Agent for High-Efficiency and Annealing-Free Bulk...

6
© 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 2196 www.advmat.de www.MaterialsViews.com wileyonlinelibrary.com COMMUNICATION Adv. Mater. 2012, 24, 2196–2201 Cédric Renaud, Sébastien-Jun Mougnier, Eleni Pavlopoulou, Cyril Brochon,* Guillaume Fleury, Dargie Deribew, Giuseppe Portale, Eric Cloutet, Sylvain Chambon, Laurence Vignau, and Georges Hadziioannou* Block Copolymer as a Nanostructuring Agent for High- Efficiency and Annealing-Free Bulk Heterojunction Organic Solar Cells Dr. C. Renaud, S.-J. Mougnier, Dr. E. Pavlopoulou, Dr. C. Brochon, Dr. G. Fleury, D. Deribew, Dr. E. Cloutet, Prof. G. Hadziioannou Université de Bordeaux CNRS, Laboratoire de Chimie des Polymères Organiques (LCPO) UMR 5629, 16 avenue Pey-Berland, F-33607 Pessac Cedex, France E-mail: [email protected]; [email protected] Dr. S. Chambon, Dr. L. Vignau Laboratoire d’Intégration du Matériau au Système (IMS) Université de Bordeaux CNRS, UMR 5218, F-33400 Talence, France Dr. G. Portale Netherlands Organization for Scientific Research (NWO) DUBBLE-CRG at the ESRF, Grenoble, France DOI: 10.1002/adma.201104461 Three main requirements for the industrial development of polymer solar cells have to be addressed in order to obtain a competitive technology: a fabrication process compatible with common polymer printing technologies, enhanced lifetime sta- bility, and an improved power conversion efficiency (PCE). [1–3] The active layer nanostructure in bulk heterojunction organic solar cells plays a key role on the properties of charge transfer, transport, and consequently on the PCE. Ideally a cocontinuous network of donor–acceptor domains with a length scale com- parable to the exciton diffusion length is required. [4] To obtain an optimized nanostructured active layer, an annealing process (thermal and/or solvent) is commonly performed, leading to an increase in photovoltaic performance. As an example, for the classical poly(3-hexylthiophene) (P3HT):phenyl-C 61 -butyric acid methyl ester (PCBM) blend, the PCE rises from around 1% to 5% after annealing (the average value being 3%). [5–7] Currently, the implementation of common printing technolo- gies for the fabrication of polymer solar cells on a mechanically flexible polymer substrate is impeded by this annealing step. Indeed, while thermal annealing leads to dramatic failure of the polymer solar cell architecture due to poor thermal and dimen- sional stabilities of the substrate, solvent annealing is not suit- able for roll-to-roll processes as it requires long exposure times to solvent vapors which do not fit the requirements of indus- trial mass production. [8] Moreover, via oxidative mechanisms or thermal decomposition, these processes can damage other parts of the devices such as the electrodes. [5] In order to over- come the limitations inherent to conventional thermal and sol- vent annealing, chemical processing additives can be used. [9] As an example, alkanethiol-based compounds have been studied in this way. [10–12] Polymer additives were also successfully added to the P3HT:PCBM blend to optimize the charge transport. [13] In this Communication, the use of a block copolymer as a nanostructuring agent for the polymer/fullerene derivative blend is explored as a novel approach for the fabrication of annealing-free polymer solar cells. Block copolymers are well known to self-assemble through microphase separation into highly ordered nanostructures both in bulk and thin film config- urations. [14–17] For photovoltaic applications, block copolymers are envisioned to act as a promoter for the self-organisation of the active layer. In the late ‘90s, the Hadziioannou group intro- duced the use of donor–acceptor rod–coil block copolymers in order to nanostructure the active layer of solar cells, the rod block being a donor semiconducting polymer and the coil block being a polymer grafted with an acceptor molecule, such as a fullerene. [18–20] Following these pioneer works, other groups have used the same methodology but in all cases the PCE of the solar cells did not exceed 1%. [21–23] More recently, rod–coil block copolymers composed of a donor and an acceptor block have been used as nanostructuring agents in P3HT:PCBM blends, resulting in an improved mor- phology [24–29] and a significant enhancement of the solar cell efficiency with respect to that of the pristine blend (without the additive) was observed (up to 50% after annealing). This improvement in PCE was attributed to the nanostructuring of the P3HT donor phase in the form of nanoribbons, leading to an increase of the interfacial area between the donor and acceptor phases as well as an improved hole transport. However such a methodology is problematic when C 60 -grafted polymer blocks are used, since the long annealing times promote the macrophase separation of PCBM and its crystallization, leading to a dramatic degradation of the efficiency. [26,29] Sary et al. blended P3HT- b-poly(4-vinylpyridine) (P4VP) rod– coil block copolymers with PCBM to form the active layer of polymer solar cell devices. In this case the P3HT rod block acts as the electron donor material, while the P4VP coil block acts as a molecular dispersant of PCBM through noncovalent interactions. [30,31] By this approach, only moderate PCE were obtained (around 1.5%; yet a record for a binary blend system comprising a semiconducting block copolymer as the donor and PCBM as the acceptor). This was attributed to a poorly developed cocontinuous network structure and to the presence of an important volume fraction of non-conducting material introduced in the blend.

Transcript of Block Copolymer as a Nanostructuring Agent for High-Efficiency and Annealing-Free Bulk...

2196

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATI

ON

Cédric Renaud , Sébastien-Jun Mougnier , Eleni Pavlopoulou , Cyril Brochon ,* Guillaume Fleury , Dargie Deribew , Giuseppe Portale , Eric Cloutet , Sylvain Chambon , Laurence Vignau , and Georges Hadziioannou *

Block Copolymer as a Nanostructuring Agent for High-Effi ciency and Annealing-Free Bulk Heterojunction Organic Solar Cells

Three main requirements for the industrial development of polymer solar cells have to be addressed in order to obtain a competitive technology: a fabrication process compatible with common polymer printing technologies, enhanced lifetime sta-bility, and an improved power conversion effi ciency (PCE). [ 1–3 ] The active layer nanostructure in bulk heterojunction organic solar cells plays a key role on the properties of charge transfer, transport, and consequently on the PCE. Ideally a cocontinuous network of donor–acceptor domains with a length scale com-parable to the exciton diffusion length is required. [ 4 ] To obtain an optimized nanostructured active layer, an annealing process (thermal and/or solvent) is commonly performed, leading to an increase in photovoltaic performance. As an example, for the classical poly(3-hexylthiophene) (P3HT):phenyl-C 61 -butyric acid methyl ester (PCBM) blend, the PCE rises from around 1% to 5% after annealing (the average value being 3%). [ 5–7 ] Currently, the implementation of common printing technolo-gies for the fabrication of polymer solar cells on a mechanically fl exible polymer substrate is impeded by this annealing step. Indeed, while thermal annealing leads to dramatic failure of the polymer solar cell architecture due to poor thermal and dimen-sional stabilities of the substrate, solvent annealing is not suit-able for roll-to-roll processes as it requires long exposure times to solvent vapors which do not fi t the requirements of indus-trial mass production. [ 8 ] Moreover, via oxidative mechanisms or thermal decomposition, these processes can damage other parts of the devices such as the electrodes. [ 5 ] In order to over-come the limitations inherent to conventional thermal and sol-vent annealing, chemical processing additives can be used. [ 9 ] As

© 2012 WILEY-VCH Verlag Gwileyonlinelibrary.com

Dr. C. Renaud , S.-J. Mougnier , Dr. E. Pavlopoulou , Dr. C. Brochon , Dr. G. Fleury , D. Deribew , Dr. E. Cloutet , Prof. G. Hadziioannou Université de Bordeaux CNRS, Laboratoire de Chimie des Polymères Organiques (LCPO) UMR 5629, 16 avenue Pey-Berland, F-33607 Pessac Cedex, France E-mail: [email protected]; [email protected] Dr. S. Chambon , Dr. L. Vignau Laboratoire d’Intégration du Matériau au Système (IMS) Université de Bordeaux CNRS, UMR 5218, F-33400 Talence, France Dr. G. Portale Netherlands Organization for Scientifi c Research (NWO) DUBBLE-CRG at the ESRF, Grenoble, France

DOI: 10.1002/adma.201104461

an example, alkanethiol-based compounds have been studied in this way. [ 10–12 ] Polymer additives were also successfully added to the P3HT:PCBM blend to optimize the charge transport. [ 13 ]

In this Communication, the use of a block copolymer as a nanostructuring agent for the polymer/fullerene derivative blend is explored as a novel approach for the fabrication of annealing-free polymer solar cells. Block copolymers are well known to self-assemble through microphase separation into highly ordered nanostructures both in bulk and thin fi lm confi g-urations. [ 14–17 ] For photovoltaic applications, block copolymers are envisioned to act as a promoter for the self-organisation of the active layer. In the late ‘90s, the Hadziioannou group intro-duced the use of donor–acceptor rod–coil block copolymers in order to nanostructure the active layer of solar cells, the rod block being a donor semiconducting polymer and the coil block being a polymer grafted with an acceptor molecule, such as a fullerene. [ 18–20 ] Following these pioneer works, other groups have used the same methodology but in all cases the PCE of the solar cells did not exceed 1%. [ 21–23 ]

More recently, rod–coil block copolymers composed of a donor and an acceptor block have been used as nanostructuring agents in P3HT:PCBM blends, resulting in an improved mor-phology [ 24–29 ] and a signifi cant enhancement of the solar cell effi ciency with respect to that of the pristine blend (without the additive) was observed (up to 50% after annealing). This improvement in PCE was attributed to the nanostructuring of the P3HT donor phase in the form of nanoribbons, leading to an increase of the interfacial area between the donor and acceptor phases as well as an improved hole transport. However such a methodology is problematic when C 60 -grafted polymer blocks are used, since the long annealing times promote the macrophase separation of PCBM and its crystallization, leading to a dramatic degradation of the effi ciency. [ 26 , 29 ]

Sary et al. blended P3HT- b -poly(4-vinylpyridine) (P4VP) rod–coil block copolymers with PCBM to form the active layer of polymer solar cell devices. In this case the P3HT rod block acts as the electron donor material, while the P4VP coil block acts as a molecular dispersant of PCBM through noncovalent interactions. [ 30 , 31 ] By this approach, only moderate PCE were obtained (around 1.5%; yet a record for a binary blend system comprising a semiconducting block copolymer as the donor and PCBM as the acceptor). This was attributed to a poorly developed cocontinuous network structure and to the presence of an important volume fraction of non-conducting material introduced in the blend.

mbH & Co. KGaA, Weinheim Adv. Mater. 2012, 24, 2196–2201

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATIO

N

Herein we continue our previous work and we propose to use P3HT- b -P4VP block copolymers as nanostructuring agents in P3HT:PCBM bulk heterojunctions in order to promote the structural organization of the active layer at room temperature, while impeding in the same time the macrophase separation and crystallisation of PCBM by taking advantage of the non-covalent supramolecular interactions between the P4VP block and the PCBM moities. Moreover, the use of the P3HT- b -P4VP block copolymer as a surfactant allows us to decrease the frac-tion of the non-conducting material introduced into the blend. Most importantly, we demonstrate in this work that the addition of a low weight fraction of a well-designed P3HT- b -P4VP semi-conducting diblock copolymer results in high effi cient solar cells without necessitating any additional thermal or solvent annealing of the active layer. The PCEs obtained are far beyond the state-of-the-art performance of nonannealed devices and, in fact, they are comparable to those achieved for the thermal annealed pristine P3HT:PCBM bulk heterojunctions.

The rod–coil P3HT- b -P4VP copolymer utilized herein was synthesized according to a previously described strategy by Mougnier et al., using a synthetic route designed for organic electronic requirements. [ 32 ] The synthesis procedure is pre-sented in Scheme 1 . The coil block (P4VP) is obtained by nitroxide mediated radical polymerization (NMRP) from a P3HT macroinitiator without the use of any transition metal catalyst which is one of the key requirements for highly pure materials, since metallic contamination that can lead to poor device stability is avoided. The macroinitiator was obtained

© 2012 WILEY-VCH Verlag GmAdv. Mater. 2012, 24, 2196–2201

Scheme 1 . Synthesis of the diblock P3HT- b -P4VP.

via an intermolecular radical 1,2-addition from a well-defi ned acrylate-monofunctionalized P3HT (see Experimental Section for details). In order to design an effi cient nanostructuring agent, a relatively low molecular weight copolymer was targeted (less than 10 kg mol − 1 ) with a P4VP weight fraction of around 66%. The obtained block copolymer was characterized by 1 H NMR and SEC (Supporting Information (SI), Figure S01 and S02, respectively) which allow us to quantitatively determine its number average molecular weight and composition (P3HT block M n = 2500 g mol − 1 and P4VP block M n = 5000 g mol − 1 ).

Concerning device fabrication, the inverted architecture was adopted in this work since Sary et al. had shown that the poor performance of conventional ITO/PEDOT:PSS/P3HT- b -P4VP:PCBM/Al solar cells was attributed to the presence of a P4VP-rich layer on top of the poly(3,4-ethylenedioxythiophene):-poly(styrenesulfonate) (PEDOT:PSS) fi lm. [ 30 ] This led to a hole collection barrier and/or interfacial dipoles that impeded the overall photovoltaic performance of the cells. The inverted architecture using metal oxides was fi rst reported in 2006. [ 33–35 ] It has the advantage to eliminate the use of PEDOT:PSS as well as to allow the use of a low work function metal as the back con-tact in the device. For the inverted solar cells prepared herein titanium oxide (TiO x ) was used as the electron selective layer and molybdenum oxide (MoO 3 ) as the hole selective layer (see experimental part for details). The conversion procedure from the titanium precursor to TiO x as well as the MoO 3 thickness were optimised on inverted P3HT:PCBM cells so as to obtain photovoltaic performance comparable to those of conventional solar cells. [ 35 , 36 ]

The effect of the incorporation of P3HT- b -P4VP as an addi-tive in 1:1 w/w P3HT:PCBM blends on the performance of the resulting solar cells was studied for both as cast and thermal annealed devices. Various copolymer weight fractions, x , were investigated, from 0% up to 10% with respect to the total weight of the blend. The photovoltaic characteristics obtained are reported in Figure 1 , where (a), (b), (c) and (d) show the var-iations in open circuit voltage ( V oc ), short circuit current density ( J sc ), fi ll factor (FF) and PCE, respectively, when the composition in P3HT- b -P4VP x is increased. Representative current density ( J )–voltage ( V ) curves are presented in the SI, Figure S03, and they correspond to the data acquired for the reference (pure P3HT:PCBM) and the 8% copolymer-containing solar cells. We choose to present the curves obtained for this weight frac-tion since it exhibits the best performances with and without thermal treatment.

Concerning the pristine P3HT:PCBM devices, the thermal annealing treatment leads to an increase in PCE from 1 ± 0.1% to 2.7 ± 0.1%, as displayed in Figure 1 d. This performance improvement is attributed to thermally induced crystallisation and nanostructuring of the blend components which results in enhanced charge generation and improved charge transport. [ 5 ]

When P3HT- b -P4VP is added as a nanostructuring agent a considerable improvement in the overall performance is observed with respect to the pristine P3HT:PCBM devices, for both the as cast and annealed cells. In the case of annealed devices, the addition of P3HT- b -P4VP slightly increases all photovoltaic parameters ( V oc , J sc , FF and PCE). The highest effi ciency is achieved when 8% of copolymer is added and PCE increases from 2.7% to 4.3%. However, the most astonishing

2197wileyonlinelibrary.combH & Co. KGaA, Weinheim

21

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATI

ON

Figure 1 . V oc (a), J sc (b), FF (c), and PCE (d), obtained for as cast ( � ) and annealed ( � ) solar cells at different P3HT- b -P4VP weight fractions x. The error bars were calculated by averaging the results obtained over 16 devices.

effect of the addition of the agent is obtained for the annealing-free devices. The addition of 2% of P3HT- b -P4VP within the active layer leads to an increase in V oc from 0.49 ± 0.02 V to 0.54 ± 0.01 V while it remains constant when the copolymer content increases further, from 2% to 10%. More importantly, the incorporation of the block copolymer strongly improves J sc and FF. In fact, J sc is doubled when 4% of P3HT- b -P4VP is added, from 5.0 ± 0.4 mA cm − 2 to 10.5 ± 0.3 mA cm − 2 , while FF increases by 50%, from 0.40 ± 0.02 to 0.59 ± 0.02. Conse-quently, a PCE of 3.4 ± 0.2% is achieved for the 4% copolymer-containing devices, which suggests a three-fold enhancement of the effi ciency. For copolymer concentrations higher than 4% J sc , FF and PCE remain roughly constant with respect to the P3HT- b -P4VP weight fraction. It is worth noticing that the PCE of the as cast solar cells that contain 4% of P3HT- b -P4VP is even higher than the 2.7% effi ciency achieved for the pristine P3HT:PCBM bulk heterojunctions that were thermally treated at high temperature.

In order to complete the study of the J – V curves and deeper understand the mechanisms involved in this increase of effi -ciency, the evolution of the series ( R s ) and shunt resistances ( R sh ) was investigated as a function of the P3HT- b -P4VP con-tent (SI, Figure S04). For high performance photovoltaic cells a low R s and a high R sh values are required. A high value of R sh is related to the minimization of leakage currents and to a low charge carrier recombination yield, while a low value of R s is associated to a decrease of the intrinsic resistances, including the resistances at the contacts. In the case of annealed devices, R s is independent of the P3HT- b -P4VP content while R sh is enhanced. Concerning the non-annealed devices, the P3HT- b -P4VP-containing solar cells show lower R s as well as higher R sh with respect to the pristine P3HT:PCBM blends. The increase

98 wileyonlinelibrary.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Wein

in R sh is closely related to the enhancement of J sc and FF observed in Figure 1 and sug-gests lower leakage currents than in case of the pristine P3HT:PCBM devices. [ 4 ] This out-come is further supported by the decrease in the reverse saturation current density observed in the dark J – V curves, upon addi-tion of P3HT- b -P4VP (SI, Figure S05). A decrease in leakage currents indicates a reduction in charge losses by trapping and/or by charge recombination in the P3HT- b -P4VP-containing devices. [ 37 , 38 ]

Complementary to the J – V characteri-zation, external quantum effi ciency (EQE) measurements were conducted. The data pre-sented in Figure S06 (SI) show an increase in EQE when the block copolymer is added to the blend. This result suggests an enhanced exciton dissociation yield and/or reduced charge recombination. It is noted that the increase in the integrated EQE is consistent with the increase in J sc and the decrease in charge losses discussed above.

We also checked the devices performance when a higher molecular weight P3HT- b -P4VP copolymer is utilized. The molecular weight of the rod block was kept constant

(2.5 kg mol − 1 ) while that of the coil block varied from 5 to 28 kg mol − 1 . The results are presented in the SI (Figure S07). In the investigated range of P4VP volume fractions in the copoly mer, there is no signifi cant effect on the PCE, neither for the as cast, nor for the annealed devices. However a new series of P3HT- b -P4VP copolymers is currently under study in order to access the whole range of P4VP block volume fractions.

The results presented so far concerning the improved device performance and the reduced charge losses that are observed upon addition of P3HT- b -P4VP, even at a low concentration and without the necessity for thermal annealing, should originate from structural changes in the active layers that are induced by the incorporation of the block copolymer. As the P3HT- b -P4VP copolymer is utilized herein as a nanostructuring agent, it is expected to act as a compatibilizer between the P3HT and PCBM phases, due to the chemical affi nity of the two blocks with the components of the blend. Therefore it is expected to reside at the P3HT–PCBM interface, with the P3HT block lying in the P3HT domain and the P4VP block lying in the PCBM domain, stabilizing, thus, the structure.

In order to get insight into the morphology of the copolymer-containing fi lms, grazing incidence X-ray diffraction (GIXD) measurements have been performed on fi lms coated on TiOx/ITO substrates, prepared following the same methodology as for the active layers of the devices. The 2D images recorded for the as cast P3HT:PCBM blend and the non-annealed fi lm that con-tains 4% of P3HT- b -P4VP are presented in Figure 2 a. By radi-ally averaging with respect to the centre of the incident beam, we obtain the corresponding 1D intensity vs. scattering vector, q , profi les, which are presented in the upper part of Figure 2 b. The integrated intensity of the (100) refl ection of P3HT, which corre-sponds to the alkyl chain periodicity, appears to decrease slightly

heim Adv. Mater. 2012, 24, 2196–2201

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATIO

N

Figure 2 . a) 2D GIXD images recorded for the pristine and the 4% P3HT- b -P4VP-containing P3HT:PCBM blend. Both correspond to as-cast fi lms, spin-coated on TiOx/ITO substrates to mimic the device preparation conditions. b) The corresponding 1D radially averaged intensity plot with respect to the scattering vector q , and the two intensity linecuts, across the in-plane ( q xy ) and out-of-plane ( q z ) directions. In all three cases the contribution to scattering from the TiOx/ITO substrate has been subtracted.

after the incorporation of the copolymer in the blend, which sug-gests that the crystallinity of P3HT is reduced. This is also evident in the UV–vis absorption data (SI, Figure S08). The intensity of the P3HT absorption band located between 450–650 nm slightly decreases with increasing the fraction of copolymer in the blend, which is indicative of a small decrease in the optical density of the fi lm and suggests less densely-packed chains in the crystal-line phase. [ 39 ] Additionally, a close look at the (100) diffraction peak shows that the peak width is the same for the two fi lms, indicating that the size of the P3HT domains is not affected by the presence of the copolymer. We suppose, thus, that although the crystallinity of P3HT decreases after addition of the P3HT- b -P4VP copolymer, this decrease is relatively small. Moreover, it is interesting to note in the same diffraction pattern that the intensity of the peak located at around 1.4 Å − 1 , which is assigned to PCBM, also decreases upon addition of P3HT- b -P4VP. This suggests that less PCBM crystallites are embedded in the fi lms and testifi es that the P4VP blocks interact with the PCBM domains, inhibiting the formation of big PCBM domains. [ 29 ] Finally, by comparing the in-plane and out-of-plane diffraction patterns (Figure 2 b, intensity vs. q xy and intensity vs. q z , respec-tively), we can see that the copolymer incorporation results in an increase of the (100) peak intensity in the in-plane direction and a subsequent decrease of the intensity of the same peak in the out-of-plane direction. In the same time, the intensity of the (010) refl ection of P3HT, which is located at around 1.65 Å − 1 and is assigned to the π – π stacking periodicity, decreases in the in-plane diffraction pattern. All these strongly suggest that the pop-ulation of the face-on oriented P3HT crystallites increases upon incorporation of P3HT- b -P4VP in the active layer.

© 2012 WILEY-VCH Verlag GmbH & Co. KGaA, WeinAdv. Mater. 2012, 24, 2196–2201

To sum up, the results obtained on the structural properties of the as cast fi lms indicate that when P3HT- b -P4VP is added in the P3HT:PCBM blend the crystallinity of P3HT is slightly decreasing, however the P3HT crystallites prefer to orient themselves with faces-on. This is well-desirable in photo-voltaic devices, since hole extraction through the electrodes is enhanced, [ 40 ] resulting in an improved performance of the devices. On the other hand, PCBM crystallization is inhibited by the presence of the P4VP blocks. We suppose that this also contributes to the improvement of the device performance, by decreasing the size of the PCBM domains, increasing, thus, the interfaces available for exciton dissociation. This results in increased exciton dissociation and reduced charge recombination, as confi rmed by the EQE and J – V data. Unfortunately we cannot provide experimental proof on the PCBM domain size by applying the Scherrer’s equa-tion on our data, due to the broadness of the PCBM refl ection and the overlapping with the (010) P3HT peak, that hinder the accu-rate estimation of the peak full width at half maximum. Still, our results strongly suggest that the improved performance is related to the formation of a well-optimized nanoscale

structure that allows better exciton dissociation and charge transport. A more elaborate structural investigation is currently performed that will allow a deeper understanding of the phys-ical mechanisms involved in the observed enhancement of the device performance.

In conclusion, we have reported a systematic investigation of the photovoltaic performance in P3HT:PCBM-based solar cell devices when a well-designed P3HT- b -P4VP block copolymer is added as a nanostructuring agent. The results have shown that the PCE can reach a plateau value of 3.4% as soon as 4% w/w of P3HT- b -P4VP is added without performing any (thermal and/or solvent) annealing treatment. In comparison, a PCE of only 2.7% is obtained after a high temperature annealing treat-ment without any addition of copolymer (with the conventional P3HT:PCBM 1:1 w/w) under the same experimental condi-tions, i.e., in the frame of inverted solar cells. By performing a thermal annealing treatment, the PCE reaches 4.3% in case of the 8% w/w P3HT- b -P4VP-containing cells. The analysis of the J – V characteristics and the initial structural characterization suggest that the increase in performance is related to the for-mation of a well-optimized nanoscale structure that allows for a more effi cient exciton dissociation and charge transport, by low-ering the charge recombination and/or trapping. The simplicity of adding a well designed block copolymer as a nanostructuring agent in the P3HT:PCBM classical active layer appears to be a promising formulation process to enhance the performance of polymer solar cells while allowing in the same time the imple-mentation of a low cost fabrication printable roll-to-roll process on fl exible polymer substrates. A more elaborated morpholog-ical characterization that will allow for a deeper understanding

2199wileyonlinelibrary.comheim

2200

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATI

ON

of the physical mechanisms involved in this enhancement of

performance is currently under investigation.

Experimental Section P3HT with number average molecular weight M n = 60000 g mol − 1 (determined by SEC using PS standards calibration), dispersity (ratio between weight average molecular weight and number average molecular weight M w / M n ) Ð = 1.4, regioregularity (determined by 1 H NMR) = 99%, was synthesized via the Grignard metathesis route. [ 41 , 42 ] PCBM was purchased from Solaris.

Synthesis of the Diblock P3HT-b-P4VP : ω -MAMA-SG1 terminated poly(3-hexylthiophene) macroinitiator was synthesized according to the litterature. [ 32 ] In a dried Schlenk, ω - Blocbuilder terminated P3HT (10 mg, 0.005 mmol) was dissolved in degassed toluene (0.1 mL) and stirred for 1 h. At this point, distilled 4-vinylpyridine (0.9 mmol) was added to the reaction mixture. After stirring for 5 min, the Schlenk was put in an oil bath at 115 ° C and stirred for 5 min. The Schlenk tube was poured into liquid nitrogen in order to quench the polymerization. Then, the reaction mixture was diluted with degassed toluene (0.9 mL) and stirred for 5 min. The block copolymer was isolated by precipitation in cold diethyl ether (10 mL). The copolymer was fi ltered and dried under vacuum for 24 h at 90 ° C to remove remaining monomers. See SI data for more details (SEC analysis, 1 H NMR).

Device Fabrication : P3HT- b -P4VP-containing inverted solar cells were fabricated using the following structure: ITO/TiO x /active layer/MoO 3 /Ag. The inverted device structure was prepared as follows: substrates (glass coated with ITO) were cleaned in an ultrasonic bath in a series of solvents acetone, ethanol and isopropanol. Glass/ITO substrates were coated with an organic precursor of titania (Titanium (IV) isopropoxide, Aldrich, 99.999%) diluted in absolute ethanol (C = 0.05 M ) and stabilized with HCl (pH = 1.9) by spin-coating. Conversion of the precursor to TiO x via hydrolysis was achieved by storing the fi lms for 1 h in air at room temperature. This procedure was derived from the work of Burgos et al. [ 43 ] The thickness of the TiOx layer was 80 nm measured using an Alpha-step IQ surface Profi lometer. All procedures after TiOx deposition were performed in an inert-atmosphere glove box under nitrogen (O 2 and H 2 O < 0.5 ppm). Subsequently, the photoactive layer (P3HT:PCBM 1:1 wt%) was spin-coated on the top of the TiO x layer from ortho -dichlorobenzene ( o -DCB) solutions at 50 ° C. The thickness of the photoactive layer is typically in the range of 100–150 nm. Molybdenum oxide (MoO 3 ) and the silver electrode were thermally evaporated (10 nm/80 nm) through a shadow mask with a pressure of 10 − 7 mbar. The layer of MoO 3 is electron-blocking and hole-selective, in order to improve holes extraction from the active layer to the silver electrode. The active area of the devices was 8.4 mm 2 . In some devices a thermal annealing treatment was performed before MoO 3 and silver deposition on a temperature-controlled hotplate at 165 ° C for 20 min. Devices were left to cool down to room temperature before testing.

A P3HT:PCBM 1:1 w/w solution (40 mg mL − 1 ) was prepared in o -DCB and a P3HT- b -P4VP solution (5 mg mL − 1 ) was prepared in chloroform. The two solutions were mixed in a way to obtain a P3HT- b -P4VP in P3HT:PCBM blend at different copolymer weight ratios x . x is defi ned with respect to the blend amount. Chloroform had to be added in the fi nal solutions so as to have the same ratio between o -DCB and chloroform in all solutions. For the reference, the P3HT:PCBM 1:1 solution in o-DCB was complemented with chloroform as well. See SI, Table S01, for more details about the preparation of the solutions. The P3HT:PCBM:P3HT- b -P4VP solution was then spin-coated at 1000 rpm for 60 s on top of TiO x .

Device Characteristic Measurements : Current density-voltage curves were recorded using a Keithley 4200 SCS , under an illumination of 100 mW/cm 2 from a K.H.S. Solar Cell test 575 solar simulator with AM1.5G fi lters. A Triax Monochromator was utilized for the EQE measurements. Four 8.4 mm 2 photovoltaic cells were fabricated on each substrate. In order to check the reproducibility of the performance, the measurements have been performed on two series of eight solar cells prepared separately from two different solutions.

wileyonlinelibrary.com © 2012 WILEY-VCH Verlag

[ 1 ] M. Jørgensen , K. Norrman , F. C. Krebs , Solar Energy Mater. Solar Cells 2008 , 92 , 686 .

[ 2 ] F. C. Krebs , Solar Energy Mater. Solar Cells 2008 , 92 , 715 . [ 3 ] F. C. Krebs , Solar Energy Mater. Solar Cells 2009 , 93 , 394 . [ 4 ] H. Hoppe , N. S. Sariciftci , J. Mater. Res. 2004 , 19 , 1924 . [ 5 ] W. Ma , C. Yang , X. Gong , K. Lee , A. J. Heeger , Adv. Funct. Mater.

2005 , 15 , 1617 . [ 6 ] M. T. Dang , L. Hirsch , G. Wantz , Adv. Mater. 2011 , 23 , 3597 . [ 7 ] G. Li , V. Shrotriya , J. Huang , Y. Yao , T. Moriarty , K. Emery , Y. Yang ,

Nat. Mater. 2005 , 4 , 864 . [ 8 ] F. C. Krebs , S. A. Gevorgyan , B. Gholamkhass , S. Holdcroft ,

C. Schlenker , M. E. Thompson , B. C. Thompson , D. Olson , D. S. Ginley , S. E. Shaheen , H. N. Alshareef , J. W. Murphy , W. J. Youngblood , N. C. Heston , J. R. Reynolds , S. Jia , D. Laird , S. M. Tuladhar , J. G. A. Dane , P. Atienzar , J. Nelson , J. M. Kroon , M. M. Wienk , R. A. J. Janssen , K. Tvingstedt , F. Zhang , M. Andersson , O. Inganäs , M. Lira-Cantu , R. de Bettignies , S. Guillerez , T. Aernouts , D. Cheyns , L. Lutsen , B. Zimmermann , U. Würfel , M. Niggemann , H.-F. Schleiermacher , P. Liska , M. Grätzel , P. Lianos , E. A. Katz , W. Lohwasser , B. Jannon , Solar Energy Mater. Solar Cells 2009 , 93 , 1968 .

[ 9 ] A. Pivrikas , H. Neugebauer , N. S. Sariciftci , Solar Energy 2011 , 85 , 1226 .

[ 10 ] N. E. Coates , I.-W. Hwang , J. Peet , G. C. Bazan , D. Moses , A. J. Heeger , Appl. Phys. Lett. 2008 , 93 , 072105 .

[ 11 ] A. Pivrikas , P. Stadler , H. Neugebauer , N. S. Sariciftci , Org. Elec-tronics 2008 , 9 , 775 .

[ 12 ] J. Peet , C. Soci , R. C. Coffi n , T. Q. Nguyen , A. Mikhailovsky , D. Moses , G. C. Bazan , Appl. Phys. Lett. 2006 , 89 , 252105 .

[ 13 ] M.-C. Chen , D.-J. Liaw , W.-H. Chen , Y.-C. Huang , J. Sharma , Y. Tai , Appl. Phys. Lett. 2011 , 99 , 223305 .

[ 14 ] F. S. Bates , G. H. Fredrickson , Physics Today 1999 , 52 , 32 . [ 15 ] L. Leibler , Macromolecules 1980 , 13 , 1602 .

Grazing Incidence X-ray Diffraction Measurements : GIXD experiments were performed on the Dutch-Belgian Beamline (DUBBLE CIG), station BM26B, at the European Synchrotron Radiation Facility (ESRF), Grenoble, France. The energy of the X-rays was 12 keV and the angle of incidence was set at 0.15 ° . The diffracted intensity was recorded by a Frelon CCD camera and it was normalized by the incident photon fl ux and the acquisition time. The scattering vector q is defi ned with respect to the center of the incident beam and has a magnitude of q = (4 π / λ ) sin θ , where 2 θ is the scattering angle and λ is the wavelength of the X-ray beam.

Supporting Information Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements CR and SJM contributed equally to this work. Authors would like to acknowledge the fi nancial support from Arkema Corporation, Aquitaine County and the French Research Agency (ANR) through the Habisol program via the project “Spirwind ANR-08-HABISOL-014”. Finally the ESRF and NWO are acknowledged for allocating beamtime at the Dutch-Belgian beamline (DUBBLE) for the GIXD experiments.

Received: November 22, 2011 Revised: December 29, 2011

Published online: March 22, 2012

GmbH & Co. KGaA, Weinheim Adv. Mater. 2012, 24, 2196–2201

www.advmat.dewww.MaterialsViews.com

CO

MM

UN

ICATIO

N

[ 16 ] B. D. Olsen , X. Li , J. Wang , R. A. Segalman , Macromolecules 2007 , 40 , 3287 .

[ 17 ] W. van Zoelen , G. ten Brinke , Soft Matter 2009 , 5 , 1568 . [ 18 ] B. de Boer , U. Stalmach , P. F. van Hutten , C. Melzer , V. V. Krasnikov ,

G. Hadziioannou , Polymer 2001 , 42 , 9097 . [ 19 ] U. Stalmach , B. de Boer , A. D. Post , P. F. van Hutten ,

G. Hadziioannou , Angew. Chem. Int. Ed. 2001 , 40 , 428 . [ 20 ] U. Stalmach , B. de Boer , C. Videlot , P. F. van Hutten ,

G. Hadziioannou , J. Am. Chem. Soc. 2000 , 122 , 5464 . [ 21 ] S. P. Economopoulos , C. L. Chochos , V. G. Gregoriou , J. K. Kallitsis ,

S. Barrau , G. Hadziioannou , Macromolecules 2007 , 40 , 921 . [ 22 ] S. M. Lindner , S. Hüttner , A. Chiche , M. Thelakkat , G. Krausch ,

Angew. Chem. Int. Ed. 2006 , 45 , 3364 . [ 23 ] M. Sommer , M. Thelakkat , Europ. Phys. J.—Appl. Phys. 2006 , 36 ,

245 . [ 24 ] J. U. Lee , J. W. Jung , T. Emrick , T. P. Russell , W. H. Jo , J. Mater.

Chem. 2010 , 20 , 3287 . [ 25 ] F. Richard , C. Brochon , N. Leclerc , D. Eckhardt , T. Heiser ,

G. Hadziioannou , Macromol. Rapid Commun. 2008 , 29 , 885 . [ 26 ] K. Sivula , Z. T. Ball , N. Watanabe , J. M. J. Fréchet , Adv. Mater. 2006 ,

18 , 206 . [ 27 ] M. Urien , H. Erothu , E. Cloutet , R. C. Hiorns , L. Vignau , H. Cramail ,

Macromolecules 2008 , 41 , 7033 . [ 28 ] C. Yang , J. K. Lee , A. J. Heeger , F. Wudl , J. Mater. Chem. 2009 , 19 ,

5416 . [ 29 ] V. Gernigon , P. Leveque , C. Brochon , J.-N. Audinot ,

N. Leclerc , R. Bechara , F. Richard , T. Heiser , G. Hadziioannou , Europ. Phys. J.—Appl. Phys. 2011 , 56 , 34107 .

© 2012 WILEY-VCH Verlag GAdv. Mater. 2012, 24, 2196–2201

[ 30 ] N. Sary , F. Richard , C. Brochon , N. Leclerc , P. Lévêque , J.-N. Audinot , S. Berson , T. Heiser , G. Hadziioannou , R. Mezzenga , Adv. Mater. 2009 , 22 , 763 .

[ 31 ] N. Sary , L. Rubatat , C. Brochon , G. Hadziioannou , R. Mezzenga , Macromolecular Symposia 2008 , 268 , 28 .

[ 32 ] S.-J. Mougnier , C. Brochon , E. Cloutet , S. Magnet , C. Navarro , G. Hadziioannou , J. Polym. Sci. Part A 2012 , DOI: 10.1002/pola.26023.

[ 33 ] G. Li , C. W. Chu , V. Shrotriya , J. Huang , Y. Yang , Appl. Phys. Lett. 2006 , 88 , 253503 .

[ 34 ] M. S. White , D. C. Olson , S. E. Shaheen , N. Kopidakis , D. S. Ginley , Appl. Phys. Lett. 2006 , 89 , 143517 .

[ 35 ] C. Waldauf , Appl. Phys. Lett. 2006 , 89 , 233517 . [ 36 ] C. Tao , Appl. Phys. Lett. 2008 , 93 , 193307 . [ 37 ] G. Li , J. Appl. Phys. 2005 , 98 , 043704 . [ 38 ] A. Zen , J. Pfl aum , S. Hirschmann , W. Zhuang , F. Jaiser ,

U. Asawapirom , J. P. Rabe , U. Scherf , D. Neher , Adv. Funct. Mater. 2004 , 14 , 757 .

[ 39 ] W. C. Tsoi , S. J. Spencer , L. Yang , A. M. Ballantyne , P. G. Nicholson , A. Turnbull , A. G. Shard , C. E. Murphy , D. D. C. Bradley , J. Nelson , J.-S. Kim , Macromolecules 2011 , 44 , 2944 .

[ 40 ] H. Sirringhaus , P. J. Brown , R. H. Friend , M. M. Nielsen , K. Bechgaard , B. M. W. Langeveld-Voss , A. J. H. Spiering , R. A. J. Janssen , E. W. Meijer , P. Herwig , D. M. de Leeuw , Nature 1999 , 401 , 685 .

[ 41 ] R. S. Loewe , S. M. Khersonsky , R. D. McCullough , Adv. Mater. 1999 , 11 , 250 .

[ 42 ] A. Yokoyama , R. Miyakoshi , T. Yokozawa , Macromolecules 2004 , 37 , 1169 .

[ 43 ] M. Burgos , M. Langlet , J. Sol–Gel Sci. Technol. 1999 , 16 , 267 .

2201wileyonlinelibrary.commbH & Co. KGaA, Weinheim