You are on page 1of 13

Macromolecular

Chemistry and Physics

Full Paper

Advanced Functional Polymers for Increasing


the Stability of Organic Photovoltaics
Eva Bundgaard, Martin Helgesen, Jon E. Carl, Frederik C. Krebs,
Mikkel Jrgensen*

The development of new advanced polymers for improving the stability of OPV is reviewed.
Two main degradation pathways for the OPV active layer are identied: photochemically
initiated reactions primarily starting in the side chains and morphological changes that
degrade the important nanostructure. Chemical units can be introduced that impart an
increased stability. Similarly, the morphological degradation of the optimal nanostructure can
be reduced. Active polymers and blends with acceptor material are used to create nanoparticle links with controlled size. Most of these advanced polymers and processing methods have
only been utilized in small-scale devices
prepared by standard techniques such as
spin coating, but a few cases of roll-to-roll
processed solar cells with heat-cleaved side
chains are discussed.

1. Introduction
Organic photovoltaics (OPV) is a high-interest fast-paced
research area with the main effort being to increase the
power conversion efciency (PCE), which is now around
10%.[13] Although this is important and may also be necessary to compete with existing photovoltaic technologies, it
is equally important to improve the stability[46] and later
transfer these results from laboratory-scale devices to rollto-roll (R2R) manufacturing.[79]
The stability and degradation of OPV are complex
issues due to the possible different structures of a
device.[46] In the present review, we focus on the active
layer that contributes to several degradation pathways
that are important to the overall stability of OPV. Since
E. Bundgaard, M. Helgesen, J. E. Carl,
F. C. Krebs, M. Jrgensen
Department of Energy Conversion and Storage,
Technical University of Denmark, Frederiksborgvej 399,
DK-4000 Roskilde, Denmark
E-mail: mijq@dtu.dk

1546

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

this layer is responsible for absorbing the light, it is also


subject to photodegradation with a great variation due
to differences in the chemical structure.[1015] In most
OPVs, the donor material is a conjugated polymer with
alkyl side chains, and the dominant mode of degradation is reaction with oxygen.[1620] One previously popular
type of polymer was the poly(phenylene vinylene) (PPV)
class of materials that have been abandoned because it
is very susceptible to attack from oxygen at the exocyclic vinylene bonds.[21] These polymers were then superseded by poly(3-hexylthiophene) (P3HT), which is much
more resistant, but still undergoes oxidative degradation
starting at the side chains in a reaction type that may
be common to the many polymers used now (Figure 1
top).[13,22,23] In order to improve the efciency of OPV, a
large number of different polymers have been synthesized
and tested, while the stability issue has been somewhat
neglected. In order to at least give a qualitative overview,
a number of studies have focused on ranking the photochemical stability of conjugated polymers according to
which monomers are used.[10,11,24] This has led to a rule
of thumb specifying a subset of monomers that impart

wileyonlinelibrary.com

DOI: 10.1002/macp.201300076

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

such as in, for example, 9,9-dialkyluorene and N-alkylcarbazole, promote


degradation (Figure 1 bottom). This is
clearly demonstrated in Figure 2 where
the donor and acceptor monomers
are illustrated according to increasing
stability.[11]
Another very important issue is
maintaining an optimal nanomorphology of the active layer, which is
commonly a blend of a donor material
such as a conjugated polymer and an
acceptor of the fullerene type.[26,27] The
Figure 1. Two of the key reactions in the oxidation mechanism of P3HT suggested by
accepted view is that this mixture is
Manceau et al.[13,25] (top). One of the key reactions in the oxidation mechanism of polyideally a so-called bulk heterojunction,
uorenes suggested by Grisorio et al.[15] (bottom).
where the materials are microphase
separated to create an interpenetrating
network of donor and acceptor domains. Due to the
stability.[11] The general trends seem to be that fused sysshort exciton diffusion length, a domain size of approxitems are more stable while side chains especially those
mately 10 nm is believed to be the optimum.[2830] This
attached in quaternary positions or at hetero atoms
structure may fortuitously be created
when a solution of the active materials
is spin-coated or R2R processed and
later developed further by procedures
such as heat or solvent annealing.[3135]
Unfortunately, the optimum structure
for best performance is usually not at
the thermodynamic equilibrium. For
a standard type OPV composed of a
polymer and the PCBM acceptor, the
active layer is not static after the nished PV device has been prepared and
evolves over time depending on conditions such as exposure to heat.[28,29]
Phenyl-C61-butyric acid methyl ester
(PCBM) is observed to aggregate in clusters decreasing the interfacial between
donor polymer and PCBM, leading to a
decrease in photocurrent. The problems
of designing and maintaining the nanomorphology of the active layer in OPV
devices are therefore very critical to
the performance and stability and have
been the subject of many studies.

2. Discussion
2.1. Cleavable Side Chains

Figure 2. Qualitative ranking of donor (left) and acceptor (right) units for OPV polymers.
Reproduced with permission.[11] Copyright, 2011, Royal Society of Chemistry.

www.MaterialsViews.com

In the standard fabrication methods of


OPV devices, the conjugated polymers
used need to be reasonably soluble in
organic solvents, which are usually

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1547

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

Figure 3. Cleavable side chains attached to the polymer backbone. After a thermal or acid treatment, the solubilizing groups
are eliminated leaving an insoluble material consisting of the
polymer backbone.

with the side chains are avoided. Thermal elimination of


the side chains leads to a more rigid conjugated material,
characterized by its high glass transition temperature (Tg),
which has been demonstrated to strongly suppress morphological changes in high-Tg polymer:PCBM active layers
and provides high thermal stability of the photovoltaic
characteristics during long-term operation.[28,45,46] This
observation indicates that the use of a high-Tg polymer as a
donor material can effectively reduce the free movement of
the fullerene molecules within the active layer of the photovoltaic device, as long as the temperature is kept below
the Tg, as also shown previously.[28,29] The most recent
developments in this research direction have largely been
through the use of thermo-cleavable tertiary esters, where
the side chain is attached to the active conjugated polymer
backbone through a heat-labile ester bond. As illustrated in
Figure 4, thermocleavage of poly[3-(2-methyl-2-hexyl)-oxycarbonylbithiophene] (P3MHOCT) can proceed in two distinct steps with increasing temperature by initial removal
of the solubilizing tertiary substituent, leaving a carboxylic-acid-substituted thiophene (P3CT), which upon further heating to around 300 C will decarboxylate, leaving
native PT.[37,39] The thermocleaved materials, P3CT and PT,
are insoluble in all solvents and have demonstrated superior stability, compared with P3MHOCT, in photovoltaic
devices. The best performing devices are those heated to
300 C, containing native PT, which has reached efciencies up to 1.5%.
Ever since this procedure to solution-processed PT with
high stability was published, a lot of efforts have been put
into nding novel materials that combine the benecial
features of thermocleavable materials, in terms of stability, with lower bandgap donoracceptor polymers for
higher performance in OPVs. Some of the novel polymers

attained by incorporating a large number of alkyl side


chains. The ratio of carbon atoms in the side chains to
non-hydrogen atoms in the conjugated core is often similar, but once the devices have been assembled, these side
chains are actually detrimental in a number of ways, they
do not contribute to the light absorption nor to the carrier
mobility and further they decrease the stability. The side
chains are often the preferred rst point of attack of reactive oxygen species, ultimately resulting in the degradation
of the conjugated structure.[14,15,25] Several investigations
have thus focused on the possibility of removing the side
chains once the active layer has been deposited.[3638] This
has been accomplished using two different approaches
with either thermocleavable or acid-cleavable side chains
(Figure 3).[38,39]
The softness of the active layer has also been shown to
enhance the permeability of solar cells that can ease the
diffusion of small molecules such as oxygen and water,
which can cause reaction with both polymer and the electrodes.[14,18,22,4043] In addition, by using time of ight secondary ion mass spectroscopy and isotopic labeling (18O and
H218O), it has also been reported that the electrode materials
are able to diffuse through the entire layered structure to the
counter electrode in a typical polymer solar cell [i.e., indium/
tin oxide (ITO)/poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS)/active layer/metal electrode].[20] As a consequence, the solid state
of the active layer, in general, cannot be
regarded as xed, which will be a requirement to avoid morphological instability
and diffusion processes. Clearly, it is a
challenge to prepare harder polymer
materials because solubilizing side chains
are required to solution process polymers
Figure 4 . Thermolytic elimination of the side chains on P3MHOCT. At about 200 C,
into thin lms and the alkyl chains, in
the tertiary group is removed as a volatile alkene leaving the insoluble P3CT and above
300 C decarboxylation occurs leaving the native polythiophene.
addition, has a major impact on the nal
morphology of the active layer.
2.1.1. Thermocleavable Side Chains
The use of thermocleavable side chains
introduced for polythiophene (PT) polymers by Liu et al.[44] has proven to be one
of the most successful ways of increasing
the stability of conjugated polymers since
the photochemical reactions associated

1548

Figure 5. Chemical structures of novel donoracceptor polymers, bearing thermally


removable ester groups, for PSCs. eh, ethylhexyl; hx, hexyl.

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

Table 1. Optical and photovoltaic data for thermocleavable materials according to Figure 4 and 5. Solar cell data are based on a glass/ITO/
PEDOT:PSS/polymer: PC61BM/Al device geometry.

Polymer
P3MHOCT

maxa)
[nm]

Egopt
[eV]

Cleaving
temperature

Voc
[V]

Jsc
[mA cm2]

FF

b)
[%]

Ref.

511

1.99

pristine

0.73

4.90

0.40

1.27

[47]

P3CT

482

2.00

225 C

0.4 0.6

13

0.29 0.35

0.1 0.4

[39]

PTc)

482

1.97

300 C

0.75

3.26

0.41

0.84

[47]

a1

825

1.24

pristine

0.55

4.31

0.51

1.21

[48]

832

1.18

225 C

0.50

2.87

0.45

0.64

[48]

a2

608

1.66

pristine

0.69

5.68

0.35

1.37

[49]

a2c)

595

1.71

225 C

0.59

3.68

0.33

0.72

[49]

a2c)

604

1.65

265 C

0.70

4.58

0.38

1.22

[49]

a3

621

1.80

pristine

0.76

4.98

0.36

1.36

[50]

a3c)

603

1.64

225 C

0.66

5.93

0.37

1.45

[50]

c)

a1

c)

a)In

lm; b)AM1.5G corresponding to 100 mW cm2 white light; c)Thermocleaved.

that have been developed are shown in Figure 5 and the


of dithienylbenzothiadiazole and copolymerized with a
photovoltaic characteristics are given in Table 1. Copolymer
CPDT unit.[49] Placing the ester group on the thiophene unit
a1 is based on dithienylthienopyrazine, bearing thermoallows processing of three chemically different thin lms
cleavable benzoate esters on the pyrazine ring, alternating
from the same soluble precursor polymer (a2) through
with a dialkyl-substituted cyclopentadithiophene (CPDT)
elimination and decarboxylation, as seen for P3MHOCT
resulting in a very low bandgap of 1.24 eV.[48] The thermal
(Figure 4). The thermal behavior of a2 has been analyzed
with TGA and solid-state NMR, which showed that the
behavior of the polymer was investigated by thermo gravielimination of the tertiary substituent takes place around
metric analysis (TGA), which indicates that the alkyl part
200 C as expected and decarboxylation happens already at
of the ester groups is eliminated around 200 C. CO2 evolu265 C. Without thermal treatment, a2:PCBM solar cells give
tion, as measured by TGA in conjunction with mass specefciencies up to 1.37%. Upon heating the devices to 225 C,
trometry (MS) of the carrier gas (TGAMS), has not been
the performance drops to around 0.72% and then increases
reported to take place before decomposition of this polymer
again to 1.22% after heating at 265 C. P3MHOCT:PCBM
system, which initiates around 400 C. Thus, decarboxylasolar cells demonstrate the same thermal trend (Table 1)
tion of the benzoic acid group does not occur as readily as
and the authors explain it by undesirable phase segregation
observed for the carboxylic-acid-substituted polythiophene
in the active layer that commences prior to thermocleavage
P3CT (Figure 4). When utilized in bulk heterojunction solar
at 225 C. However, the extensive phase segregation, as
cells, a1 shows a moderate performance up to 1.21%
measured by atomic force microscopy (AFM), is reduced
in unannealed devices. After a thermal treatment and
drastically when annealing at 265 C (Figure 6).
thermocleavage around 225 C, the PCE drops to 0.64%,
which is mainly due to a large drop in the current together
Following this promising work, wherein thermowith minor drops in the open-circuit voltage (Voc) and ll
cleavage of the tertiary esters could be optimized in a way
so that the initial active layer morphology is not altered
factor (FF). Other donor units (i.e., dialkoxybenzene, uorene, and thiophene), besides CPDT, were
also tested with the dithienylthienopyrazine in this work, and a general observation was that the photovoltaic performance drops after elimination of the
tertiary substituent. The authors ascribe
this trend to severe macrophase separation of the polymer/PCBM domains, at
high annealing temperatures, that occur
prior to thermal cleavage of the tertiary
substituent.[48] For a2, the tertiary ester
Figure 6. AFM topography images (2 m 2 m) of a2:PCBM solar cells annealed at different
temperatures. Reproduced with permission.[49] Copyright, 2010, American Chemical Society.
was incorporated on the thiophene unit

www.MaterialsViews.com

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1549

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

Figure 7. Evolution of the normalized absorption during accelerated


photochemical ageing in air at 100 solar intensities. a3 cleaved refers
to an a3 lm that has been thermocleaved at 225 C. Reproduced
with permission.[50] Copyright 2011, Royal Society of Chemistry.

the polymer was followed during constant illumination at


100 solar intensities (Figure 7) using a lens-based sunlight
concentration setup.[55]
In terms of T50, which is the time it takes a polymer
lm to reach half of its initial absorbance, Figure 7 shows
that the stability of a3 is more than twice that of P3HT,
whereas elimination of the tertiary substituents further
increases T50 by 50%, which clearly emphasizes the unfavorable effect of solubilizing groups on the photochemical stability of conjugated polymers. To imitate the active
layer of a device, PCBM was also included in the lm more
than doubling the T50 compared with polymer only samples. The authors ascribe this effect to the effective radical
scavenging properties of PCBM that can reduce chemical
transformations of the materials induced by formed
radicals in the active layer. In terms of thermal stability,
optical microscopic images of a3/PCBM lms before and
after thermal annealing at 150 C indicate a high stability. The a3/PCBM lms show a homogeneous nature
even after annealing for 96 h (Figure 8). In stark contrast,
thermal annealing of a P3HT/PCBM lm induces the formation of many crystals, which reveal severe macrophase
separation even after only 4 h (Figure 8).

unfavorably, a3 was developed by copolymerization of


dithienylthiazolo[5,4-d]thiazole (DTZ) and silolodithiophene
(SDT).[50] For solubility, a3 has incorporated alkyl groups
on the SDT unit and thermally removable ester groups on
the DTZ unit that, as anticipated, can be eliminated around
200 C. With the use of TGAMS, decarboxylation is only
2.1.2. Thermocleavage in a Roll-to-Roll Process
observed at temperatures well in excess of 300 C. Thus, for
So far, a signicant limitation to the use of thermocleavable
solar cell applications, only the low-temperature elimimaterials in OPVs is that the cleavage temperature does
nation of the tertiary substituent is of interest because of
not comply with processing on exible plastic substrates
the device failure at higher temperatures (>300 C). The
relevant for large-scale R2R-coated polymer solar cells
SDT unit was preferred as a donor unit, as a substitute for
CPDT, because recent studies have shown
that substitution of the bridging carbon
atom with silicon can give a signicant
photochemical stability improvement of
polymers by a signicant factor[51], but
higher carrier mobility[52] and enhanced
interchain packing[53,54] have also been
reported. What is noteworthy about a3
is that the photovoltaic performance in
a3:PCBM solar cells improves after the
tertiary substituents have been eliminated from the DTZ unit by a thermal
treatment around 225 C. Hence, the initial morphology in the active layer has
not been altered unfavorably during the
thermal treatment. Generally, the opencircuit voltage (VOC) drops 0.1 V after
annealing a3:PCBM solar cells at 225 C
but together with a signicant increase
in the current, the efciency is enhanced
from 1.36 to 1.45% (Table 1). Evaluation
of the photochemical stability of a3 was
Figure 8. Optical microscopy images of a3 cleaved/PCBM and P3HT/PCBM lms before
estimated by an accelerated degradation
and after annealing at 150 C. Reproduced with permission.[50] Copyright 2011, Royal
method where the photodegradation of
Society of Chemistry.

1550

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

heating and destruction of the substrate.


Compared with using an oven, the
pulsed light can selectively heat the polymer lm, thereby keeping the temperature below the substrates-deformation
temperature (e.g., PET 140 C).
Figure 9 shows initial ash test on a3
lms where 13 pulses (power density
6 KW cm2), delivered by a commercial
photonic sintering system, were sufFigure 9. Photographic images of a3 thin lms. a) Flashed a3 lms, b) ashed a3 lm
cient to achieve thermocleavage and
immersed into ODCB where non-ashed lm areas was washed out, c) whereas ashed
high solvent resistance, as can be seen
thermocleaved areas remained on the substrate demonstrating a high solvent resist[
57
]
in Figure 9b, where the ashed polymer
ance. Reproduced with permission. Copyright 2012, Royal Society of Chemistry.
lm is immersed into ODCB. Only the
(PSCs). As examples, poly(lactic acid), poly(ethylene terephnon-ashed lm areas are washed out, whereas ashed
thalate) (PET), and poly(ethylene naphthalenate) support
thermocleaved areas remained on the substrate. The demtemperatures up to 65, 140, and 180 C, respectively. Thus,
onstration of a high solvent resistance is also revealed in
if the thermal procedure is to be transferred to exible subthe UV-vis absorption spectra before and after washing of
strates, it will be necessary to nd alternative methods to
the a3 lm (Figure 10). Only the pristine sample is washed
bring down the cleavage temperatures. Not until this has
off, whereas the changes in the spectrum were insignibeen realized can thermocleavable materials be considered
cant for the ashed polymer lm. Large-area R2R-coated
as potential candidates for large-scale production of OPV. A
solar cell modules based on a3 (active area = 35.5 cm2)
possible solution to this problem has been suggested[56,57]
were processed with an inverted device geometry on exthat has focused on the processing options and explored
ible PET substrates using the photonic sintering system
the use of light-induced heat generation (a technique
for thermocleaving of the active layer as illustrated in
dubbed ashing) where the absorption of light by a
Figure 11. The active layer lm was chemically modied
material generates heat through non-radiative energy disand insolubilized efciently and importantly enough the
sipation and exothermic photochemical reactions. For this
PET foil did not show any deformation after exposure to
purpose, the key is to use high-intensity pulsed light shortthe high-intensity light-enabling subsequent coating
duration pulses (<2 ms), with high power density, which is
steps. The performance of the nal solar cell modules
primarily absorbed in the polymer lm due to absorption
was ca. 0.5%, which is a factor of 34 lower compared
differences. The process is rapid enough to avoid excessive
with a spin-coated reference device. The authors ascribe
this to the generation of minor cracks in the ITO layer in
ashed areas that could lead to reduced charge transport
and thereby hamper the performance of the modules.
Overall, functional polymer solar cell modules could be
printed and coated using high-intensity pulsed light for
contactless chemical modication of the active layer lm,
though the authors anticipate that there will be some
restrictions to the choice/quality of active materials using
this process.
2.2. Acid-Catalyzed Cleavage

Figure 10. UV-vis absorption spectra of a3 thin lms spin-coated


on glass substrates and then ashed at 35 mm distance from the
lamp housing (power density ca. 6 kW cm2). Subsequently, the
samples were washed by immersion in ODCB for 5 min. Only the
pristine a3 sample washed off, whereas the changes in the UV-vis
were negligible for the ashed a3 lm. Reproduced with permission.[57] Copyright 2012, Royal Society of Chemistry.

www.MaterialsViews.com

The cleavage of side chains by heating is of great importance to ensure a stable morphology as described in the
sections above. However, from an energy analysis point
of view, the use of high temperatures in the production of
solar cells is not favorable,[58] increasing the energy payback time and the energy return factor thus making the
OPV less competitive with other PV technologies. Thus,
nding a method to decrease the temperature at which
the side chains are cleaved is of high importance. This can
be done by using a high-power light source (ashing) as

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1551

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

in the reaction temperature from 300


to 150 C for the formation of PT. This
temperature would not be harmful to
a PET substrate; however, the acid-catalyzed reaction was not utilized in OPV
devices. Studies where acid has been
used as a catalyst to lower the temperature have also been carried out for poly
(2-methoxy-5-(3 ,7 -dimethyloctyloxy1,4-phenylenevinylene)
(MDMO-PPV)
and poly(3-octyl-2,5-thienylenevinylene)
(O-PTV).[60] Here, the polymers were preFigure 11. (Left) Schematic drawing of the ash lamp housing mounted on a R2R system pared by both heating and acidic catalin a variable distance to the substrate. Not to scale. A = simplied R2R setup, B = high- ysis from dithiocarbamate precursors,
voltage pulse-forming generator, C = lamp housing, D = reector, E = xenon arc ash
and it was found that the temperature
lamp, F = substrate, df = focus distance, d = substrate distance. (Right) Photographic
could be lowered from 180 to 70 C.
image of a R2R process during a high-intensity light pulse. Reproduced with permis[57] Copyright 2012, Royal Society of Chemistry.
Recently, the use of silane side chains
sion.
on the PT backbone replaced the tertiary
esters (see Figure 12) and the cleavage with acid as a catadescribed above. Unfortunately, this method eventually
lyst was now carried out at room temperature.[61] Solidled to a lower photovoltaic performance for the ashed
state NMR and TGA studies showed complete removal of
roll-coated cells compared with a reference spin-coated
the silane side chains to form pristine PT at room tempercell. Therefore, an alternate method of using an acid as a
ature. This is a major improvement for both the energy
catalyst to decrease the cleaving temperature was develpayback time and the energy return factor. Similar
oped. This was demonstrated in a recent work[59] wherein
studies were carried out for polyphenylene[62] wherein it
the use of an acid as a catalyst for the cleavage of the
was found that the removal of silane side chains from the
polymer P3MHOCT to the carboxylic acid and further
polyphenylene backbone could be carried out in boiling
decarboxylation to form PT (Figure 4) resulted in a decrease
toluene. However, no reports have been published on the
use of the acid-catalyzed cleavage of side chains or acidcatalyzed precursor polymers in OPV and this work is still
in progress.

3. Morphological Stability
3.1. Block Copolymers

Figure 12. Top: cleavage reactions of silane side chains at room


temperature using acid catalyst and tertiary ester groups at
300 C.[61] Bottom: spin-coated lms of the silane polymer before
and after acid cleavage and a solvent resistance test. Reproduced
with permission. Copyright, American Chemical Society.

1552

One possible solution to control the morphology of the


active layer in OPV is to design the active material so that it
will spontaneously form the desired ordered structure. In
this approach, blocks of the donor and acceptor materials
are joined in a single-polymer strand. The electronic properties, processing, and the morphology, which the polymer
will have, are controlled by the variation in structure and
chain length.[63]
There are, in general, three different kinds of block
copolymers, the coilcoil, the coilrod, and the rodrod
block copolymers, where the coil is a exible polymer
and a rod is a rigid crystalline polymer. The copolymers
self-assemble into different morphologies depending
on different factors and off course only the morphologies allowing for charge transfer and charge transport
to the electrodes are favorable for OPV. The different
morphologies, which can be formed from coilcoil block

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

Figure 13. Nanostructures obtained for coilcoil block copolymers.


S = spheres, C = cylinders, G = gyroid, and L = lamella. The red and
blue colors represent the two polymers, A and B, respectively,
in the block copolymer. The volume fraction of polymer A is A.
Reproduced with permission.[63] Copyright 2011, Wiley-VCH.

facilitate mixing of immiscible materials such as P3HT


and PCBM, which upon annealing phase separates. The
presences of the block copolymer based on P3HT and
fullerene maintain the nanoscale mixing and ensure
high efciency of the OPV devices.[63] Recently, Miyanishi
et al.[67] described a block copolymer based on 3-hexylthiophene with controllable length with fullerene moieties attached onto the side chain of the polymers. The
copolymer was applied in OPV and an FF of 63% and a PCE
of 2.15% was obtained. The copolymer formed nanostructures of around 20 nm mainly due to the crystallization
of the thiophene block and aggregation of the fullerene
groups.[67] Another example of a block copolymer based on
P3HT was recently published by Renaud et al.,[68] who prepared a copolymer of P3HT and poly(4-vinylene-pyridine)
(P4VP), P3HT-b-P4VP, and used it in OPV as an nanostructuring agent to stabilize the nanostructure of the active
layer. It is noteworthy that the block copolymer increases
the photovoltaic parameters when applied in OPV even
without annealing at higher temperatures, which is normally used to affect the morphology of the active layer
when pristine P3HT and PCBM are used. This is ascribed
to the fact that the P4VP block interacts with the PCBM
and inhibits the formation of large PCBM domains. A
similar procedure was used by Gernigon et al.[69] who
prepared a rodcoil block copolymer and used it as a
compatibilizer. They found that the highest photovoltaic performances were achieved for a 2 wt% addition of
the block copolymer due to the lamella-like morphology
achieved upon annealing. Increasing the compatibilizer
content resulted in an increase in the number of PCBM
crystallites upon annealing giving a poorer photovoltaic
performance. In a similar study by Lee et al.,[70] a C60 endcapped P3HT block copolymer was used as a compatilizer
in a P3HT/PCBM blend and it was found that the addition
of the block copolymer prevented the formation of larger
PCBM domains upon annealing at 150 C for a long time

copolymers, are shown in Figure 13. The different morphologies depend on the FloryHuggins parameter ,[64,65]
which describes the miscibility of the two polymers in
the block copolymer, the total degree of polymerization of
the copolymer N; and the relative volume fraction, . The
different morphologies vary for different polymers, that
is to say, a certain morphology cannot be obtained for
the same value of for different polymers.[63] The morphologies S for coilcoil, where spheres of the two components are formed, are not favorable for OPV, all others
especially the lamella structure L are usable in OPV if the
electron and hole mobility are high and the domains are
aligned perpendicular to the electrodes. However, the
coilcoil copolymers cannot be used in the active layer
in OPV alone, but can be used as templates or compatibilizers (see below), this is ascribed to the low charge-carrier
mobility of the copolymer (i.e., there are no overlapping
of orbitals.)[65] For the active layer, one has to use the
coilrod or rodrod copolymers.
For coilrod and rodrod copolymers, the accessible
morphologies are very different from the coilcoil. The
rodrod copolymers have little conformational exibility.
For the rodcoil copolymer, the morphology is dominated by lamellae or liquid-crystalline structures and
not the spherical structures observed
by coilcoil copolymers. The lamellae
morphology is the most favorable for
OPV since this morphology assure an
ideal charge transfer to the electrodes
in the device.[66] In Figure 14, the morphologies for liquidcrystal rodcoils
are shown. The hockey puck morphology, Figure 14E, is observed for rod
coil copolymers with a large fraction of
coil polymer, this type of morphology,
where part of the copolymer domain
is isolated, is off course not useful for
OPV.[63]
Figure 14 . Nanostructures obtained for coilcoil block copolymers. a = nematic, b =
Block copolymers have been used
bilayer smectic A, c = mono layer smectic A, d = monolayer smectic C, e = hockey pucks.
as compatibilizers, that is to say, the
Red represents the rod and blue the coil. Reproduced with permission.[63] Copyright 2011,
block copolymer is used as additive to
Wiley-VCH.

www.MaterialsViews.com

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1553

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

Figure 15. P3HT with bromoalkyl side chains and the conceptual steps in a crosslinking reaction. Reproduced with permission.[75] Copyright
2009, Wiley-VC.

(8 days) (i.e., increasing the morphological stability of the


blend). This was mainly ascribed to the block copolymer
reducing the interfacial tension between the P3HT and
PCBM. For more details on the use of block copolymers of
polythiophene and PCBM or perylene bisimide as compatilizers in a P3HT/PCBM blend, the reader is directed
elsewhere.[71]
The use of the block copolymers as templates for a nanostructure has also been studied.[63] In this case, the block
copolymer lm is rst allowed to form ordered phases
and then one of the block components is removed leaving
voids that can then be lled with another polymer as
demonstrated by Crossland et al.[72]. A somewhat similar
approach was studied by Andreasen et al.[73] using copper
nanoparticles in a matrix of thermo cleavable P3MHOCT
polymer. A lm of the mixture was heated to cleave of the
side chains of P3MHOCT to rigidify the morphology and
make the polymer insoluble. The copper particles where
then removed with a chemical agent coordinating the
metal. Finally, the voids were lled with PCBM.
The use of block copolymers in OPV to control the morphology of the active layer is of great potential mainly
due to the great variations, which can be made with these
polymers. However, the variations are also a drawback
of block copolymers, the morphology of the active layer
and the electronic properties are highly dependent on the
structure, the chain length, and further the processing
and charge transfer. Thus, the optimal morphology is very
difcult to achieve. Besides this, the synthesis is not easy
and optimization has to be done with respect to purity
and cost.
3.2 Crosslinking
Another possibility is to freeze the morphology once it
has been achieved by crosslinking reactions. Active groups
can be incorporated into the structure, usually at the end of
the side chains that can be induced to react after the active
layer and optimal structure have been created. Several different active groups have been used for this purpose such
as oxetane,[74] alkyl-bromide,[7577] azide,[36,78] vinyl,[79] and

1554

glycidyl.[29] Kim et al.[75] demonstrated that by introducing


bromine groups to the end of the alkyl chain in P3HT
crosslinking could be achieved by UV-light (254 nm) illumination for 10 min (Figure 15). Photo-crosslinking of thin
lms with 10% bromine functionality resulted in solvent
resistance without affecting the molecular packing of the
polymer, hence retaining high charge-carrier mobility. The
bromine-functionalized P3HT polymer was further applied
in BHJ OPVs and devices subjected to photo-crosslinking
showed enhanced thermal stability when compared with
just P3HT:PCBM blends after annealing at 150 C.
The use of bromide groups as a way of inducing
polymer crosslinking has also been applied for a low
bandgap copolymer based on the octylthieno[3,4-c]pyrrole-4,6-dione (TPD) unit.[76] By introducing 16% bromide
functionality to the octyl solubilizing group of TPD, the
produced BHJ OPVs attained a PCE of 4.6% after 72 h of
thermal annealing at 150 C, with an initial PCE of 3.3%.
The increase in PCE was attributed to the consequence of
crosslinking on the -stacking, improving the electronic
properties of the polymer thus increasing the device
performance.
Miyanishi et al.[79] achieved crosslinking by using a
P3HT derivative (P3HNT) with incorporated terminal
vinyl groups on all the hexyl chains. Crosslinking
between the polymer strains could be induced by thermal
treatment of the thin lms at 150 C for 2 h and combined with PCBM, the crosslinking suppressed the formation of PCBM crystallites. The polymer crosslinking
resulted in improved thermal stability of BHJ OPV devices
when compared with P3HT:PCBM devices, due to a stabilization of the lm morphology, which resulted in an
increased voltage and a more stable current. Nam et al.[78]
used azide-substituted side chains of P3HT, which enabled the attachment of the polymer to PCBM by a cycloaddition reaction after thermal treatment of the modied
P3HT:PCBM lm. With only 15% azide functionality in
the polymer, the chemical bonding between the polymer
and PCBM-reduced phase separation and growth of PCBM
crystallites was observed when annealing at 150 C for
24 h. When applied in BHJ OPVs, the thermal stability of

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

via the miniemulsion; and 3) the use of


various nanoparticles such as quantum
dots, ZnO, TiO2, or carbon nanotubes
mixed with the polymer to create
a stable morphology.[82] The latter
approach wherein nanoparticles are
mixed with the polymer is not within
the scope of this review.
The nanoparticle grafting approach
was described by Krebs et al.[83]
Figure 16. TQ-type low bandgap polymer with 10% crosslinkable side chains containing wherein nanoparticles with P3HT
bromo, azide vinyl, or oxetane groups (left) and normalized PCE of BHJ OPVs based on TQ1, chains attached to a silica core (Si(OR) )
4
TQ-Br, TQ-N3, TQ-Vinyl, and TQ-oxetane recorded under constant illumination at ambient
of 4 nm were synthesized and applied
[
81
]
atmosphere (right). Reproduced with permission. Copyright 2012, Royal Society of
in nano-structured inks with PCBM.
Chemistry.
The inks were then used in R2R slot
die-coated OPVs, and based on the OPV
the devices was signicantly improved compared with
data, the optimal size of the silica core was discussed.
P3HT:PCBM, while still keeping the original device efSimilar organosilica nanoparticles were prepared by
ciency of more than 3%.
Senkovskyy et al.[84] where the core of the nanoparticles
Another approach is to crosslink the fullerene. Dress et
was 460 nm. The studies in the two papers showed that
al.[29] and Zhu et al.[80] have explored this using an epoxythe optimum size of the nanoparticles is of 2050 nm
functionalized fullerene, resulting in a polymerizable
represented as nanoparticle C in the Figure 17. This was
PCBM derivative (PCBG). The polymerization reaction could
ascribed to the chains being too close in a large silica
be induced by heating the thin lms at 140 C or by applinanoparticle (A in Figure 17) and to far apart in the
cation of an acid catalyst. The morphology of P3HT:PCBG
small silica nanoparticle (B in Figure 17).[85]
blends lm was stabilized by the polymerization of the
The grafting of poly(9,9-dioctyl-uorene) chains on
fullerenes compared with P3HT:PCBM blend lms. Howsilica nanoparticles was described by Tkachov et al.,[86]
ever, when the P3HT:PCBG blend was applied in BHJ OPVs,
and it was found that the high grafting density of the
a signicant degradation of the currentvoltage characteruorene chains causes the chains to adopt an ordered and
istics was observed among others due to a reduction of the
planar conformation. The use of these nanoparticles in
FF. The origin of this degradation could not be established.
OPV was not described.
Carl et al.[81] compared different types of crosslinkable
Other nanoparticles have also been described in the
chemistries in the TQ-type low bandgap polymer. Degraliterature. The grafting of functionalized P3HT chains
dation studies were carried out either with dark storage
on ZnO nanoparticles was described by Li et al.[87] and
between measurements or under constant illuminathe reported photovoltaic performances were better
tion to emphasize morphological or photochemical stathan regular P3HT/ZnO nanoparticle hybrid OPV. The
bility, respectively. It was concluded that the crosslinking
nanoparticle approach was also studied by Andersen
reaction does not affect the photochemical stability of
et al.[85] wherein silicon nanoparticles were grafted by
the polymers and that morphological stability can be
oligo(phenylvinylene) or oligo(3-hexylthiophene) chains.
increased as seen both in retarding the growth of PCBM
The reason for using silicon (SiR4) instead of silica (Si(OR)4)
domains and a slower degradation of the PCE. When these
is that for silicon, the absorption is much greater toward
materials were tested under continuous irradiation at
longer wavelengths compared with silica. This should
AM1.5 conditions, the crosslinking did not increase the
stability signicantly, which indicates that other factors
as photochemical degradation might play a more important role for the device stability (Figure 16).
3.3. Pre-structuring in Nanoparticles
The use of nanoparticles in the active layer to create
a stable morphology has been demonstrated by several groups using different nanoparticle approaches:
1) grafting of polymer chains on a silicon or silica nanoparticle; 2) creating nanoparticles of polymer in water

www.MaterialsViews.com

Figure 17. Optimum type of the nanoparticles is found to be nanoparticle C. Reproduced with permission.[85] Copyright 2012, Elsevier.

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1555

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

4. Conclusion

Figure 18. The polymer nanoparticles in water prepared by the


approach.

result in much higher efciencies. However, the authors


found that the method of grafting polymers/oligomers
onto silica nanoparticles to be difcult with difculties in
purication and the poor device performance was mainly
ascribed to pinholes and oxidation at the surface of the
nanoparticles.
Another approach toward nanoparticles was developed by Landfester and co-workers[8892] and further
developed by Andersen et al.[93] who studied the nanoparticles of P3HT and low-bandgap polymers blended
with PCBM in water in OPV. The nanoparticles were
prepared as shown in Figure 18 where the polymer was
dissolved in chloroform and then mixed with a water/
surfactant mixture to form an emulsion using sonication. Evaporation of the chloroform from the emulsion
resulted in a suspension of polymer/PCBM nanoparticles
in water. The nanoparticles of three different lowbandgap polymers were studied in R2R-processed OPV,
and it was found that the morphology of the active layer
was different for the three polymers having different
nanoparticle sizes. Further, the morphology of the R2Rcoated devices was different when compared with spincoated, perhaps explaining the lower device performance
achieved for the R2R-coated devices. The nanoparticles
in water dispersion were also used in another process
wherein the active layer consisted of polythiophene
(n-PT) derived from P3MHOCT and PCBM.[94] An emulsion of P3MHOCT/PCBM nanoparticles in water prepared
by the Landfester method given above was put through
a nebulizer to achieve vapor droplets and then heated to
vaporized the water and cleave off the side chains at the
same time forming n-PT/PCBM nanoparticles that could
be collected as a lm.
The miniemulsion approach was also used by Stapleton et al.[95] to prepare nanoparticles of polyuorene
and PCBM and the OPV showed higher photovoltaic properties compared with regular bulk heterojunctions based
on polyuorene. This was mainly ascribed to: 1) creation
of a coreshell nanoparticle structure, which enables
efcient electron and hole transport; 2) migration of the
surfactant away from the particle interface; and 3) efcient charge separation and transport by optimizing the
particle size.

1556

The design and synthesis of the active


materials for OPV remain a major scientic endeavor partially driving this
eld of research. Most of the effort is
still concentrated at achieving higher
PCE by optimizing optical and electronic
modied Landfester parameters. Achieving this goal is very
complex, involving harvesting as much
of the spectrum as possible, creating
charge carriers, and transporting them to the electrodes.
One of the important factors in most OPVs is obtaining a
microphase-separated structure the bulk heterojunction.
This can either be designed into the materials by using
block copolymers of donor and acceptor moieties or prestructuring in nanoparticles, or more commonly obtained
by post-production steps such as annealing. The latter
technique depends on the serendipitous formation of
domains of donor and acceptor with the right size distribution in a bicontinuous network. Little is actually known
about the driving forces for the process, and it is difcult to
predict and control. Once the structure has been obtained,
it is possible to arrest it by crosslinking reactions. Another
even more radical solution is to remove exible side chains
by heat or acid treatment. This reduces the mobility of the
system and presumably arrests further structural changes.
Another very important factor that has begun to attract
some attention is the degradation of OPV. Again this is a
complex issue involving many different aspects in part
due to the multi-layer structure of OPV. Organic materials exposed to sunlight and the atmosphere decomposes
in many ways and the active components of OPV are no
exception. The photostability of conjugated polymers
has been shown to be highly variable, but now the rule
of thumb has been discovered to select the most stable
materials.
If OPV is to succeed outside the laboratory, all of this
knowledge on photochemical and morphological stability
together with optimal light harvesting and carrier generation and transport has to be engineered into the active
materials. Despite the huge effort already dedicated,
there are still many challenges to overcome.

Acknowledgements: This work was supported by the Danish


National Research Foundation and by the Danish Strategic
Research Council (DSF 2104-05-0052 and 2104-07-0022).

Received: February 5, 2013; Published online: June 28, 2013; DOI:


10.1002/macp.201300076

Keywords: block copolymers; crosslinking; degradation; highperformance polymers; OPV

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

Macromolecular
Chemistry and Physics

Advanced Functional Polymers for Increasing the Stability of Organic Photovoltaics

www.mcp-journal.de

[1] Z. He, C. Zhong, S. Su, M. Xu, H. Wu, Y. Cao, Nat. Photonics


2012, 6, 593.
[2] Z. M. Beiley, M. D. McGehee, Energy Environ. Sci. 2012, 5,
9173.
[3] Z. He, C. Zhong, X. Huang, W.-Y. Wong, H. Wu, L. Chen, S. Su,
Y. Cao, Adv. Mater. 2011, 23, 4636.
[4] M. Jrgensen, K. Norrman, S. A. Gevorgyan, T. Tromholt,
B. Andreasen, F. C. Krebs, Adv. Mater. 2012, 24, 580.
[5] N. Grossiord, J. M. Kroon, R. Andriessen, P. W. M. Blom, Org.
Electron. 2012, 13, 432.
[6] M. Jrgensen, K. Norrman, F. C. Krebs, Sol. Energy Mater. Sol.
Cells 2008, 92, 686.
[7] R. Sndergaard, M. Hsel, D. Angmo, T. T. Larsen-Olsen,
F. C. Krebs, Mater. Today 2012, 15, 36.
[8] Y. Galagan, I. G. de Vries, A. P. Langen, R. Andriessen,
W. J. H. Verhees, S. C. Veenstra, J. M. Kroon, Chem. Eng. Processing 2011, 50, 454.
[9] F. C. Krebs, T. Tromholt, M. Jrgensen, Nanoscale 2010, 2,
873.
[10] A. Distler, P. Kutka, T. Sauermann, H. J. Egelhaaf, D. M. Guldi,
D. Di Nuzzo, S. C. J. Meskers, R. A. J. Janssen, Chem. Mater.
2012, 24, 4397.
[11] M. Manceau, E. Bundgaard, J. E. Carl, O. Hagemann,
M. Helgesen, R. Sndergaard, M. Jrgensen, F. C. Krebs,
J. Mater. Chem. 2011, 21, 4132.
[12] M. Manceau, A. Rivaton, J. L. Gardette, S. Guillerez,
N. Lematre, Sol. Energy Mater. Sol. Cells 2011, 95, 1315.
[13] M. Manceau, A. Rivaton, J. L. Gardette, S. Guillerez,
N. Lematre, Polym. Degrad. Stab. 2009, 94, 898.
[14] S. Chambon, M. Manceau, M. Firon, S. Cros, A. Rivaton,
J. L. Gardette, Polymer 2008, 49, 3288.
[15] R. Grisorio, G. Allegretta, P. Mastrorilli, G. P. Suranna, Macromolecules 2011, 44, 7977.
[16] M. Hermenau, M. Riede, K. Leo, S. A. Gevorgyan, F. C. Krebs,
K. Norrman, Sol. Energy Mater. Sol. Cells 2011, 95, 1268.
[17] M. V. Madsen, K. Norrman, F. C. Krebs, Proc. SPIE 2010,
7777.
[18] K. Norrman, F. C. Krebs, Sol. Energy Mater. Sol. Cells 2006, 90,
213.
[19] K. Norrman, F. C. Krebs, Surf. Interface Anal. 2004, 36, 1542.
[20] F. C. Krebs, K. Norrman, Prog. Photovoltaics Res. Appl. 2007,
15, 697.
[21] S. Chambon, A. Rivaton, J. L. Gardette, M. Firon, L. Lutsen, J.
Polym. Sci., Part A: Polym. Chem. 2007, 45, 317.
[22] M. Manceau, A. Rivaton, J. L. Gardette, Macromol. Rapid
Commun. 2008, 29, 1823.
[23] H. Hintz, H. J. Egelhaaf, H. Peisert, T. Chasse, Polym. Degrad.
Stab. 2010, 95, 818.
[24] T. Tromholt, M. V. Madsen, J. E. Carl, M. Helgesen, F. C. Krebs,
J. Mater. Chem. 2012, 22, 7592.
[25] M. Manceau, J. Gaume, A. s. Rivaton, J. L. Gardette, G. Monier,
L. Bideux, Thin Solid Films 2010, 518, 7113.
[26] B. C. Thompson, J. M. J. Frechet, Angew. Chem. Int. Ed. 2008,
47, 58.
[27] G. Li, R. Zhu, Y. Yang, Nat. Photonics 2012, 6, 153.
[28] S. Bertho, G. Janssen, T. J. Cleij, B. Conings, W. Moons,
A. Gadisa, J. DHaen, E. Goovaerts, L. Lutsen, J. Manca,
D. Vanderzande, Sol. Energy Mater. Sol. Cells 2008, 92, 753.
[29] M. Drees, H. Hoppe, C. Winder, H. Neugebauer, N. S. Sariciftci,
W. Schwinger, F. Schfer, C. Topf, M. C. Scharber, Z. Zhu,
R. Gaudiana, J. Mater. Chem. 2005, 15, 5158.
[30] J. S. Moon, J. K. Lee, S. Cho, J. Byun, A. J. Heeger, Nano Lett.
2009, 9, 230.
[31] F. Padinger, R. S. Rittberger, N. S. Sariciftci, Adv. Funct. Mater.
2003, 13, 85.

www.MaterialsViews.com

[32] W. L. Ma, C. Y. Yang, X. Gong, K. Lee, A. J. Heeger, Adv. Funct.


Mater. 2005, 15, 1617.
[33] G. Li, V. Shrotriya, Y. Yao, Y. Yang, J. Appl. Phys. 2005, 98,
043704.
[34] M. Campoy-Quiles, T. Ferenczi, T. Agostinelli, P. G. Etchegoin,
Y. Kim, T. D. Anthopoulos, P. N. Stavrinou, D. D. C. Bradley,
J. Nelson, Nat. Mater. 2008, 7, 158.
[35] Y. Zhao, Z. Y. Xie, Y. Qu, Y. H. Geng, L. X. Wang, Appl. Phys.
Lett. 2007, 90, 043504.
[36] F. C. Krebs, H. Spanggaard, Chem. Mater. 2005, 17, 5235.
[37] M. Bjerring, J. S. Nielsen, A. Siu, N. C. Nielsen, F. C. Krebs, Sol.
Energy Mater. Sol. Cells 2008, 92, 772.
[38] S. A. Gevorgyan, F. C. Krebs, Chem. Mater. 2008, 20, 4386.
[39] X. Han, X. W. Chen, S. Holdcroft, Adv. Mater. 2007, 19, 1697.
[40] H. J. Kim, C. H. Cho, H. Kang, H. H. Cho, B. J. Kim, A. R. Han,
M. Y. Lee, J. H. Oh, J. M. J. Frchet, Chem. Mater. 2012, 24, 215.
[41] M. Lira-Cantu, K. Norrman, J. W. Andreasen, F. C. Krebs,
Chem. Mater. 2006, 18, 5684.
[42] K. Norrman, N. B. Larsen, F. C. Krebs, Sol. Energy Mater. Sol.
Cells 2006, 90, 2793.
[43] S. Chambon, A. Rivaton, J. L. Gardette, M. Firon, Sol. Energy
Mater. Sol. Cells 2008, 92, 785.
[44] J. S. Liu, E. N. Kadnikova, Y. X. Liu, M. D. McGehee,
J. M. J. Frchet, J. Am. Chem. Soc. 2004, 126, 9486.
[45] J. Vandenbergh, B. Conings, S. Bertho, J. Kesters,
D. Spoltore, S. Esiner, J. Zhao, G. Van Assche, M. M. Wienk,
W. Maes, L. Lutsen, B. Van Mele, R. A. J. Janssen, J. Manca,
D. J. M. Vanderzande, Macromolecules 2011, 44, 8470.
[46] J. Zhao, S. Bertho, J. Vandenbergh, G. Van Assche, J. Manca,
D. Vanderzande, X. Yin, J. Shi, T. Cleij, L. Lutsen, B. Van Mele,
Phys. Chem. Chem. Phys. 2011, 13, 12285.
[47] T. Tromholt, S. A. Gevorgyan, M. Jrgensen, F. C. Krebs,
K. O. Sylvester-Hvid, ACS Appl. Mater. Interfaces 2009, 1,
2768.
[48] M. Helgesen, F. C. Krebs, Macromolecules 2010, 43, 1253.
[49] M. Helgesen, M. Bjerring, N. C. Nielsen, F. C. Krebs, Chem.
Mater. 2010, 22, 5617.
[50] M. Helgesen, M. V. Madsen, B. Andreasen, T. Tromholt,
J. W. Andreasen, F. C. Krebs, Polym. Chem. 2011, 2, 2536.
[51] M. Helgesen, M. Manceau, F. C. Krebs, T. J. Srensen, Polym.
Chem. 2011, 2, 1355.
[52] J. H. Hou, H. Y. Chen, S. Q. Zhang, G. Li, Y. Yang, J. Am. Chem.
Soc. 2008, 130, 16144.
[53] H. Y. Chen, J. Hou, A. E. Hayden, H. Yang, K. N. Houk, Y. Yang,
Adv. Mater. 2010, 22, 371.
[54] M. Morana, H. Azimi, G. Dennler, H.-J. Egelhaaf, M. Scharber,
K. Forberich, J. Hauch, R. Gaudiana, D. Waller, Z. Zhu,
K. Hingerl, S. S. van Bavel, J. Loos, C. J. Brabec, Adv. Funct.
Mater. 2010, 20, 1180.
[55] T. Tromholt, M. Manceau, M. Helgesen, J. E. Carl, F. C. Krebs,
Sol. Energy Mater. Sol. Cells 2011, 95, 1308.
[56] F. C. Krebs, K. Norrman, ACS Appl. Mater. Interfaces 2010, 2,
877.
[57] M. Helgesen, J. E. Carl, B. Andreasen, M. Hsel, K. Norrman,
R. Sndergaard, F. C. Krebs, Polym. Chem. 2012, 3, 2649.
[58] N. Espinosa, M. Hsel, D. Angmo, F. C. Krebs, Energy Env. Sci.
2012, 5, 5117.
[59] R. Sndergaard, K. Norrman, F. C. Krebs, J. Polym. Sci., Part A:
Polym. Chem. 2012, 50, 1127.
[60] H. Dilin, J. Vandenbergh, F. Banishoeb, P. Adriaensens,
T. J. Cleij, L. Lutsen, D. J. M. Vanderzande, Macromolecules
2011, 44, 711.
[61] E. Bundgaard, O. Hagemann, M. Bjerring, N. C. Nielsen,
J. W. Andreasen, B. Andreasen, F. C. Krebs, Macromolecules
2012, 45, 3644.

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1557

Macromolecular
Chemistry and Physics

E. Bundgaard et al.

www.mcp-journal.de

[62] S. Jakob, A. Moreno, X. Zhang, L. Bertschi, P. Smith,


A. D. Schlter, J. Sakamoto, Macromolecules 2010, 43,
7916.
[63] P. D. Topham, A. J. Parnell, R. C. Hiorns, J. Polym. Sci., Part B:
Polym. Phys. 2011, 49, 1131.
[64] P. J. Flory, Principles of Polymer Chemistry, Cornell University
Press, Ithaca, NY, USA 1953.
[65] P.-G. Gennes, Scaling Principles in Polymer Physics, Cornell
University Press, Ithaca, NY, USA 1979.
[66] P. T. Wu, G. Ren, C. Li, R. Mezzenga, S. A. Jenekhe, Macromolecules 2009, 42, 2317.
[67] S. Miyanishi, Y. Zhang, K. Hashimoto, K. Tajima, Macromolecules 2012, 45, 6424.
[68] C. Renaud, S.-J. Mougnier, E. Pavlopoulou, C. Brochon,
G. Fleury, D. Deribew, G. Portale, E. Cloutet, S. Chambon,
L. Vignau, G. Hadziioannou, Adv. Mater. 2012, 24, 2196.
[69] V. Gernigon, P. Lvque, C. Brochon, J.-N. Audinot, N. Leclerc,
R. Bechara, F. Richard, T. Heiser, G. Hadziioannou, Eur. Phys.
J. Appl. Phys. 2011, 56, 34107.
[70] J. U. Lee, J. W. Jung, T. Emrick, T. P. Russell, W. H. Jo, J. Mater.
Chem. 2010, 20, 3287.
[71] M. Sommer, S. Huettner, M. Thelakkat, J. Mater. Chem. 2010,
20, 10788.
[72] E. J. W. Crossland, M. Kamperman, M. Nedelcu, C. Ducati,
U. Wiesner, D. -M. Smilgies, G. E. S. Toombes, M. A. Hillmyer,
S. Ludwigs, U. Steiner, H. J. Snaith, Nano Lett. 2009, 9, 2807.
[73] J. W. Andreasen, M. Jrgensen, F. C. Krebs, Macromolecules
2007, 40, 7758.
[74] J. Farinhas, Q. Ferreira, R. E. Di Paolo, L. Alcer, J. Morgado,
A. Charas, J. Mater. Chem. 2011, 21, 12511.
[75] B. J. Kim, Y. Miyamoto, B. Ma, J. M. J. Frchet, Adv. Funct.
Mater. 2009, 19, 2273.
[76] G. Grifni, J. D. Douglas, C. Piliego, T. W. Holcombe, S. Turri,
J. M. J. Frchet, J. L. Mynar, Adv. Mater. 2011, 23, 1660.
[77] K. Yao, L. Chen, T. Hu, Y. Chen, Org. Electron. 2012, 13,
1443.
[78] C. Y. Nam, Y. Qin, Y. S. Park, H. Hlaing, X. Lu, B. M. Ocko,
C. T. Black, R. B. Grubbs, Macromolecules 2012, 45, 2338.
[79] S. Miyanishi, K. Tajima, K. Hashimoto, Macromolecules 2009,
42, 1610.

1558

[80] Z. Zhu, S. Hadjikyriacou, D. Waller, R. Gaudiana, Z. Zhu, J.


Macromol. Sci., Pure Appl. Chem. 2004, 41 A, 1467.
[81] J. E. Carl, B. Andreasen, T. Tromholt, M. V. Madsen,
K. Norrman, M. Jrgensen, F. C. Krebs, J. Mater. Chem. 2012,
22, 24417.
[82] B. R. Saunders, M. L. Turner, Adv. Colloid Interface Sci. 2008,
138, 1.
[83] F. C. Krebs, V. Senkovskyy, A. Kiriy, IEEE J. Selected Topics
Quantum Electron. 2010, 16.
[84] V. Senkovskyy, R. Tkachov, T. Beryozkina, H. Komber,
U. Oertel, M. Horecha, V. Bocharova, M. Stamm,
S. A. Gevorgyan, F. C. Krebs, A. Kiriy, J. Am. Chem. Soc. 2009,
131, 16445.
[85] T. R. Andersen, Q. Yan, T. T. Larsen-Olsen, R. Sndergaard,
Q. Li, B. Andreasen, K. Norrman, M. Jrgensen, W. Yue, D. Yu,
F. C. Krebs, H. Chen, E. Bundgaard, Polymer 2012, 4, 1242.
[86] R. Tkachov, V. Senkovskyy, M. Horecha, U. Oertel, M. Stamm,
A. Kiriy, Chem. Commun. 2010, 46, 1425.
[87] F. Li, Y. Du, Y. Chen, L. Chen, J. Zhao, P. Wang, Sol. Energy
Mater. Sol. Cells 2012, 97, 64.
[88] T. Kietzke, D. Neher, M. Kumke, R. Montenegro, K. Landfester,
U. Scherf, Macromolecules 2004, 37, 4882.
[89] T. Kietzke, D. Neher, K. Landfester, R. Montenegro,
R. Guntner, U. Scherf, Nat. Mater. 2003, 2, 408.
[90] K. Landfester, Adv. Mater. 2001, 13, 765.
[91] T. Piok S. Gamerith, C. Gadermaier, H. Plank, F.P. Wenzl,
S. Patil, R. Montenegro, T. Kietzke, D. Neher, U. Scherf,
K. Landfester, E. J. W. List, Adv. Mater. 2003, 15, 800.
[92] M. Antonietti, K. Landfester, Prog. Polym. Sci. 2002, 27, 689.
[93] T. R. Andersen, T. Larsen-Olsen, B. Andreasen, A. P. L. Bttiger,
J. E. Carl, M. Helgesen, Eva Bundgaard, K. Norrman,
J. W. Andreasen, M. Jrgensen, F. C. Krebs, ACS Nano 2011, 5,
4188.
[94] Y. X. Nan, X. L. Hu, T. T. Larsen-Olsen, B. Andreasen,
T. Tromholt, J. W. Andreasen, D. M. Tanenbaum, H. Z. Chen,
F. C. Krebs, Nanotechnology 2011, 22, 475301.
[95] A. Stapleton, B. Vaughan, B. Xue, E. Sesa, K. Burke, X. Zhou,
G. Bryant, O. Werzer, A. Nelson, A. L. D. Kilcoyne, L. Thomsen,
E. Wanless, W. Belcher, P. Dastoor, Sol. Energy Mater. Sol. Cells
2012, 102, 114.

Macromol. Chem. Phys. 2013, 214, 15461558


2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.MaterialsViews.com

You might also like