You are on page 1of 8

Solar Energy Materials & Solar Cells 116 (2013) 262–269

Contents lists available at SciVerse ScienceDirect

Solar Energy Materials & Solar Cells


journal homepage: www.elsevier.com/locate/solmat

The influence of the nematic phase on the phase separation of blended


organic semiconductors for photovoltaics
Steven A. Myers a,n, Manea S. Al Kalifah a,1, Chunghong Lei a, Mary O'Neill a,
Stuart P. Kitney b, Stephen M. Kelly b
a
Department of Physics and Mathematics, University of Hull, Hull HU6 7RX, UK
b
Department of Chemistry, University of Hull, Hull HU6 7RX, UK

art ic l e i nf o a b s t r a c t

Article history: Differential scanning calorimetry in combination with atomic force microscopy is used to examine the
Received 28 November 2012 phase separation of a blended nematic liquid crystalline electron-donor and crystalline perylene electron-
Received in revised form acceptor mixture. Separate domains of donor and acceptor material are mostly retained in the blend,
25 April 2013
although a small proportion of the acceptor, increasing with increasing donor concentration, is mixed in
Accepted 7 May 2013
Available online 6 June 2013
with the donor domains. Annealing in the nematic phase allows the donor and acceptor molecules to move
and generate phase-separated domains of the required size, thus enhancing the performance of bulk
Keywords: heterojunction photovoltaic devices based on these blends. We show that the optimum annealing
Photovoltaics temperature can be controlled by manipulation of the temperature range of the nematic phase of the
Liquid crystals
donor.
Nematic
& 2013 Elsevier B.V. All rights reserved.
Self-assembly
Nano-morphology
Phase transitions

1. Introduction within the individual domains of electron donors and electron


acceptors. Thioenothiophene polymers, which have high-
Columnar, smectic and, more recently, nematic liquid crystals temperature liquid crystalline phases, were blended with soluble
have been shown to be very promising charge-transporting fullerene derivatives [7,8], with optimized devices having a power
organic semiconductors due to their ability to spontaneously conversion efficiency of 2.5%. It is noted that annealing in the
self-assemble into highly ordered domains in uniform, thin films liquid crystal phase improved the performance of photovoltaic
over large areas [1–3]. This combination of properties allied with devices incorporating donors with either columnar or lamellar
broad absorption spectra render them particularly suitable as polymeric phases [6,9,10]. A nematic liquid crystal composite with
active materials for organic photovoltaics, such as bulk hetero- a porous surface and sub-micron scaled grooves has also been
junction devices based on blends of liquid crystalline electron used to provide a distributed interface to vertically separate
donors and crystalline electron acceptors. Photogenerated excitons electron-donating and electron-accepting films in a bilayer PV
are dissociated at the interface between the phase-separated device [11,12]. However, quite surprisingly, the use in OPVs of low-
domains of the donor and acceptor components. The separated molar-mass nematic liquid crystals (small molecules) has not been
electrons and holes then travel along different and distinct con- studied to any significant degree [13]. The nematic phase of such
tinuous pathways in opposite directions to the electrodes. Photo- compounds possesses a much lower viscosity than that of either
voltaic cells comprising a coronene compound as a donor with highly ordered columnar liquid crystals or high molecular weight
columnar liquid crystalline phases blended with an electron liquid crystalline polymers, which are exceptionally viscous mate-
acceptor showed power conversion efficiencies up to 1.5% under rials. Therefore, nematic liquid crystals should offer significant
standard measurement conditions [4–6]. The performance of these advantages in controlling the morphology of the donor–acceptor
organic photovoltaic (OPV) cells was optimized by effective use of composites for organic photovoltaics, especially when using thermal
thermal annealing to control the degree of molecular ordering annealing to control domain size and morphology.
This work addresses the important question of how the self-
n
assembly properties of liquid crystals can be exploited to control
Corresponding author. Tel.: +44 1482465396.
E-mail address: S.A.Myers@hull.ac.uk (S.A. Myers).
the morphology of the donor–acceptor composites for organic
1
Current Address: Department of Physics, Faculty of Science and Art at Alrass, photovoltaics. Our model system consists of a nematic donor
Qassim University, Kingdom of Saudi Arabia. blended with a crystalline perylene based acceptor. We propose

0927-0248/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.solmat.2013.05.006
S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269 263

a combination of macroscopic and nanoscale techniques, differ- Poly(3,4-ethylenedioxythiophene) (PEDOT) doped with poly(styr-
ential scanning calorimetry and atomic force microscopy, as an enesulfonate) (PSS) supplied by Baytron (P AI4083) was sonicated
effective method to investigate how the liquid crystalline phase is for 5 min, spin cast onto the ITO substrate at 4000 rpm for 30 s and
modified by blending and whether the donors and acceptors are then annealed at 120 1C for 30 min followed by 5 min at 220 1C.
intimately mixed or occupy distinct domains. We examine why The active layer was spincast from 1.5% by weight solution of the
annealing in the nematic phase improves morphology and opti- donor and acceptor compounds mixed in various mole ratios in
mises PV performance. We choose two distinct donor–acceptor chlorobenzene at 2000 rpm for 30 s. This was followed by anneal-
blends, where both LC donors are electrically and optically very ing on a hot-plate under various conditions as described later. The
similar but have significant different LC temperature ranges, to cathode consisted of a thin layer (∼0.6 nm) of lithium fluoride (LiF)
confirm the relevance of the LC phase. and a layer (100 nm) of aluminium deposited by thermal evapora-
tion under vacuum. The completed PV device was mounted in a
vacuum sealed test chamber for characterisation outside a glove
2. Experimental box. The devices were illuminated with a Xenon lamp (Bausch &
Lomb), dispersed through a monochromator and attenuated with
Table 1 shows the chemical structure of the materials used in neutral light filters over an area of 0.25 cm2. The current–voltage
this investigation. The synthesis of the electron-donor materials characteristics of the photovoltaic devices were measured in an
1 and 2 is described in the Supplementary information. The inert atmosphere using a Visual-Basic controlled picoammeter.
synthesis of perylene acceptor 3 has been reported previously
[12]. Cyclic voltammetry and absorption spectroscopy are used to
obtain the ionization potential and electron affinity of the materi- 3. Results and discussion
als, as discussed in the supplementary information.
A Perkin-Elmer differential scanning calorimeter (DSC 7) was 3.1. Phases of binary donor–acceptor mixtures
used to measure the phase transition temperatures of the pure
materials and blends, which are summarized in Table 2. Blends The ionization potential and electron affinity of liquid crystal-
were prepared by dissolving either donor 1 or donor 2 along with line compounds 1 and 2 are the same (5.52 eV and 3.15 eV,
the appropriate amount of acceptor 3 in dichloromethane as a respectively) within experimental error. Both of them act as
solvent and then drop-casting the resultant solution into a electron donors, when blended with the crystalline electron
standard aluminium differential scanning calorimetry (DSC) pan acceptor 3, see Table 1, which has an electron affinity of 4.19 eV.
(Perkin-Elmer) using a pipette. Around 4–5 mg of material was Fig. 1 shows the DSC traces for the donors 1 and 2 and the acceptor
used for each DSC sample. The pan was prepared in this way in 3, the blend comprising 2 and 3 as well as the blend 1 and 3 in a
order to simulate the blend mixture produced during thin film 1:3 mol ratio. The transition temperatures of these compounds
processing. Donor 1 and acceptor 3 were mixed in a 1:3 mol ratio, and their blends, as well as those of other blends of 2 and 3 with
whilst donor 2 and acceptor 3 were mixed in mole ratios of 4:3, different mole ratios, are summarised in Table 2.
1:1, 1:2 and 1:3. A standard indium sample was used to calibrate
the DSC prior to all measurements. The scan rate was 10 1C min−1
and two heating–cooling cycles were performed for each blend. Table 2
The morphology of the blended thin films was investigated with The glass transition temperature (tg), melting point (Cr–N and Cr–I), clearing point
an Atomic Force Microscope (AFM) developed by Molecular (N–I) and recrystallisation temperature (I–Cr) for each of the four blends studied.
The fifth column 5 records the enthalpy of melting (ΔE, per gram) of the acceptor
Imaging (Agilent) in the tapping mode. MAC Levers Type II molecule 3.
cantilevers (Agilent) were used for all measurements. Picoscan
V.5.2 was used to produce the images, RMS roughness and height Material Blend molar tg Cr–N Cr–I ΔE N–I I–Cr
distribution data. Fast Fourier Transforms (FFT) and the power ratio (D:A) (1C) (1C) (1C) (J/g of 3) (1C) (1C)
spectral density data (PSD) were obtained using Gwyddion free-
1 80 218
ware software. We investigated areas of 5  5 mm2 and 1  1 mm2, 2 94 180 303
each area being composed of 1024  1024 data points. 3 – – 274 19.1 – 256
Soda lime glass substrates were provided by UQG optics with a 2 and 3 1:3 93 180 256 16.4 – 209
100 nm thick indium tin oxide (ITO) layer, giving a sheet resistance 1:2 93 180 257 11.5 – 206
1:1 94 181 250 10.4 – 181
of ≤20 Ω/sq. Acid etching was used to remove strips of ITO from
4:3 94 180 250 8.7 – 180
the edges to allow contact to the cathode. The substrates 1 and 3 1:3 80 – 265 10.1 206 225
were cleaned by sonification and treated with oxygen plasma.

Table 1
The chemical structure of the liquid crystalline electron-donor materials 1 and 2 and the crystalline perylene acceptor 3.

3
264 S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269

26 22
Cr N N I
24 Tg

Heat Flow (mW)

Heat Flow (mW)


Tg N I 21
22
20
20
19
18 I N 1st Cycle
2nd Cycle
16 18
0 50 100 150 200 250 300 0 50 100 150 200 250 300 350
Temperature (°C) Sample Temperature (°C)
40
24 Cr I
35 Tgof 2 Cr of 3
N of 2
Heat Flow (mW)

Heat Flow (mW)


30 Cr I 22
25
20
20
18
15 I Cr 1st Cycle
10 16 2nd Cycle
I C of 3
0 50 100 150 200 250 300 -50 0 50 100 150 200 250 300 350
Sample Temperature (°C) Sample Temperature (°C)

26
1st Cycle Cr I of 3
2nd Cycle
24
Heat Flow (mW)

Tg of 1 N I of 1
22

20
Crystallisation of 3

18
I N of 1 I Cr of 3
0 50 100 150 200 250 300
Sample Temperature (°C)
Fig. 1. DSC scans for compounds (a) 1, (b) 2, (c) 3, blends (d) 2 and 3 and (e) 1 and 3 in a 1:3 mol ratio.

The DSC data shown in Fig. 1a indicate that the donor 1 is a temperature decreases only slightly for blends with higher con-
nematic glass at room temperature. It has a glass transition at centrations of donor material 2 and most probably corresponds to
80 1C on heating and a nematic–isotropic phase transition at the melting of the acceptor component 3 containing traces of the
218 1C. Donor 2 is crystalline at room temperature and melts to donor 2. On cooling of the 1:3 blend, both components of the
form a nematic phase at 180 1C as illustrated in Fig. 1b. The mixture remain isotropic until 209 1C, when the isotropic–crystal
clearing point (N–I) is substantially higher (+85 1C) than that of (I–Cr) phase transition, i.e., recrystallisation, is observed. The
1, see Table 1. The nematic phase is formed again on cooling degree of super-cooling observed depends on the mole ratio of
2 from the isotropic liquid followed by the formation of a glassy the mixture. These results suggest that most of donor 2 is present
phase (vitrification) on further cooling significantly below the as phase-separated nematic domains in the blend alongside
melting point. As a consequence a nematic glass transition is crystalline domains of the acceptor 3. However, the phase separa-
observed at 94 1C in the second heating cycle of 2. Acceptor tion of the two components is not complete as shown by examin-
3 exhibits no observable mesophase as shown in Fig. 1c. Com- ing the enthalpy change per gram of acceptor 3 integrated over the
pounds 2 and 3 were mixed together in four different mole ratios melting points for the different blends. As the concentration of
(4:3, 1:1, 1:2 and 1:3). Fig. 1d shows that the 1:3 mole ratio blend donor increases in the samples, proportionally less acceptor is
of donor 2 and acceptor 3 has a crystal–nematic melting transition incorporated into the distinct phase-separated domains. This
at 180 1C in the first heating cycle, most probably attributable to implies that small, but varying, fractions of donor and acceptor
melting of compound 2 present in the blend. This melting are intimately mixed.
transition is replaced by a weak glass transition at temperature The first heating and cooling cycle of the DSC scan of the blend
93 1C in the second heating cycle, implying that the nematic phase containing donor 1 and acceptor 3 in the mole ratio 1:3, see Fig. 1e,
has vitrified on cooling significantly below the melting point of shows distinct transitions for each material. The glass transition
both 2 and 3. Table 2 shows that the glass transition temperature temperature and clearing point of the blend are very similar to
(tg) and melting point (Cr–N) during heating, are fairly constant for those of the pure donor 1 material. However, the melting point
all mole ratios of the blends comprising components 2 and 3 and and the recrystallisation transition of the blend are similar, but a
that they correlate closely with those of 2. A double-peaked little lower, than those of the acceptor 3. However, we note that
crystal–isotropic transition (Cr–I) is observed on heating in both the enthalpy of the recrystallisation transition is only 68% of that
cycles at about 256 1C, see Fig. 1d. As Table 2 shows, this at the melting point, suggesting that a significant amount of
S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269 265

3 remains mixed with 1 on cooling. This is confirmed in the second Table 4


cycle where an exothermic recrystallisation transition is observed Data from bulk heterojunction PV devices obtained on excitation with 22 mW cm−2
at 466 nm. The blended film contained donor 1 and acceptor 3 in a 1:3 ratio by
on heating at 239 1C, which is probably attributable to the acceptor
mole. Different annealing temperatures were used.
3 partially re-crystallising out of the blend. Thus, the DSC mea-
surements suggest that mostly separate phases of donor and Device Max. anneal. temp. η EQE Jsc Voc FF
acceptor are retained in the blend. number (1C) (%) (%) (A cm−2) (V)

D9 – 0.4 5.2 0.436 0.95 0.21


3.2. Photovoltaic devices D10 50 0.32 4.6 0.381 0.85 0.22
D11 100 0.65 7.1 0.592 0.95 0.26
D12 120 1.1 8.3 0.708 1.0 0.36
Fig. 2 shows the current–voltage characteristics for a bulk D13 150 0.65 6.6 0.54 0.95 0.28
heterojunction device containing a blend of 1 and 3 in the mole D14 200 0.19 2.6 0.21 0.85 0.24
ratio 1:3. The devices were irradiated with light of wavelength
466 nm and of different irradiances as shown in the figure.
Photovoltaic devices were fabricated using the liquid crystalline
compound 2 as the donor and the crystalline perylene derivative
1.0 120C
3 as the acceptor, mixed in different ratios as collated in
200C
Tables 3 and 4, which also tabulate the power conversion effi-
ciency (η), external quantum efficiency (EQE), short circuit current 0.8
density (Jsc), open circuit voltage (Voc) and fill factor (FF) for the
devices on irradiation with 24 mW cm−2 at a wavelength of

 (%)
0.6
462 nm. The maximum value of Voc approaches the difference
equal to 1.34 eV between the LUMO energy of the acceptor and the 0.4
HOMO energy of the donor. The fill factor is poor for all the
devices, but improves on annealing at 200 1C. The results show
0.2
that the performance of devices is better for blends with a greater
percentage of the acceptor 3 in the blend. Table S1 in the
Supporting information shows that the performance deteriorates 0.0
0.5 1.0 1.5 2.0 2.5 3.0
for blend ratios of 2 and 3 greater than 3:1. Similar trends were
found for the blend of 1 and 3, see Table S2. It is also improved by
Blend Ratio Acceptor/Donor (Mol)
Fig. 3. The η of bulk heterojunction photovoltaic devices as a function of the mole
ratio of acceptor 3 to the donor 2 in the blended film annealed at either 120 1C or
200 1C.
22 mW cm-2
6.0x10-4
14.7 mW cm-2
9.28 mW cm-2 annealing at 200 1C as summarized in Fig. 3, which shows η as a
4.0x10-4 function of blend composition and annealing temperature. As
2.55 mW cm-2
Current (A)

Fig. 1d shows, the donor component of the blend is in the


Dark crystalline state at 120 1C and in the nematic phase at 200 1C,
2.0x10-4 suggesting that annealing in the (fluid) nematic phase is beneficial
for the device performance.
0.0 Based on the simulation of Peumans et al.[14], who studied the
effect of annealing at increasing temperatures on the morphology
of a donor–acceptor blend, it is assumed that the surface topo-
-2.0x10-4 graphy can give an insight into the bulk morphology and that the
lateral features of the surface roughness correlate with the donor–
-0.5 0.0 0.5 1.0 1.5 2.0 acceptor domains. Fig. 4 shows 2D atomic force microscopy images
Voltage (V) of all eight films. The main images are 5  5 mm2 with the
Fig. 2. The current versus voltage of the photovoltaic device incorporating a thin
1  1 mm2 image inset. It is clear from the images for each blend
film of a 1:3 blend of 1 and 3. The inset labels the irradiance in mW cm−2 of the ratio that the roughness increases significantly when the sample is
input light source of wavelength 466 nm. annealed at 200 1C, see Fig. 5. For the samples annealed at 120 1C,
the 1:2 ratio blend exhibits the maximum rms roughness ampli-
tude (equal to 1.6 nm). Annealing at 200 1C increases the rough-
Table 3 ness for all the mixtures up to a maximum value of 2.6 nm for the
Data from bulk heterojunction PV devices obtained on excitation with 24 mW cm−2 blend with the largest acceptor concentration. Fig. 4 also shows a
at 462 nm. Different mole ratios of donor 2 and acceptor 3 and different annealing clear difference in the texture of the samples annealed at different
temperatures are used.
temperatures. The films annealed at 120 1C show phase separation
Device D:A mole Max. anneal. η EQE Jsc Voc FF on a significantly finer spatial scale compared with those annealed
number ratio temp. (1C) (%) (%) (A cm−2) (V) at 200 1C, where many of the grains are elongated and apparently
crystalline, with well defined edges, particularly when the accep-
D1 4:3 120 0.04 0.75 6.7  10−5 0.50 0.31
tor concentration is high. The narrow diameter of the grain is of
D3 1:1 120 0.11 1.07 9.6  10−5 0.87 0.30
D5 1:2 120 0.30 2.72 2.4  10−4 0.91 0.32 the order 50 nm, which is relatively large compared with the
D7 1:3 120 0.39 3.73 3.4  10−4 0.94 0.30 expected exciton diffusion length of about 10 nm.
D2 4:3 200 0.32 2.77 2.5  10−4 0.88 0.35 It is clear from the analysis of the blend of 2 and 3 that annealing
D4 1:1 200 0.49 3.40 3.1  10−4 1.06 0.36 in the nematic phase promotes phase separation of mixtures of
D6 1:2 200 0.86 5.34 4.8  10−4 1.20 0.36
D8 1:3 200 0.98 6.46 5.8  10−4 1.17 0.35
donor and acceptor materials formed as a result of deposition from
solution by spin casting and subsequent evaporation of the solvent.
266 S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269

Fig. 4. AFM topography image for thin films of donor 2 and acceptor 3 blended in the mole ratio (a) 4:3 mol, (b) 1:1 mol, (c) 1:2 mol, and (d) 1:3 mol. The samples on the left
and right were annealed at 120 1C and at 200 1C respectively. The main image size is 5 mm  5 mm with 1 mm  1 mm inset.

We now investigate the significance of annealing in the nematic to have little effect whilst an annealing temperature of 120 1C
phase for the blended materials 1 and 3, which have very different appears to be optimal for PV device performance, which is
transition temperatures, correlating morphology and photovoltaic substantially inferior at higher annealing temperatures. Table 2
performance. shows the same trends for EQE, Jsc, Voc and the fill factor. Fig. 1e
Fig. 6 shows that the power conversion efficiency of devices (DSC) suggests that this particular blend is a nematic glass below
D9–D14 based on blends of 1 and 3 mixed in a 1:3 mol ratio is very 108 1C, and then forms a nematic phase up to the clearing point of
sensitive to the annealing temperature. Annealing at 50 1C appears 206 1C. Some of the dissolved acceptor crystallizes out at higher
S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269 267

3 3.3. Discussion
1μm 120°C
RMS Roughness (nm) 1μm 200°C Despite the very different temperature ranges of the nematic
phases, the best photovoltaic performance is found for the blends
when the bulk heterojunction donor–acceptor films are annealed
2 in the nematic phase. This is particularly surprising for the blend
of 2 and 3, since annealing in the nematic phase also results in
high surface roughness, which is normally associated with poor
photovoltaic performance [16,17]. The DSC for both blends sug-
gests that the donor and acceptor components retain their
1 individual phase behaviour to a large extent in the blended film.
Annealing in the fluid nematic phase then allows the material to
move to phase-separated regions of improved size. Other authors
find improved photovoltaic performance following annealing in
1 2 3
liquid crystalline phases, for example, Sun et al. [18] reported
Blend Mole Ratio (Acceptor/Donor) increases of 200–400% in the Jsc using PCBM with liquid crystalline
Fig. 5. RMS roughness measured from the 1 mm  1 mm image of each sample porphyrins. The improvement in efficiency was attributed to
shown in Fig. 4. thermally induced alignment of the porphyrins in the columnar
discotic phase. This alignment provides efficient charge transport
along the columnar axis and optimized light harvesting. Recently
an increase in efficiency from 0.22% to 0.65% was reported for a
1.2 solar cell containing a blend of poly(3-hexylthiophene) mixed
1.0 with a liquid crystalline fullerene derivative, N-methyl-2-{4-[6-(4′-
cyanobiphenyl-4-oxy)hexyloxy]phenyl}-3,4-fulleropyrrolidine,
0.8 when annealing above the liquid crystalline transition tempera-
 (%)

ture. [19].
0.6 The intensity dependence of the photocurrent density can be
expressed by Jsc ¼AIinα, where Iin is the incident optical irradiance
0.4
and A and α are constants; α¼1 implies monomolecular recombi-
0.2 nation whereas the recombination of nongeminate electrons and
holes is bimolecular with α¼0.5. Fig. 9 shows α as a function of the
50 100 150 200 annealing temperature for both blend systems with the donor:
T (°C) acceptor mole ratio 1:3. It shows that annealing at the optimum
Fig. 6. η of bulk heterojunction photovoltaic devices as a function of annealing
temperature produces photovoltaic devices, which have the high-
temperature. Blends of donor 1 and acceptor 3 in the mole ratio 1:3 were used in est values of α, showing mono-molecular recombination. At
the devices. temperatures below the nematic phase the molecules are not
sufficiently mobile in the vitrified glassy state to move from their
temperatures. Fig. 7 shows the surface morphology of the films on non-ideal positions reached on evaporation of the solvent from the
annealing at 50 1C, 120 1C and 200 1C. The rms roughness increases spin-coated film. The domain size is small, and possibly not well
with temperature and is equal to 0.5 nm, 1.0 nm and 1.7 nm for the phase-separated, increasing the probability of recombination of
three films studied. The AFM image shown in Fig. 7c clearly shows the free electrons and holes. This result agrees with a recent study
that the acceptor has crystallized out at 200 1C, indicating that the that found the optimum domain size for poly(3-hexylthiophene):
nematic clearing point is lower than that determined for a bulk PCBM blends to be around 6 nm, with respect to photovoltaic
sample using the DSC. This is not surprising as phase transition performance [20]. Domains smaller than this were found to result
temperatures tend to be lower in thin films than in the bulk [15]. in a poorly percolated domain structure and hence increase the
Fig. 8 shows the normalized power spectral density for the three probability of charge recombination. For the blends studied,
films. The plot is split into four regions to aid analysis of the bimolecular recombination also becomes more probable on the
domain size. The lowest frequency (i) relates to domain sizes formation of large acceptor domains following high temperature
larger than 100 nm, the second section (ii) relates to domain sizes annealing of the blended 1 and 3.
from 20 nm to 100 nm, the third (iii) relates to domains from 5 nm Another key result is that the device performance of the blend
to 20 nm and the highest frequency range (iv) indicates domains of donor 2 and acceptor 3 is best for the highest concentration of
smaller than 5 nm. Ideally, a morphology with domain sizes in the acceptor in the blend, which is well in excess of an equal mole
region (iii) will allow the majority of excitons to reach a donor– ratio with the donor. We have found this to be a general result
acceptor interface within the assumed 10 nm diffusion length. found for other bulk heterojunction devices involving a nematic
Domain sizes smaller than 5 nm, whilst good for charge separa- liquid crystalline donor combined with the acceptor 3. It may be
tion, may not allow efficient charge transport. These results show associated with more complete phase separation of the donor and
that annealing in the nematic phase at 120 1C allows optimal acceptor, which also occurs with greater acceptor concentration as
phase separation by creating more domains of the required size. indicated by DSC in donor rich samples; proportionally less
Larger domains are obtained by annealing at 200 1C, whilst acceptor is incorporated into the distinct phase separated
annealing at low temperatures, i.e., in the nematic glassy state, domains, resulting in poorer photovoltaic devices.
gives a larger fraction of smaller domains. AFM images of samples
which were annealed in the nematic phase at 150 1C show a
similar morphology to that obtained at 120 1C. However, the rms 4. Conclusion
roughness of the former sample is significantly higher, equal to
1.7 nm, which may account for the inferior performance of the We show that differential scanning calorimetry combined with
corresponding photovoltaic device. atomic force microscopy is an extremely effective method to
268 S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269

Fig. 7. AFM topography image for thin films of donor 1 and acceptor 3 blended in the mole ratio 1:3 on annealing at (a) 50 1C, (b) 120 1C and (c) 200 1C. The main image size
is 5 mm  5 mm with a 1 mm  1 mm inset.

1.0
Blend: 1 and 3
1 Blend: 2 and 3
50C
0.9
0.1 120C
200C
Normalized PSD

Slope α

0.01 0.8
1E-3
0.7
1E-4
(i) (ii) (iii) (iv)
1E-5 <100nm <20nm 0.6
>100nm <5nm
>20nm >5nm
1E-6 0 50 100 150 200
0.01 0.1 1
T (°C)
Spatial frequency (nm)-1
Fig. 9. Variation of α, defined in text, with annealing temperature for devices based
Fig. 8. Radially integrated power spectral density obtained using Fourier analysis of on blended thin films of donor and acceptors, mixed in the mole ratio 1:3; α¼ 1
a 2-dimensional topography image of the surface of films of the blended materials implies monomolecular recombination whereas the recombination of nongeminate
1 and 3 annealed for the temperatures stated in the inset. electrons and holes is bimolecular with α ¼0.5.
S.A. Myers et al. / Solar Energy Materials & Solar Cells 116 (2013) 262–269 269

examine the phase separation of blended donors and acceptors for hexabenzocoronenes in organic field-effect transistors and solar cells,
bulk heterojunction photovoltaic devices. Such studies would be Advanced Functional Materials 20 (2010) 927–928.
[7] I.W. Hwang, J.Y. Kim, S. Cho, J. Yuen, N. Coates, K. Lee, M. Heeney,
usefully extended to more commonly investigated blends invol- I. McCulloch, D. Moses, A.J. Heeger, Bulk heterojunction materials composed
ving crystalline polymeric donors and fullerene acceptors. Our of poly(2,5-bis(3- tetradecylthiophen-2-yl)thieno[3,2-b;]thiophene): ultra-
results also highlight the usefulness of the nematic phase in fast electron transfer and carrier recombination, Journal of Physical Chem-
istry C 112 (2008) 7853–7857.
controlling phase separation. As demonstrated here, the temperature [8] N.C. Cates, R. Gysel, Z. Beiley, C.E. Miller, M.F. Toney, M. Heeney, L. McCulloch,
range of the nematic phase can be varied independently of the M.D. McGehee, Tuning the properties of polymer bulk heterojunction solar
electronic properties of the molecule. Hence, effective low tempera- cells by adjusting fullerene size to control intercalation, Nano Letters 9 (2009)
4153–4157.
ture annealing of blends could be achieved by the incorporation of [9] J.P. Schmidtke, R.H. Friend, M. Kastler, K. Müllen, Control of morphology in
molecules with low temperature nematic phases. These molecules efficient photovoltaic diodes from discotic liquid crystals, Journal of Chemical
could also be polymerizable so that the optimized morphology is Physics 124 (2006).
[10] K. Yao, Y. Chen, L. Chen, F. Li, X. Li, X. Ren, H. Wang, T. Liu, Mesogens mediated
subsequently locked in by crosslinking.
self-assembly in applications of bulk heterojunction solar cells based on a
conjugated polymer with narrow band gap, Macromolecules 44 (2011)
2698–2706.
Appendix A. Supporting material [11] M. Carrasco-Orozco, W.C. Tsoi, M. O'Neill, M.P. Aldred, P. Vlachos, S.M. Kelly,
New photovoltaic concept: liquid-crystal solar cells using a nematic gel
template, Advanced Materials, 18 (2006) 1754-1758.-+.
Supplementary data associated with this article can be found in [12] W.C. Tsoi, M. O'Neill, M.P. Aldred, S.P. Kitney, P. Vlachos, S.M. Kelly, Dis-
the online version at http://dx.doi.org/10.1016/j.solmat.2013.05.006. tributed bilayer photovoltaics based on nematic liquid crystal polymer
networks, Chemistry of Materials 19 (2007) 5475–5484.
[13] C. Lei, M.S. Al Khalifah, M. O'Neill, M.P. Aldred, S.P. Kitney, P. Vlachos,
S.M. Kelly, Calamitic liquid crystal blends for organic photovoltaics, in:
References Proceedings of SPIE—The International Society for Optical Engineering, 2008.
[14] P. Peumans, S. Uchida, S.R. Forrest, Efficient bulk heterojunction photovoltaic
cells using small-molecular-weight organic thin films, Nature 425 (2003)
[1] M. Funahashi, J.I. Hanna, High carrier mobility up to 0.1 cm2 V−1 s−1 at 158–162.
ambient temperatures in thiophene-based smectic liquid crystals, Advanced [15] R.A.L. Jones, The dynamics of thin polymer films, Current Opinion in Colloid
Materials 17 (2005) 594–598. and Interface Science 4 (1999) 153–158.
[2] A.M. Van De Craats, J.M. Warman, A. Fechtenkötter, J.D. Brand, M.A. Harbison, [16] S.H. Lee, D.H. Kim, J.H. Kim, G.S. Lee, J.G. Park, Effect of metal-reflection and
K. Müllen, Record charge carrier mobility in a room-temperature discotic surface-roughness properties on power-conversion efficiency for polymer
liquid-crystalline derivative of hexabenzocoronene, Advanced Materials photovoltaic cells, Journal of Physical Chemistry C 113 (2009) 21915–21920.
11 (1999) 1469–1472. [17] H.C. Hesse, J. Weickert, M. Al-Hussein, L. Dassel, X. Feng, K. Müllen,
[3] K.L. Woon, M.P. Aldred, P. Vlachos, G.H. Mehl, T. Stirner, S.M. Kelly, M. O'Neill, L. Schmidt-Mende, Discotic materials for organic solar cells: effects of
Electronic charge transport in extended nematic liquid crystals, Chemistry of chemical structure on assembly and performance, Solar Energy Materials
Materials 18 (2006) 2311–2317. and Solar Cells 94 (2010) 560–567.
[4] L. Schmidt-Mende, A. Fechtenkötter, K. Müllen, E. Moons, R.H. Friend, [18] Q. Sun, L. Dai, X. Zhou, L. Li, Q. Li, Bilayer and bulk heterojunction solar cells
J.D. MacKenzie, Self-organized discotic liquid crystals for high-efficiency using liquid crystalline porphyrins as donors by solution processing, Applied
organic photovoltaics, Science 293 (2001) 1119–1122. Physics Letters 91 (2007) 253505.
[5] J. Li, M. Kastler, W. Pisula, J.W.F. Robertson, D. Wasserfallen, A.C. Grimsdale, [19] P. Wang, K. Yao, L. Chen, Y. Chen, F. Li, H. Wang, S. Yu, Self-assembled
J. Wu, K. Müllen, Organic bulk-heterojunction photovoltaics based on alkyl mesogens modified fullerene for efficiently stable bulk heterojunction solar
substituted discotics, Advanced Functional Materials 17 (2007) 2528–2533. cells, Solar Energy Materials and Solar Cells 97 (2012) 34–42.
[6] W.W.H. Wong, T. Birendra Singh, D. Vak, W. Pisula, C. Yan, X. Feng, [20] Y.M. Nam, H. Huh, W.H. Jo, Optimization of thickness and morphology of
E.L. Williams, K.L. Chan, Q. Mao, D.J. Jones, M. Chang-Qi, K. Müllen, active layer for high performance of bulk-heterojunction organic solar cells,
P. BÃuerle, A.B. Holmes, Solution processable fluorenyl hexa-peri- Solar Energy Materials and Solar Cells 94 (2010) 1118–1124.

You might also like