Sei sulla pagina 1di 10

Solar Energy Materials & Solar Cells 120 (2014) 675–684

Contents lists available at ScienceDirect

Solar Energy Materials & Solar Cells


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

Inter-diffused ordered bulk heterojunction organic


photovoltaics: optimized morphology for efficient
exciton dissociation and charge transport
Buyoung Jung a,1, Kangmin Kim a,1, Jungwon Kim a, Sehwan Kim b,
Eunkyoung Kim b, Woochul Kim a,n
a
School of Mechanical Engineering, Yonsei University, 50 Yonsei-ro, Seodaemun-gu, Seoul 120-749, Republic of Korea
b
Department of Chemical and Biomolecular Engineering, Yonsei University, 50 Yonsei-ro, Seodaemun-gu, Seoul 120-749, Republic of Korea

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

Article history: Effective control of the morphology can enhance the performance of organic photovoltaics.
Received 15 August 2013 The morphology of an active layer needs to have a large interfacial area between donors and acceptors
Received in revised form for efficient exciton dissociation and continuous, direct charge transport paths to electrodes for high
28 September 2013
charge transport efficiency. These two requirements are usually contradictory. Here, we propose a
Accepted 21 October 2013
Available online 10 November 2013
morphology that meets these requirements nearly simultaneously, called an inter-diffused ordered bulk
heterojunction (IDOBHJ). This novel structure exhibited 9% higher performance based on the Monte Carlo
Keywords: simulation than an optimized disordered bulk heterojunction. The main reasons for superior perfor-
Organic photovoltaics (OPVs) mance were attributed to the comparable short circuit current density and a higher fill factor. We also
Monte Carlo method
implemented the IDOBHJ morphology by the experimental nanoimprint technique combined with
Morphology
thermal annealing process to confirm simulation results. Our experimental results are indeed consistent
with the theoretical analysis.
& 2013 Elsevier B.V. All rights reserved.

1. Introduction critical for high EQE, i.e., power conversion efficiency (PCE).
The OPVs can be categorized by various morphologies, but, in this
Organic photovoltaics (OPVs) have been gaining attention due study, we will focus on disordered bulk heterojunction (DBHJ) and
to their light weight, flexibility, and low manufacturing cost [1–4]. ordered bulk heterojunction (OBHJ).
The performance of OPVs is proportional to the integration of The DBHJ possesses a large interfacial area, which leads to high
external quantum efficiency (EQE) versus photon wavelength, also exciton diffusion efficiency. When the interfacial area between a
called the incident photon to current efficiency (IPCE). The EQE at donor and acceptor is large, the exciton diffusion to the interface is
certain photon wavelengths, λ, is calculated from the ratio of light efficient; the generated Frenkel excitons, which possess limited
absorption (ηabs), exciton diffusion (ηextD), charge transfer (ηchrT), diffusion length [11,12], can only be dissociated into the interfacial
and charge collection efficiency (ηchrC) [5–8]. area. However, when the interfacial area is large, the charge
EQ EðλÞ ¼ ηabs ðλÞ  ηextD ðλÞ  ηchrT ðλÞ  ηchrC ðλÞ ð1Þ transport path to the electrodes is generally long and non-straight,
which reduces the charge collection efficiency. Thus, most DBHJ
As shown in Eq. (1), a high EQE is required for high light studies have focused on morphology control to ensure efficient
absorption, exciton diffusion, charge transfer, and charge collec- charge colleciton while maintaining high exciton diffusion [13–18].
tion. We let ηchrT ¼ 1 due to the rapidity (  10–100 dfs) of the So, various computational simulation including Monte Carlo
process in organic donor–acceptor photovoltaics [8]. However, method [8,19–25] and drift-diffusion modeling [26,27] are con-
other parameters are usually interdependent, wherein increasing ducted to reveal the correlation of morphology and performance.
one parameter decreases the others [9]. While the light absorption Maturová et al. [26] calculated current–voltage characteristics on
mostly depends on the active layer thickness, both the exciton different degrees of nanoscale phase separation in MDMO-PPV:
diffusion and charge transport show strong dependency on the PCBM system. In the modeling, they observed that the short-
active layer morphology [10]. Thus, the active layer morphology is circuit current is enhanced in finer phase separation due to a
reduction in bimolecular recombination caused by lateral move-
n
Corresponding author. Tel.: þ 82221235816.
ment of photo generated electrons to the fullerene-rich phase.
E-mail address: woochul@yonsei.ac.kr (W. Kim). However, at high bias, vertical electron transport is enhanced and
1
These authors contributed equally. lateral movement is reduced, causing a significant field-dependent

0927-0248/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.solmat.2013.10.019
676 B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684

carrier extraction for coarse morphologies. In their other paper to the interfacial area per unit volume of 0.48 to 0.13 nm, 1
[28], they concluded that the BHJ should have a phase separated respectively.
morphology with feature sizes that are smaller than  50 nm for The morphologies of OBHJ were constructed with pillar sizes of 18,
best performance. All of these simulations represent that, through 30 and 45 nm. Intervals between pillars are set to be the same as the
morphology control on DBHJ, charge collection can be enhanced. respective pillar sizes. In the IDOBHJ morphologies, we chose to insert
Also, there are various experimental morphology control methods DBHJ morphologies into the OBHJ morphologies. This is because of the
to achieve this aim, [29–32] such as the thermal annealing reasons as follows; we suspect that inter-diffusion occurs during the
[13,33,34], solvent annealing, [14] slow cooling by controlling the post-annealing process mainly because of the very rapid diffusion of
spin coating time [17], using a solvent additive [35], and solid fullerene derivatives into the amorphous polymer region [47,48] along
additives as block-copolymer [36,37]. While these methods with accompanying the polymer-chains-crystallization [41]. However,
enhance the charge transport efficiency of DBHJ [28,38,39], we this cannot be captured well enough by the existing morphology
found that the charge transport efficiency of DBHJ is still less than generation methods such as the Ising–Hamiltonian method [20,43] or
0.8 yet that of OBHJ is close to 1.0 (see Fig. 3). the modified Cahn–Hilliard method [19,24,25]. Also, although it is
Coakley et al. suggested the use of OBHJ to overcome the poor known that phase separation in organic photovoltaics is usually a
charge collection efficiency of DBHJ [40]. Unlike DBHJ, charges are consequence of solvent evaporation and thermal annealing [25], the
collected through straight pathways in the OBHJ. Additionally, relevant modeling on evaporation driven processes is still at an early
Aryal et al. reported that the vertical chain alignment of P3HT phase [49].
formed in the OBHJ can enhance the charge-carrier mobility [41]. Once the morphology was generated, light absorption was calcu-
This argument was supported by Zhou et al. [42], who reported lated. Analysis on the light absorption was conducted by solving
that P3HT nanogratings showed 60 times greater hole mobility Maxwell's equation using the finite differential time domain method
than a non-optimized shape. On the other hand, interfacial area for a glass/ITO/PEDOT:PSS/active layer/Al [50,51]. Uniform light absor-
per unit volume between donors and acceptors of OBHJ, which ption everywhere in the active layer was not assumed as it is not valid
directly influencing to the exciton diffusion efficiency, is generally for OBHJ. The location of the light absorption is very important
an order of magnitude smaller than that of DBHJ, which will be because it determines the position of the exciton generation and
shown later in this paper (see Fig. 3). influences the exciton dissociation. If light is absorbed far from the
In this study, we propose a morphology satisfying both the donor/acceptor interface, generated excitons typically recombine due
previously contradictory high exciton diffusion and high charge to their limited diffusion length of approximately 10 nm [11,12].
collection efficiency. We called this morphology the inter-diffused Alternatively, if light is absorbed near the donor/acceptor interface,
ordered bulk heterojunction (IDOBHJ). This morphology hybri- the possibility of dissociation is high. In the simulation, excitons were
dized advantages of both DBHJ and OBHJ; high exciton diffusion generated with the rate of 1000 s  1 nm  2 corresponding to 1 sun
efficiency and high charge collection efficiency. A schematic of the condition, i.e., 100 mW/cm2.
IDOBHJ structure is shown in Fig. 1. Both morphologies of DBHJ The lattice-based kinetic Monte Carlo (MC) method which traces
(Fig. 1(a)) and OBHJ (Fig. 1(b)) coexist in the IDOBHJ (Fig. 1(c)); movement of each charge or exciton has been developed to
straight pillars, i.e., the morphology of OBHJ, for which the width is characterize photovoltaic performance in a device. Nelson et al.
not greater than the exciton diffusion length, are observed. Near [52] initially applied the MC method to simulate charge transport
the pillar structures, the morphology of DBHJ, i.e., the inter- in MDMO-PPV/PCBM system. They found that charge recombination
diffused layer, are present as shown in Fig. 1(d). In our theoretical is slow enough not to hinder short circuit current and claimed that
analysis, we confirmed that this IDOBHJ can achieve higher power space charge effects may be the reason for observing dependency of
conversion efficiency over those of OBHJ and DBHJ. Also, we photocurrent on temperature. One of the main advantages of the MC
presented experimental proofs supporting this analysis. method is that this method can be used even under complex
morphology. Watkins et al. [20] utilized this fact and studied the
dependence of the internal quantum efficiency (IQE) of an organic
bulk heterojunction solar cell on the device morphology. They used
2. Theoretical analysis and experimental demonstration the Kawasaki spin-exchange dynamics to create various DBHJ
morphologies. They found that the IQE is indeed a strong function
2.1. Theoretical analysis of the morphology and that OBHJ possesses 1.5 times higher IQE than
that of DBHJ. Marsh et al. [43] updated the Monte Carlo modeling by
Morphology can be simulated in a few different ways such as including dark injection, so their modeling can present the full
Ising–Hamiltonian method [20,43], modified Cahn–Hilliard method current–voltage characteristics including short-circuit current,
[19,24,25], etc. The Ising–Hamiltonian method [20,43] was used for open-circuit voltage, and fill factor. They examined that the mor-
generating DBHJ morphologies. In DBHJ, two different materials, phology, light intensity, charge mobility, and recombination rate are
which were initially mixed randomly in an active layer, were com- key parameters on OPVs performance. Interestingly, they have
bined with the same materials and these evolved to larger domains. shown that a tenfold increase in mobility produces a double increase
Donor and acceptor sites were equally generated in the 60  60  30 in the maximum power output in a bilayer device. Meng et al. [23]
sites (x, y, z) of a Cartesian coordinate. The length of each site was compared the MC simulation results to experimental data and found
3 nm [44], and only one material was assigned to a site. According to that the fill factors are identical. They suggested a feature size has to
the swap probabilities, two materials could swap, and these swap be around  10 nm to achieve an optimal energy conversion effi-
events occurred at neighboring site pairs. By repeating the swapping ciency. Later, they improved the MC method by incorporating the
event, DBHJ morphologies were generated. This method is consistent Poisson equation to consider a nonuniform electrostatic potential
with experimental data showing that domain size grows proportion- that depends on the charge distribution [53]. The I–V curve based on
ally to thermal annealing time [45,46]. For example, Lei et al. [21] this modeling showed good agreement with experimental results.
demonstrated the consistency of the Ising–Hamiltonian method with Recently, Heiber et al. [54] further improved the exciton dissociation
the P3HT/PCBM system. The domain size is defined as the distance at mechanism in DMC modeling by including exciton delocalization and
which the pair distribution probability, the ratio of the same material hot charge separation effect. They found that both delocalized
located in the distance, becomes 0.5 [25]. The domain sizes of exciton and hot charge separation effect could reduce recombination
generated DBHJ are evaluated from 3 to 14 nm which corresponds so these effects should be considered. Kipp et al. [55] updated the
B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684 677

Fig. 1. Schematics of the morphologies simulated in this study: (a) disordered bulk heterojunction (DBHJ), (b) ordered bulk heterojunction (OBHJ) and (c) inter-diffused
ordered bulk heterojunction (IDOBHJ). The DBHJ structure has large interfacial areas, which is beneficial for exciton dissociation. The OBHJ has direct charge pathways, which
enhance the charge transport efficiency. The IDOBHJ structure possesses benefits of both DBHJ and OBHJ. Schematic (d) shows a route to fabricate IDOBHJ from OBHJ by the
inter-diffusion process.

conventional charge injection mechanism of previous MC method by Table 1


including the charge injection rate prefactor. From this treatment, the The parameters used in theoretical analysis.
electrode grid-size, electrode charge density, and injection timescales
could be corrected for manage the dark current. Parameters[23] Values in the simulation
The DMC modeling combined with the first reaction method
Polymer dielectric constant (ε) 3.5
(FRM) was used in this report to simulate excitons and charges Electric constant (ε0) 8.854  10  12 F m  1
transport. The detailed simulation procedure is described in else- Temperature (T) 298 K
where [8,20,22,23] so we present only brief explanation here. All Active layer thickness (L) 90 nm
the events explained below are occurred in sequence according to Boltzmann constant (kB) 8.6  10  5 eV K  1
Lattice constant (a0) 3 nm
the shortest execution time, τq,—the first reaction method (FRM). Forster radius (R0) 3.16 nm
1 Exciton generation rate 1,000 sec  1 nm  2
τq ¼  lnðXÞ ð2Þ Charge transfer rate (WchrT) 5 ps  1
W Exciton recombination rate (Wexr) 0.002 ps  1
here, X is a random number (0oXo1) and W is the rate of an event. Exciton lifetime (τe) 500 ps
Charge recombination rate[75] 0.925  10  6 ps  1
The event with the shortest execution time is executed first and this
Charge injection barrier (EIB) 0.5 eV
time becomes the iteration time in the whole system. The execution Electron mobility[76] (μe) 8  10  3 cm2/V-s
time of all carriers in a region corresponding to the Coulombic radius, Hole mobility[77] (μh) 5.6  10  3 cm2/V-s
i.e., 15 nm in this study, by the previous executed event are recalcu- Gaussian standard deviation (s) 0.062 eV
lated and accumulated to the event list. Parameters used in the Twice the polaronic binding energy (Er) 0.187 eV
Electrode work function difference (ΔΦ) 0.5 eV
simulation are presented in the Table 1. An exciton is diffused accor-
ding to the exciton hopping rate, Wexh;
  rate, Wexr. When a free charge is generated, the charge hops with the
1 R0 6
W exh ¼ ð3Þ hopping rate, Wchh, i.e., Marcus theory,
τe Rij
!
where τe, Ro and Rij are exciton lifetime, Förster radius and distance ðEj  Ei þ Er Þ2
W chh ¼ V hop exp  ð4Þ
between the original and destination sites, i and j respectively. If the 4Er kB T
exciton meets an interface between a donor and an acceptor, it is
dissociated into a free electron and a hole with the charge transfer where Ei and Ej denote energies of initial site i and possible hopping
rate, WchrT. Otherwise it recombines with the exciton recombination site j respectively. Er, kB, and T is twice the polaronic binding energy,
678 B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684

Boltzmann constant and absolute temperature (298 K). The prefactor, smallest pillar size in 25 nm did outperform the blend a little bit.
Vhop, comes from the Einstein relationship [20,43] which is We suspect that the degree of inter-diffusion could be different in the
  both cases, which should affect the device performance. In the P3HT/
6kB Tμe=h Er
V hop ¼ 2
exp ð5Þ PCBM [57] case, the morphology they got could be close to the
qa0 4kB T
morphology of OBHJ but they could get the morphology of IDOBHJ in
here, μe/h and a0 is electorn (or hole) mobility and the lattice constant. the P3HT/ P8TBT [58]. The degree of inter-diffusion is a function of
The site energies (Ei, Ej) reflect electric field influenced by the electrode materials, annealing temperature, annealing time, etc. So, the post
work function difference (Δϕ), external voltage (F) and the Coulombic annealing temperature of 120 1C may be enough for the P3HT and
interaction between charges. Here, qi and qj are the site charges. P8TBT to be inter-diffused each other but may not be the sufficient
temperature for the P3HT and PCBM. In our case, we raised the post
ðΔϕ  qFÞ qi qj 1
Ei ¼ xþ∑ ð6Þ annealing temperature to 150 1C and confirmed with the secondary
L Rij 4πε 0 ε Rij
ion mass spectroscopy (SIMS) data that inter-diffusion did occur nicely
here, L, x, ε0, ε and qi/j are an active layer thickness, a distance from an between P3HT and PCBM. Details in the fabrication step are presented
electrode, vacuum permittivity, dielectric constant and site charges, as follows. A patterned indium–tin-oxide (ITO) glass was sonicated in
respectively. Also, the effect of dark current has been considered since detergent, DI water, acetone and IPA sequentially. The ITO glass was
charges are generated not only by the exciton dissociation but also then dried with N2 gas stream and treated with UV ozone cleaner.
from the charge injection from electrodes. The energy barrier, U, After cleaning, the ITO glass was moved into a glove box to spin-coat a
between the metallic electrode and active layer is set to be, buffer and a donor layer. Poly (3,4-ethylenedioxythiophene): poly
(styrenesulfonate) (PEDOT:PSS, Clevios P VP AL 4083) was spin-coated
q2 ðΔϕ  qFÞ on the ITO as a buffer layer at 3000 rpm yielding 40 nm thickness and
U ¼ EIB  þ x ð7Þ
16πε0 εa0 L dried on a hotplate at 140 1C during 10 min. The 100 nm thickness of
here, charge injection barrier, EIB, corresponds to the energy difference poly-3-hexylthiophene (rr-P3HT, 96%, Rieke Metals) dissolved in 1,2-
between the Fermi level of Al and the LUMO of PCBM. The second and dichlorobenzene (20 mg L  1) was spin-coated on the buffer layer. The
third term consider effects of the image charge and external electric P3HT layer was then nano-imprinted by a pre-patterned PDMS mold
field. The charge injection follows the Miller–Abrahams equation, immediately (see Fig. 2(a)). The mold can transfer 100 nm patterns in
8   diameter to the P3HT donor. Scanning electron micrographs show
Ej  Ei
6kB Tμ < exp  kB T : ðEj 4Ei Þ patterned P3HT (Fig. 2(b)). The stamped P3HT/PDMS was cured by
W chi ¼  ð8Þ
qa20 : 1 : ðEj r Ei Þ thermal annealing for 60 min at 150 1C. C60 (99.5%, Sigma-Aldrich)
was purified by the gradient sublimation method [59] for three times
Finally, charges are either collected to the electrodes or recom- before loading at thermal evaporator. C60 was then thermally
bined with other charges. evaporated with 100 nm thick on the pattered P3HT as an acceptor
layer. Especially, the evaporation was chosen rather than the spin
2.2. Device fabrication and characterization coating because the solvent based spin coating method is known to
disrupt the patterned P3HT layer [60]. Subsequently, LiF and Al were
We have tried to realize this IDOBHJ experimentally by the follo- deposited on top of the C60 layer for a metal electrode. The post
wing procedure; once a donor layer was deposited onto a substrate, it thermal annealing (PA) at 150 1C was expected to induce inter-
was patterned by nanoimprinting [56,57] to form pillar structures. diffusion of the two materials, i.e., donor and acceptor, so the mor-
Then, an acceptor layer was deposited onto the patterned donor layer. phology of IDOBHJ was constructed (Fig. 2(c)).
Finally, thermal annealing at 150 1C caused both the donor and The J–V characteristics of the OPVs were measured with an
acceptor to inter-diffuse, as shown in Fig. 1(d) [16,48]. He et al. have AM1.5G filtered solar simulator (SAN-EI, XES-301S) at 1 sun
fabricated OBHJ out of P3HT/P8TBT [58] and P3HT/PCBM [57] using condition (100 mW/cm2) and a source meter unit (SMU, Keitheley
the ‘double nanoimprint technique'. In both papers they did perform 2400). To measure inter-diffusion rate, we synthesized C60D36
the post-annealing at 120 1C; inter-diffusion of donors acceptors may and counted deuterium, i.e., D, using the time-of-flight secondary
have occurred. However, they achieved contradictory results—The PCE ion mass spectrometry (TOF-SIMS, ION-TOF). X-ray diffractometry
of P3HT/P8TBT OBHJ did outperform than that of P3HT/P8TBT DBHJ (XRD, Rigaku) was used for analyzing crystal packing of the P3HT.
structure, however the PCE of the post-annealed P3HT/PCBM OBHJ did Absorbance was measured with an UV/Vis/NIR spectrometer
not outperform than that of the DBHJ; only the P3HT/PCBM with the (PerkinElmer, Lambda 750).

Fig. 2. (a) Schematics of the device fabrication process. The P3HT was used as a donor and the C60 was used as an acceptor. (b) SEM images of the imprinted P3HT donor.
(c) The morphology of the IDOBHJ.
B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684 679

3. Results and discussion nanometers in OBHJ should have sufficient interfacial area, but it
would be difficult to implement [61] and was thus ruled out in the
3.1. Theoretical analysis analysis. In the case of DBHJ (Fig. 3(b)), the exciton dissociation
efficiency increases with the interfacial area and reaches approxi-
Fig. 3 shows exciton dissociation efficiencies, charge transport mately 100% when the interfacial area per unit volume is approxi-
efficiencies and internal quantum efficiencies (IQE) of various mately 0.4 nm  1 which corresponds to the domain size of 6 nm.
morphologies plotted versus the interfacial area per unit volume. This is because an exciton travels a shorter distance before it is
The IQE is described as bellow. dissociated into an electron and a hole. However, the charge
transport efficiency decreases with increases in the interfacial area
IQ E ¼ ηextD  ηchrT  ηchrC ð9Þ
as the charge transport pathways to electrodes become longer and
here, exciton dissociation efficiency is defined as the ratio of the more non-linear and, in some cases, become disconnected. Thus,
total number of dissociated excitons after charge transfer to the the possibility of charge recombination increases when charges
number of generated excitons and is the multiplication of exciton are traveling long time in the active layer. Long lasting, free
diffusion and charge transfer efficiency. In this analysis the charge charges in a disconnected site especially interact with other
collection efficiency is the ratio of the total number of extracted charges by Coulombic force and attract the opposite charge to
charges to the number of dissociated charges after charge transfer recombine. The bimolecular recombination [62] affecting by
process. As shown in the figures, the exciton dissociation efficien- domain size [63] or trap depth [64] is crucial in determining the
cies increase as the interfacial area increases and the charge IQE so it is considered in our model. Thus, the IQE has an optimum
transport efficiencies decrease. Interfaces are definitely beneficial value when the interfacial area per unit volume is approximately
to exciton dissociation, while they diminish charge transport 0.28 nm  1. Therefore, IDOBHJ was proposed to take advantage of
unless properly organized. Thus, optimum values exist in IQE. In the charge transport efficiency of OBHJ and exciton dissociation
the OBHJ (Fig. 3(a)), the charge transport efficiencies are nearly efficiency of DBHJ.
independent of the interfacial area and close to 100%. As long as Fig. 3(c) shows efficiencies of the IDOBHJ. The normalized inter-
excitons are dissociated to the free charge carriers, most of the diffused width is denoted as W, the inter-diffused width divided
charge carriers contribute to the current flow. However, as shown by pillar size, which was 27 nm in this study. Therefore, IDOBHJ
in Fig. 3(a) and (b), the interfacial areas in OBHJ are quite small with W ¼0 corresponds to the morphology of OBHJ. Additionally,
compared to that in the DBHJ. The highest interfacial area in the IDOBHJ with W¼ 1 corresponds to the morphology of the DBHJ.
OBHJ was based on a pillar size of 18 nm. A pillar size of a few IDOBHJ with W¼ 1 was chosen from the DBHJ with the highest IQE
shown in Fig. 3(b) and was denoted as an optimized DBHJ for
reference purpose. The charge transport efficiency of IDOBHJ
shows weak dependency on the interfacial area, indicating the
formation of straight pathways at the center of the pillars to
deliver the free charges to the electrodes while maintaining a large
interfacial area. Alternatively, the exciton dissociation rates
increase rapidly as W increases, while not disturbing the charge
transport. Based on Fig. 3, the IDOBHJ performs better than OBHJ
and DBHJ because two different morphologies, one from the OBHJ
and the other from the DBHJ, exist in the IDOBHJ. Thus, the
interfacial area per unit volume of IDOBHJ lies between those of
OBHJ and DBHJ, yet this amount is sufficient for IDOBHJ to have
such high exciton dissociation efficiency. Additionally, the charge
transport efficiency can also be optimized; the ordered morphol-
ogy of IDOBHJ is definitely beneficial to the charge transport. Even
a disordered morphology would exhibit greater charge transport
efficiency than DBHJ as the long, non-linear paths in the IDOBHJ
are limited to the inter-diffused width, so charges should travel a
shorter distance than those in DBHJ. Lyons et al. [25] studied the
effect of interface morphology on the performance of the OPVs—
performance comparison between OPVs with sharp interfaces and
with diffuse interfaces. They concluded that the OPV with sharp
interfaces has the higher Jsc than that with diffuse interfaces
because of the higher charge collection efficiency, although it
has the lower exciton diffusion efficiency, i.e., lower interfacial
area. We should emphasize that the diffuse interfaces by Lyons
et al. [25] are different to the interfaces of IDOBHJ. In fact,
interfaces of IDOBHJ should resemble to the sharp interfaces in
their paper. As we mentioned previously, we generated morphol-
ogies of IDOBHJ by inserting morphologies of DBHJ and interfaces
of DBHJ should be sharp since the Ising–Hamiltonian drive
materials to gather the same materials.
Fig. 3. Theoretical analysis of the exciton dissociation efficiency, charge transport One of the advantages of the Monte Carlo method is that the
efficiency and internal quantum efficiency (IQE) according to the interfacial area carrier trajectories can be followed [53]. We compared exciton and
per unit volume of (a) OBHJ, (b) DBHJ and (c) IDOBHJ. Here, the normalized inter- charge traveling distances of OBHJ, DBHJ and IDOBHJ, as shown in
diffused width is denoted as W, the inter-diffused width divided by pillar size,
which is 27 nm. Therefore, IDOBHJ with W¼ 0 corresponds to a morphology of
Fig. 4. The exciton traveling distance is defined as the average
OBHJ. IDOBHJ with W¼ 1 corresponds to a morphology of DBHJ. The IDOBHJ with traveling distance from exciton generation to exciton dissociation.
W ¼1 was chosen from the DBHJ with the highest IQE in Fig. 3(b). The normalized charge traveling distance, l, is defined as the average
680 B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684

Fig. 4. Simulation results of the exciton traveling distances and normalized charge
traveling distances according to the normalized inter-diffused width of IDOBHJ. The
exciton traveling distance (left) is the average traveling distance from exciton
generation to exciton dissociation. The charge traveling distance (right) is the
average distance from charge generation due to exciton dissociation to charge
extraction at the electrodes. The charge traveling distance is the average distance
between holes and electrons and is normalized to an active layer thickness, i.e.
90 nm.

distance from charge generation due to exciton dissociation to charge


extraction at the electrodes.
!
∑ N hop  a0 =N carriers
N carriers
l¼ ð10Þ
L
here, Nhop, Ncarriers and L is a number of hoppings, a number of
carriers (here, 1600 excitons and 1260 charges) and active layer
thickness, i.e., 90 nm. Fig. 4 shows traveling distances of excitons and
charges according to the normalized inter-diffused width, W, which
are proportional to the interfacial area per unit volume. The exciton
traveling distance decreases as W increases due to the increase in
interfacial area. The exciton traveling distance was reduced by more
than a factor of two when IDOBHJ was W¼ 0 and IDOBHJ was Fig. 5. Theoretical analysis of (a) J–V curve characteristics (current density–
external voltage) according to the normalized inter-diffused width, W, in IDOBHJ
W¼0.89 or 1.0. The normalized charge traveling distance increases as
and (b) power conversion efficiency (PCE) according to the normalized inter-
W increases due to the increase in interfacial area, confirming the diffused width, W. The PCEs of IDOBHJ are greater than DBHJ. Inset table present
charge transport paths in IDOBHJ with low W are relatively straight. short circuit current density (Jsc), open circuit voltage (V), fill factor (FF) and PCE
The exciton traveling distances of IDOBHJ with W¼0.89 to W¼1 are according to W.
nearly the same, yet the charge traveling distances are different, and
charges in IDOBHJ with W¼ 0.89 travel shorter distances. This clearly
indicates that IDOBHJ could outperform DBHJ, i.e. IDOBHJ with W¼1, when W increases because the exciton dissociation increases due to
because of comparable exciton dissociation efficiency and enhanced the increase in the interfacial area. The open circuit voltage is indepen-
charge transport efficiency. dent of morphology as it depends on the energy band diagrams of
Fig. 5(a) shows J–V curve characteristics (current density–external donor and acceptor materials [66]. The FF also exhibits dependency to
voltage) according to the normalized inter-diffused widths, W, in the the inter-diffused width, W, of the IDOBHJ. When W is small, i.e.,
IDOBHJ. The power conversion efficiency (PCE) is expressed as. pathways to electrodes are relatively straight and continuous, the fill
factor is high. However, when W increases, the FF decreases because
J sc  V oc  FF
PCE½% ¼  100 ð11Þ disordered and disconnected morphologies appear that act as resis-
P in
tance. When charges move directly to the electrode in response to an
where Jsc denotes the short circuit current density, Voc is the open external voltage, the J–V curve shows diode characteristics [67].
circuit voltage, FF is the fill factor and Pin as an incident power. However, by increasing the W, i.e., increasing the disordered morphol-
To achieve the high PCE, high Jsc, Voc and FF are required; however, ogy, the charge migration is related to the external voltage and
they have an interdependent relation according to the active layer influenced by the neighboring morphology. For example, although
morphology. The large interfacial area enhances the Jsc but lowers FF. an electron is forced to move to a cathode, in some cases it cannot
An optimum interfacial area per unit volume of a conventional DBHJ is move when an electron donor that cannot accommodate the electron
shown in Fig. 3(b) and in Fig. 5. Morphology control in a DBHJ is not is near the cathode. As a result, the electron must detour the obstacle.
trivial [65], so the morphology of the optimized DBHJ should be The possibility of charge recombination increases because of this long
difficult to achieve. However, it would be relatively easy to achieve a journey toward the electrode.
uniform morphology from device to device in the IDOBHJ as the Fig. 5(b) shows the power conversion efficiency (PCE) of
disordered part in the IDOBHJ morphology is smaller than that of a IDOBHJ with various normalized inter-diffused widths, W; W¼0
DBHJ. As shown in Fig. 5(a), the short circuit current density increases to W ¼1. The PCEs of IDOBHJ clearly outperform both the OBHJ,
B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684 681

i.e., the IDOBHJ with W¼ 0, and DBHJ, i.e., IDOBHJ with W¼ 1, in


which IDOBHJ surpasses OBHJ by 18–38%. The PCE of IDOBHJ
(W ¼0.89) is greater than the optimized DBHJ by 9%. The Jsc of
IDOBHJ with W¼ 0.44 is lower than the optimized DBHJ because of
the smaller interfacial area. However, the FF of IDOBHJ with
W¼0.44 is larger than optimized DBHJ by 12%. In the IDOBHJ
with W¼0.89, the short circuit current density and fill factor are
1.5% and 7% greater than the optimized DBHJ, respectively.

3.2. Experimental results and discussion

We performed the secondary ion mass spectrometer (SIMS)


analysis, as shown in Fig. 6, (i) to get the evidence of the inter-
diffusion process and (ii) to quantify the inter-diffusional rate. The
SIMS data in Fig. 6 indeed verified the existence of inter-diffusion
process. By counting the number of D (deuterium) in C60D36, [68]
the distribution of C60 in the active layer was estimated [48,60]. The
deuterium labeling method for SIMS has been used to measure
Fig. 7. The XRD data of P3HTs. The P3HT planar or pillar indicates that P3HT was
distribution of fullerene based acceptor, because they do not have a imprinted with a PDMS mold with or without pillar structures respectively. The
distinctive element compare to P3HT [69]. The 100 nm thickness of P3HT pillar/C60 was post annealed at 150 C for 30 min. The P3HT pillars show the
P3HT was spin coated on the glass/ITO/PEDOT:PSS substrate and face-on structure indicating the existence of π–π stacking.
cured by thermal annealing at 150 1C for 60 min. Then, C60D36 was
evaporated onto the cured P3HT followed by post annealing (PA) at nanoimprinted. Also, XRD of P3HT/C60 structures after the post
150 1C for various annealing time. In Fig. 6, we deduce the followings; annealing process indicate the face-on peak. This suggests that the
C60 diffused slowly into the P3HT donor layer and did not fully face-on structure formed prior to the C60 deposition was sustained
penetrate into the P3HT under given PA time. The diffusion constants against to the C60 penetration during the post annealing process.
of C60D36 into the P3HT were estimated to be around 10  15 cm2 s  1 The absorbance of glass/ITO/PEDOT:PSS/P3HT/C60 according to
for P3HT. [60,70,71] It is well known that acceptors based on C60 various post annealing (PA) time were measured and presented in
such as the phenyl-C61-butyric acid methyl ester (PCBM) cannot Fig. 8. The peaks at 550 nm and 603 nm indicate that there exist
penetrate into the crystalline region of the P3HT so they only diffuse π–π staking in the imprinted P3HT [73]. The external quantum
into amorphous region of the P3HT [48]. Therefore, it can be deduced efficiencies (EQEs) of various PA time suggest there is an optimum
that crystalline phases exist in the thermally cured P3HT. In terms of PA time for forming IDOBHJ—the sample with PA 30 min shows
device perspective, while high hole mobility in the crystalline P3HT the highest performance among others. While there is no appreci-
is useful, large size of the crystalline P3HT is not good for power able difference among absorbance data, EQEs depend on strongly
conversion efficiency—C60 cannot easily diffuse into the P3HT. So, on PA time. This indicates light absorption did not affect the EQEs.
the size of the crystalline P3HT needs to be optimized [71]. Morphological changes during the inter-diffusion process caused
As we discussed previously, the crystalline P3HT plays an impor- differences in the EQE data.
tant role on charge transport [72]. Especially π–π staking orientation The power conversion efficiencies (PCEs) of IDOBHJ with normal-
(0 1 0), i.e., the face-on structure, exhibits higher mobility than alkyl ized inter-diffused width are presented in Fig. 9. Since there exist quite
chain orientation (1 0 0), i.e., the edge-on structure [41]. The crystal difference in PCE between P3HT:PCBM and P3HT:C60, the y axis was
packings of pristine and imprinted P3HTs were compared by the drawn in double scale for the ease of comparison. In this way, the
X-ray diffraction (XRD) measurement shown in Fig. 7. For reference existence of an optimum morphology between OBHJ and DBHJ can be
purpose, we also presented XRD analysis of an annealed P3HT on seen clearly. The normalized inter-diffused width (W) in experiment
which we imprinted with PDMS of no pillar patterns. As shown in was calculated from SIMS data in Fig. 6 and normalized to the initial
Fig. 7, the face-on structures appear only when the P3HT was pillar diameter, 100 nm. Note that the simulation was based on the
P3HT/PCBM in order to ensure the accuracy of the results for ease of
comparing with other literature data [23,53]. However, we used P3HT/
C60 for experimental demonstration since it is hard to achieve the
morphology of OBHJ out of P3HT/PCBM due to the solvent permeation
[48] during the spin coating process. In this way, the un-annealed
specimen could be regarded as OBHJ and others would be either
IDOBHJ or DBHJ depending on PA time. The samples have an optimum
PA time where they show the highest PCE, which is consistent with
the theoretical analysis. The theoretical analysis predicts that there
exists optimum morphology of IDOBHJ. Based on the SIMS data in
Fig. 6, PA time controls degrees of the inter-diffused widths. As
describe above, the morphology during the post-annealing in reality
should evolve accompanying with the polymer-chains- crystallization
[41] and the very rapid fullerene derivatives diffusion through the
amorphous polymer region [47,48]. In this sense, the pillars of P3HTs
can be fragmented into grains due to the interpenetrated PCBMs. The
fill factor (FF) data suggest this hypothesis. The FF in theoretical
analysis decreases as the inter-diffused width increases because
Fig. 6. The secondary ion mass spectrometer (SIMS) data. The y axis represents for increased disordered morphology hinders the charge transport.
numbers of deuterium (D) in C60D36 and PA stands for the post annealing process. In experimental data, however, the FF has an optimum point at
682 B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684

charge collection. IDOBHJ utilizes only advantages of the high


exciton diffusion efficiency of the disordered bulk heterojunction
(DBHJ) and the high charge collection efficiency of the ordered
bulk heterojunction (OBHJ). The IDOBHJ has a large interfacial area
while not destroying direct pathways to the electrode too much.
By using dynamical Monte Carlo (DMC) method combined with
the first reaction method (FRM), the power conversion efficiency
of the IDOBHJ was demonstrated to outperform the optimized
DBHJ by approximately 9% due to a greater short circuit current
density and fill factor. For IDOBHJ, the short circuit current density
is greater by 1.5% than the optimized DBHJ because it possesses a
comparable interfacial area per unit volume yet higher charge
transport efficiency. Additionally, the fill factor was 7% greater
than that of the optimized DBHJ due to more direct and vertical
pathways to electrodes in IDOBHJ. The traveling distance analysis
of excitons and charges confirm that the large interfacial area of
Fig. 8. Experimental data of the absorbances and external quantum efficiencies IDOBHJ provides sufficient exciton dissociation and ensures effi-
(EQEs) of IDOBHJ according to various post annealing (PA) time. While there are no cient charge collection.
appreciable differences among absorbance data, EQEs depend strongly on PA time. We realized IDOBHJ structure by using the nanoimprinting
This indicates clearly light absorption did not affect the EQEs but morphological technique combined with the thermal annealing process; the
changes due to the inter-diffusion process cause these changes.
nanoimprint implements pillar structures on the P3HT. Then, C60
was evaporated onto the P3HT followed by the thermal annealing to
accelerate inter-diffusion between donors and acceptors. The sec-
ondary ion mass spectrometer was used to characterize this inter-
diffusion process. XRD analysis on the materials reveals existence of
the face-on structure in the IDOBHJ structure, which is beneficial for
efficient charge transport. The power conversion efficiencies and
external quantum efficiencies of IDOBHJ show that there exists an
optimum post annealing time. Since increase in post annealing time
leads to increase in inter-diffused width, theoretical analysis is then
clearly confirmed by these experimental results.

Acknowledgements

This work was supported by the National Research Foundation


of Korea Grant (NRF-2011-0028729, NRF-2010-0021487) by
Fig. 9. The power conversion efficiencies (PCEs) of P3HT/C60 experimental data and the Korean Government Ministry of Education, Science and Tech-
P3HT/PCBM simulation results according to the normalized inter-diffused width. nology (MEST).

References
Table 2
Experimentally measured Voc, Jsc, FF and PCE of P3HT/C60 IDOBHJ of various post
[1] S. Van Bavel, S. Veenstra, J. Loos, On the importance of morphology control in
annealing time.
polymer solar cells, Macromol. Rapid Commun. 31 (2010) 1835–1845.
[2] J. Nelson, Polymer: fullerene bulk heterojunction solar cells, Mater. Today 14
Post annealing time Normalized inter- Voc Jsc [mA/ FF PCE (2011) 462–470.
[min] diffused width [V] cm2] [%] [3] G. Li, R. Zhu, Y. Yang, Polymer solar cells, Mater. Today 6 (2012) 153–161.
[4] R. Søndergaard, M. Hösel, D. Angmo, T.T. Larsen-Olsen, F.C. Krebs, Roll-to-roll
0 0.32 0.55 3.98 0.33 0.74 fabrication of polymer solar cells, Mater. Today 15 (2012) 36–49.
20 0.56 0.55 3.95 0.41 0.91 [5] C. Deibel, V. Dyakonov, Polymer-fullerene bulk heterojunction solar cells, Rep.
30 0.6 0.54 6.80 0.48 1.77 Prog. Phys. 73 (2010) 096401-1–096401-39.
40 0.62 0.55 6.75 0.45 1.56 [6] K. Li, P.P. Khlyabich, L. Li, B. Burkhart, B.C. Thompson, J.C. Campbell, Influence
60 0.7 0.59 4.36 0.32 0.83 of exciton diffusion and charge-transfer state dissociation efficiency on the
short-circuit current densities in semi-random donor/acceptor polymer: full-
erene solar cells, J. Phys. Chem. C 117 (2013) 6940–6948.
[7] V.D. Mihailetchi, L.J.A. Koster, J.C. Hummelen, P.W.M. Blom, Photocurrent
30 min for the sample as shown in Table 2. Initial increase in the FF generation in polymer-fullerene bulk heterojunctions, Phys. Rev. Lett. 93
could be attributed to the formation of good interfacial contact (2004). (216601-216601-216601-216604).
between Al and C60 [74]. Later decrease in the FF is due to increased [8] F. Yang, S.R. Forrest, Photocurrent generation in nanostructured organic solar
cells, ACS Nano 2 (2008) 1022–1032.
sizes of inter-diffused region, which hinders charge transport pathway [9] X. Yang, J. Loos, Toward high-performance polymer solar cells: the importance
to the electrodes. of morphology control, Macromolecules 40 (2007) 1353–1362.
[10] J.A. Carr, Y. Chen, M. Elshobaki, R.C. Mahadevapuram, S. Chaudhary, Control-
ling nanomorphology in plastic solar cells, Nanomater. Energy 1 (2011) 18–26.
[11] J.J.M. Halls, K. Pichler, R.H. Friend, S.C. Moratti, A.B. Holmes, Exciton diffusion
4. Conclusion and dissociation in a poly(p-phenylenevinylene)/C60 heterojunction photo-
voltaic cell, Appl. Phys. Lett. 68 (1996) 3120–3122.
In summary, an inter-diffused ordered bulk heterojunction [12] A. Haugeneder, M. Neges, C. Kallinger, W. Spirkl, U. Lemmer, J. Feldmann,
U. Scherf, E. Harth, A. Gügel, K. Müllen, Exciton diffusion and dissociation in
(IDOBHJ) structure is suggested to satisfy both interdependent conjugated polymer/fullerene blends and heterostructures, Phys. Rev. B:
parameters in organic photovoltaics, i.e., exciton diffusion and Condens. Matter 59 (1999) 15346–15351.
B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684 683

[13] T.J. Savenije, J.E. Kroeze, X. Yang, J. Loos, The effect of thermal treatment on the (p-arylene-vinylene): PCBM bulk-heterojunction solar cells, Polymer 52 (2011)
morphology and charge carrier dynamics in a polythiophene–fullerene bulk 3819–3826.
heterojunction, Adv. Funct. Mater. 15 (2005) 1260–1266. [40] K.M. Coakley, M.D. McGehee, Conjugated polymer photovoltaic cells, Chem.
[14] G. Li, Y. Yao, H. Yang, V. Shrotriya, G. Yang, Y. Yang, Solvent Annealing effect in Mater. 16 (2004) 4533–4542.
polymer solar cells based on poly(3-hexylthiophene) and methanofullerenes, [41] M. Aryal, K. Trivedi, W. Hu, Nano-confinement induced chain alignment in
Adv. Funct. Mater. 17 (2007) 1636–1644. ordered P3HT nanostructures defined by nanoimprint lithography, ACS Nano 3
[15] M. Campoy-Quiles, T. Ferenczi, T. Agostinelli, P.G. Etchegoin, Y. Kim, (2009) 3085–3090.
T.D. Anthopoulos, P.N. Stavrinou, D.D.C. Bradley, J. Nelson, Morphology evolu- [42] M. Zhou, M. Aryal, K. Mielczarek, A. Zakhidov, W. Hu, Hole mobility enhance-
tion via self-organization and lateral and vertical diffusion in polymer:full- ment by chain alignment in nanoimprinted poly(3-hexylthiophene) nanograt-
erene solar cell blends, Nat. Mater. 7 (2008) 158–164. ings for organic electronics, J. Vac. Sci. Technol., B 28 (2010) C6M63–C66M67.
[16] D.H. Wang, J.S. Moon, J. Seifter, J. Jo, J.H. Park, O.O. Park, A.J. Heeger, Sequential [43] R.A. Marsh, C. Groves, N.C. Greenham, A microscopic model for the behavior of
processing: control of nanomorphology in bulk heterojunction solar cells, nanostructured organic photovoltaic devices, J. Appl. Phys. 101 (2007)
Nano Lett. 11 (2011) 3163–3168. 083509-1–083509-7.
[17] T. Wang, A.J. Pearson, D.G. Lidzey, R.A.L. Jones, Evolution of structure, [44] J. Hou, C. Yang, C. He, Y. Li, Poly[3-(5-octyl-thienylene-vinyl)-thiophene]: A
optoelectronic properties, and device performance of polythiophene: fullerene side-chain conjugated polymer with very broad absorption band, Chem.
solar cells during thermal annealing, Adv. Funct. Mater. 21 (2011) 1383–1390. Commun. (2006) 871–873.
[18] M. Graetzel, R.A.J. Janssen, D.B. Mitzi, E.H. Sargent, Materials interface [45] J.U. Lee, J.W. Jung, T. Emrick, T.P. Russell, W.H. Jo, Synthesis of C60-end capped
engineering for solution-processed photovoltaics, Nature 488 (2012) 304–312. P3HT and its application for high performance of P3HT/PCBM bulk hetero-
[19] I.C. Henderson, N. Clarke, On modelling surface directed spinodal decomposi- junction solar cells, J. Mater. Chem. 20 (2010) 3287–3294.
tion, Macromol. Theor Simul 14 (2005) 435–443. [46] M.H. Yun, J. Kim, C. Yang, J.Y. Kim, A simultaneous achievement of high
[20] P.K. Watkins, A.B. Walker, G.L.B. Verschoor, Dynamical Monte Carlo modelling performance and extended thermal stability of bulk-heterojunction polymer
of organic solar cells: the dependence of internal quantum efficiency on solar cells using a polythiophene-fullerene block copolymer, Sol. Energy
morphology, Nano Lett. 5 (2005) 1814–1818. Mater. Sol. Cells 104 (2012) 7–12.
[21] B. Lei, Y. Yao, A. Kumar, Y. Yang, V. Ozolins, Quantifying the relation between [47] B.A. Collins, E. Gann, L. Guignard, X. He, C.R. McNeill, H. Ade, Molecular
the morphology and performance of polymer solar cells using Monte Carlo miscibility of polymer fullerene blends, J. Phys. Chem. Lett. 1 (2010)
simulations, J. Appl. Phys. 104 (2008) 024504-1–024504-6. 3160–3166.
[22] C. Groves, R.G.E. Kimber, A.B. Walker, Simulation of loss mechanisms in [48] D. Chen, F. Liu, C. Wang, A. Nakahara, T.P. Russell, Bulk heterojunction
organic solar cells: a description of the mesoscopic Monte Carlo technique photovoltaic active layers via bilayer interdiffusion, Nano Letters 11 (2011)
and an evaluation of the first reaction method, J. Chem. Phys. 133 (2010) 2071–2078.
144110-1–144110-7. [49] K.R. Thomas, A. Chenneviere, G. Reiter, U. Steiner, Nonequilibrium behavior of
[23] L.Y. Meng, Y. Shang, Q.K. Li, Y.F. Li, X.W. Zhan, Z.G. Shuai, R.G.E. Kimber, A. thin polymer films, Phys. Rev. E: Stat. Nonlinear Soft Matter Phys. 83 (2011)
B. Walker, Dynamic Monte Carlo simulation for highly efficient polymer blend 021804-1–021804-8.
photovoltaics, J. Phys. Chem. B 114 (2010) 36–41. [50] K. Yee, Numerical solution of initial boundary value problems involving
[24] B.P. Lyons, N. Clarke, C. Groves, The quantitative effect of surface wetting layers Maxwell's equations in isotropic media, IEEE Trans. Antennas Propag. 14
on the performance of organic bulk heterojunction photovoltaic devices, J. (1966) 302–307.
Phys. Chem. C 115 (2011) 22572–22577. [51] A. Taflove, Computational electrodynamics: the finite-difference time-domain
[25] B.P. Lyons, N. Clarke, C. Groves, The relative importance of domain size, method, Artech House, Boston, 1995.
domain purity and domain interfaces to the performance of bulk- [52] J. Nelson, Diffusion-limited recombination in polymer-fullerene blends and its
heterojunction organic photovoltaics, Energy Environ. Sci. 5 (2012) influence on photocurrent collection, Phys. Rev. B: Condens. Matter 67 (2003)
7657–7663. 155209-1–155209-10.
[26] K. Maturova, S.S. van Bavel, M.M. Wienk, R.A.J. Janssen, M. Kemerink, Mor- [53] L. Meng, D. Wang, Q. Li, Y. Yi, J.-L. Bredas, Z. Shuai, An improved dynamic
phological device model for organic bulk heterojunction solar cells, Nano Lett. Monte Carlo model coupled with Poisson equation to simulate the perfor-
9 (2009) 3032–3037. mance of organic photovoltaic devices, J. Chem. Phys. 134 (2011) 124102-
[27] H. van Eersel, R.A.J. Janssen, M. Kemerink, Mechanism for efficient photo- 1–124102-7.
induced charge separation at disordered organic heterointerfaces, Adv. Funct. [54] M.C. Heiber, A. Dhinojwala, Dynamic Monte Carlo modeling of exciton
Mater. 22 (2012) 2700–2708. dissociation in organic donor-acceptor solar cells, J. Chem. Phys. 137 (2012)
[28] K. Maturová, S.S. van Bavel, M.M. Wienk, R.A.J. Janssen, M. Kemerink, 014903-1–014903-11.
Description of the morphology dependent charge transport and performance [55] D. Kipp, V. Ganesan, A kinetic Monte Carlo model with improved charge
of polymer: fullerene bulk heterojunction solar cells, Adv. Funct. Mater. 21 injection model for the photocurrent characteristics of organic solar cells, J.
(2011) 261–269. Appl. Phys. 113 (2013) 234502-1–234502-15.
[29] W. Chen, M.P. Nikiforov, S.B. Darling, Morphology characterization in organic [56] Y. Yang, M. Aryal, K. Mielczarek, W. Hu, A. Zakhidov, Nanoimprinted P3HT/C
and hybrid solar cells, Energy Environ. Sci. 5 (2012) 8045–8074. [sub 60] solar cells optimized by oblique deposition of C[sub 60], J. Vac. Sci.
[30] C.R. McNeill, Morphology of all-polymer solar cells, Energy Environ. Sci. 5 Technol., B 28 (2010) C6M104–C106M107.
(2012) 5653–5667. [57] X. He, F. Gao, G. Tu, D.G. Hasko, S. Hüttner, N.C. Greenham, U. Steiner, R.
[31] D.H. Wang, J.K. Kim, O.O. Park, J.H. Park, Analysis of surface morphological H. Friend, W.T.S. Huck, Formation of well-ordered heterojunctions in polymer:
changes in organic photovoltaic devices: bilayer versus bulk-heterojunction, PCBM photovoltaic devices, Adv. Funct. Mater. 21 (2011) 139–146.
Energy Environ. Sci. 4 (2011) 1434–1439. [58] X. He, F. Gao, G. Tu, D. Hasko, S. Hü ttner, U. Steiner, N.C. Greenham, R.
[32] H.-L. Yip, A.K.Y. Jen, Recent advances in solution-processed interfacial materi- H. Friend, W.T.S. Huck, Formation of nanopatterned polymer blends in
als for efficient and stable polymer solar cells, Energy Environ. Sci. 5 (2012) photovoltaic devices, Nano Lett. 10 (2010) 1302–1307.
5994–6011. [59] C. Yeretzian, J.B. Wiley, K. Holczer, T. Su, S. Nguyen, R.B. Kaner, R.L. Whetten,
[33] F. Padinger, R.S. Rittberger, N.S. Sariciftci, Effects of postproduction treatment Partial separation of fullerenes by gradient sublimation, J. Phys. Chem. 97
on plastic solar cells, Adv. Funct. Mater. 13 (2003) 85–88. (1993) 10097–10101.
[34] R. Hamilton, C.G. Shuttle, B. O'Regan, T.C. Hammant, J. Nelson, J.R. Durrant, [60] N.D. Treat, M.A. Brady, G. Smith, M.F. Toney, E.J. Kramer, C.J. Hawker, M.
Recombination in annealed and nonannealed polythiophene/fullerene solar L. Chabinyc, Interdiffusion of PCBM and P3HT reveals miscibility in a photo-
cells: transient photovoltage studies versus numerical modeling, J. Phys. Chem. voltaically active blend, Adv. Energy Mater. 1 (2011) 82–89.
Lett. 1 (2010) 1432–1436. [61] H.J. Park, M.-G. Kang, L.J. Guo, Large Area High, Density sub-20 nm SiO2
[35] M.-S. Su, C.-Y. Kuo, M.-C. Yuan, U.S. Jeng, C.-J. Su, K.-H. Wei, Improving device nanostructures fabricated by block copolymer template for nanoimprint
efficiency of polymer/fullerene bulk heterojunction solar cells through lithography, ACS Nano 3 (2009) 2601–2608.
enhanced crystallinity and reduced grain boundaries induced by solvent [62] J.D. Servaites, M.A. Ratner, T.J. Marks, Organic solar cells: a new look at
additives, Adv. Mater. 23 (2011) 3315–3319. traditional models, Energy Environ. Sci. 4 (2011) 4410–4422.
[36] J.H. Tsai, Y.C. Lai, T. Higashihara, C.J. Lin, M. Ueda, W.C. Chen, Enhancement of [63] C. Groves, N.C. Greenham, Bimolecular recombination in polymer electronic
P3HT/PCBM photovoltaic efficiency using the surfactant of triblock copolymer devices, Phys. Rev. B: Condens. Matter 78 (2008) 155205-1–155205-8.
containing poly(3-hexylthiophene) and poly(4-vinyltriphenylamine) seg- [64] R.C.I. MacKenzie, C.G. Shuttle, M.L. Chabinyc, J. Nelson, Extracting microscopic
ments, Macromolecules 43 (2010) 6085–6091. device parameters from transient photocurrent measurements of P3HT:PCBM
[37] D. Deribew, E. Pavlopoulou, G. Fleury, C. Nicolet, C. Renaud, S.-J. Mougnier, solar cells, Adv. Energy Mater. 2 (2012) 662–669.
L. Vignau, E. Cloutet, C. Brochon, F. Cousin, G. Portale, M. Geoghegan, [65] R. Gaudiana, Organic photovoltaics: challenges and opportunities, J. Polym.
G. Hadziioannou, Crystallization-driven enhancement in photovoltaic perfor- Sci., Part B: Polym. Phys. 50 (2012) 1014–1017.
mance through block copolymer incorporation into P3HT: PCBM blends, [66] B.P. Rand, D.P. Burk, S.R. Forrest, Offset energies at organic semiconductor
Macromolecules 46 (2013) 3015–3024. heterojunctions and their influence on the open-circuit voltage of thin-film
[38] F.C. Jamieson, E.B. Domingo, T. McCarthy-Ward, M. Heeney, N. Stingelin, J. solar cells, Phys. Rev. B: Condens. Matter 75 (2007) 115327-1–115327-11.
R. Durrant, Fullerene crystallisation as a key driver of charge separation in [67] J.H. Lee, S. Cho, A. Roy, H.-T. Jung, A.J. Heeger, Enhanced diode characteristics
polymer/fullerene bulk heterojunction solar cells, Chem. Sci. 3 (2012) of organic solar cells using titanium suboxide electron transport layer, Appl.
485–492. Phys. Lett. 96 (2010) 163303-1–163303-3.
[39] S. Rathgeber, J. Perlich, F. Kühnlenz, S. Türk, D.A.M. Egbe, H. Hoppe, R. Gehrke, [68] F. Cataldo, S. Iglesias-Groth, A. Manchado, Synthesis and FT-IR spectroscopy of
Correlation between polymer architecture, mesoscale structure and photovol- perdeuterofullerane: C60D36 evidences of isotope effect in the stability of
taic performance in side-chain-modified poly(p-arylene-ethynylene)-alt-poly C60D36, Fullerenes Nanotubes Carbon Nanostruct. 17 (2009) 378–389.
684 B. Jung et al. / Solar Energy Materials & Solar Cells 120 (2014) 675–684

[69] J.K.J. van Duren, X. Yang, J. Loos, C.W.T. Bulle-Lieuwma, A.B. Sieval, J. [74] W.L. Ma, C.Y. Yang, X. Gong, K. Lee, A.J. Heeger, Thermally stable, efficient
C. Hummelen, R.A.J. Janssen, Relating the morphology of poly(p-phenylene polymer solar cells with nanoscale control of the interpenetrating network
vinylene)/methanofullerene blends to solar-cell performance, Adv. Funct. morphology, Adv. Funct. Mater. 15 (2005) 1617–1622.
Mater. 14 (2004) 425–434. [75] R.C.I. MacKenzie, T. Kirchartz, G.F.A. Dibb, J. Nelson, Modeling nongeminate
[70] B. Watts, W.J. Belcher, L. Thomsen, H. Ade, P.C. Dastoor, A quantitative study of recombination in P3HT:PCBM solar cells, J. Phys. Chem. C 115 (2011)
PCBM diffusion during annealing of P3HT: PCBM blend films, Macromolecules 9806–9813.
42 (2009) 8392–8397. [76] E. von Hauff, J. Parisi, V. Dyakonov, Field effect measurements on charge
[71] H.W. Ro, B. Akgun, B.T. O'Connor, M. Hammond, R.J. Kline, C.R. Snyder, S. carrier mobilities in various polymer-fullerene blend compositions, Thin Solid
K. Satija, A.L. Ayzner, M.F. Toney, C.L. Soles, D.M. DeLongchamp, Poly(3- Films 511 (2006) 506–511.
hexylthiophene) and [6,6]-Phenyl-C61-butyric acid methyl ester mixing in [77] T.J. Savenije, J.E. Kroeze, X.N. Yang, J. Loos, The formation of crystalline P3HT
organic solar cells, Macromolecules 45 (2012) 6587–6599.
fibrils upon annealing of a PCBM: P3HT bulk heterojunction, Thin Solid Films
[72] A. Marrocchi, D. Lanari, A. Facchetti, L. Vaccaro, Poly(3-hexylthiophene):
511 (2006) 2–6.
synthetic methodologies and properties in bulk heterojunction solar cells,
Energy Environ. Sci. 5 (2012) 8457–8474.
[73] D. Kekuda, H.-S. Lin, M. Chyi, Wu, J.-S. Huang, K.-C. Ho, C.-W. Chu, The effect of
solvent induced crystallinity of polymer layer on poly(3-hexylthiophene)/C70
bilayer solar cells, Sol. Energy Mater. Sol. Cells 95 (2011) 419–422.

Potrebbero piacerti anche