ArticlePDF Available

Carrier motion in as-spun and annealed P3HT:PCBM blends revealed by ultrafast optical electric field probing and Monte Carlo simulations

Authors:

Abstract and Figures

Charge transport dynamics in solar cell devices based on as-spun and annealed P3HT:PCBM films are compared using ultrafast time-resolved optical probing of the electric field by means of field-induced second harmonic generation. The results show that charge carriers drift about twice as far during the first 3 ns after photogeneration in a device where the active layer has been thermally annealed. The carrier dynamics were modelled using Monte-Carlo simulations and good agreement between experimental and simulated drift dynamics was obtained using identical model parameters for both cells, but with different average PCBM and polymer domain sizes. The calculations suggest that small domain sizes in as-spun samples limit the carrier separation distance disabling their escape from geminate recombination.
Content may be subject to copyright.
2686 |Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 This journal is ©the Owner Societies 2014
Cite this: Phys. Chem. Chem. Phys.,
2014, 16,2686
Carrier motion in as-spun and annealed
P3HT:PCBM blends revealed by ultrafast optical
electric field probing and Monte Carlo simulations
Vytautas Abramavic
ˇius,*
ab
Dimali Amarasinghe Vithanage,
c
Andrius Deviz
ˇis,
a
Yingyot Infahsaeng,
c
Annalisa Bruno,
d
Samuel Foster,
d
Panagiotis E. Keivanidis,
e
Darius Abramavic
ˇius,
b
Jenny Nelson,
d
Arkady Yartsev,
c
Villy Sundstro
¨m
c
and
Vidmantas Gulbinas
a
Charge transport dynamics in solar cell devices based on as-spun and annealed P3HT:PCBM films are
compared using ultrafast time-resolved optical probing of the electric field by means of field-induced
second harmonic generation. The results show that charge carriers drift about twice as far during the first
3 ns after photogeneration in a device where the active layer has been thermally annealed. The carrier
dynamics were modelled using Monte-Carlo simulations and good agreement between experimental and
simulated drift dynamics was obtained using identical model parameters for both cells, but with different
average PCBM and polymer domain sizes. The calculations suggest that small domain sizes in as-spun
samples limit the carrier separation distance disabling their escape from geminate recombination.
1. Introduction
Diminishing sources of fossil fuels and the need to meet rising
global demands for carbon-free energy have led to renewable
sources being explored as replacements. Conjugated polymers
have been investigated as alternatives to solar cells based on
inorganic semiconductors
1
due to their light weight, flexibility,
abundance of material, low material usage and manufacturing
costs. The invention of the bulk heterojunction structure (BHJ)
using a donor and acceptor homogeneously mixed to produce the
active material
2
has aided the increase in solar cell efficiency,
which is presently 9.2% for the best reported cells.
3
To improve
device efficiency, the charge dynamics have also been investigated
and three key stages in the charge separation pathway have been
identified – charge generation,
4
transport
5
and recombination.
6
Excitons are generated when light within the absorption
spectrum of the material impinges on the devices. These excitons
very rapidly
6
separate into positive and negative charges forming
Coulombically bound electron–hole pairs (or charge transfer
states (CT)). In order to separate further, the charges have to
overcome the Columbic attraction and form mobile charges
which can move towards the electrodes through a combination
of diffusion and drift.
7
The collection of the separated charges
results in completion of the circuit and current produced by
the solar cell.
Here we study the polymer:fullerene combination poly(3-
hexylthiophene) (P3HT) and [6,6]-phenyl-C61butyric acid
methyl ester (PCBM). The method of processing P3HT:PCBM
devicesisknowntoimpacttheactive layer morphology and, as a
result, the efficiency of devices, and has therefore been extensively
studied. Several factors have been investigated with the aim
of improving device efficiency, such as the effect of solvent,
morphology, film thickness and processing conditions.
8–17
Annealing was shown to have a great impact on the conversion
efficiency of P3HT:PCBM solar cells, quite different from most
other polymer:fullerene blends. The carrier dynamics of
annealed and as-spun P3HT:PCBM films have been studied
using several techniques aiming at investigating differences in
mobility,
5,15
morphology,
14–17
EQE
14
and I–V characteristics.
5
Annealing to a high temperature changes the morphology
and enhances the hole mobility,
5,9
resulting in it being only an order
of magnitude below the electron mobility.
5
A similar effect was
achieved with slow solvent evaporation.
18
Using microsecond time
scale techniques, a large spread in mobilities and their differences
in as-spun and annealed samples have been reported.
5,15,19
The
measurements show that the two different processing methods
drastically affect the mobility and charge separation time scales.
Morphological studies have shown that high temperature
a
Center for Physical Sciences and Technology, Savanoriu 231, LT-02300 Vilnius,
Lithuania
b
Department of Theoretical Physics, Vilnius University, Sauletekio 9-III,
LT-10222 Vilnius, Lithuania. E-mail: Vytautas.Abramavicius@ff.vu.lt
c
Chemical Physics, Lund University, Box 124, 221 00 Lund, Sweden
d
Imperial College London, South Kensington Campus, London SW7 2AZ, UK
e
Center for Nano Science and Technology@PoliMi, Istituto Italiano di Tecnologia,
Via Giovanni Pascoli, 70/3, 20133 Milano, Italy
Received 31st October 2013,
Accepted 5th December 2013
DOI: 10.1039/c3cp54605e
www.rsc.org/pccp
PCCP
PAPER
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
View Journal
| View Issue
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 | 2687
results in phase separation due to crystallization of the
polymer
5,15,16,20
and formation of large PCBM clusters.
12,14–17
There is a consensus that thermal annealing results in improved
device efficiency due to enhanced phase segregation, which
consequently leads to increased charge separation efficiency,
21,22
improved hole conductivity and formation of optimized charge
transport pathways
9,15
and consequently reduced bimolecular
recombination.
23
The mechanism through which the thermal annealing
process enables higher charge carrier mobilities is now fairly
well understood. Annealing induced crystallisation of the polymer
results in larger domains (thicker lamellae) of the pure polymer
and at the same time expels fullerene molecules out of the
crystallising polymer, thereby making more fullerene available
to build a robust electron transport network.
20,24–26
It is clear from
such studies that the improvement in charge collection (reflected
through photocurrent quantum efficiency) is associated with the
growth in pure polymer and fullerene domains and resulting
improvement in charge carrier mobility relative to the recombina-
tion coefficient.
5,9
In this paper, we aim at unveiling how morphology affects
charge transport by investigating charge mobility and charge
separation at earlier timescales using electric field-induced
second harmonic generation (TREFISH)
7,27,28
and MC simula-
tions. We find the morphology to influence the mobility and
carrier separation on the ps to ns time scale. MC simulations
show that the different carrier drift kinetics in as-spun and
annealed blends may be explained by more extensive material
segregation, leading to larger P3HT and PCBM domains in
annealed material, enabling fast separation of carriers at larger
distances and preventing their geminate recombination.
2. Experiment
The experimental setup and theory have been previously
described,
7,27,28
so only a brief account is given here. TREF ISH
is a pump–probe technique, employing a femtosecond laser pulse
to excite the sample devices and generate charges, and a probe
pulse that generates the SHG signal probing the dynamics of
the charges. An applied electric field breaks the symmetry of the
material, allowing to generate the second harmonic signal of
the probe pulse. The intensity and time dependence of the
second harmonic signal monitors the electric field dynamics in
the sample. The excitation pulse (400 nm, 36 nJ per pulse) was
obtained by frequency doubling the fundamental of the Ti:Sa
laser at 800 nm; a photon density of B10
12
photons per cm
2
per
pulse for the sample was used, which is below the onset of strong
second order (non-geminate) recombination. The probing wave-
length was obtained using an optical parametric amplifier
(TOPAS) at 1200 nm, the second harmonic of which was within
the sensitivity of the photomultiplier detector. The sample device
was made using a PEDOT:PSS/ITO anode and an aluminium (Al)
cathode. The PEDOT:PSS was spun to form a 40–60 nm film and
the total device had an overall thickness of B115 nm. The sample
cells were all prepared in a clean room environment.
3. Monte-Carlo simulation model
The simulation model has been described in ref. 7. Briefly, charge
carriermotionintheP3HT:PCBMblendwasmodelledbyassuming
a cubic lattice, characterized by a lattice constant ain all three
dimensions. The lattice is divided into the donor part, where only
the hole is allowed to reside and the acceptor part for the electron.
The acceptor sites are defined by filling the lattice volume with
ellipsoids of acceptor material (see Fig. 2) with typical average
volume, which is later on used as a fitting parameter. The ellipsoids
have arbitrary proportions and they are placed in arbitrary positions
in the lattice and they overlap each other, thus mimicking the
distribution of PCBM in the actual blend. Next, the remaining space
in the lattice is filled with donor sites, which are used to create
arbitrarily oriented and folded chains representing the polymer. The
length of a chain is chosen randomly from the interval [L3, L+3],
where Lis the average length of chains. It should be noted, that
such a blend model apparently cannot reproduce the real blend
morphology, particularly of the annealed blend where a lamellar
structure is suggested to be formed. The results of the calculation
should rather be seen as a qualitative representation of morphology
to rationalize the observed carrier dynamics.
The electron and hole dynamics are controlled by site energy
properties. In the presence of an external electric field the energy of
an electron (hole) in the lattice consists of three parts: (1) the internal
site self-energy E
r
, which is assumed to be a random Gaussian value;
(2) the energy due to the constant external electric field F, and
(3) the energy due to the Coulomb interaction between charges
of opposite sign E
C
. The electron (hole) energy thus equals to:
E
f
(r)=E
r
8(Fr)+E
C
. (1)
The site self-energy is distributed according to a modified
Gaussian distribution, which is defined as a weighted sum of a
normal Gaussian distribution with addition of longer exponential
tails. The energy of the external electric field was accounted
for by projecting the site position to the electric field direction.
The electrostatic interaction energy is given by the shifted
Coulomb potential
EC¼ q
4pee0
1
reh þba (2)
Here qis the electron charge, r
eh
is the distance between the
electron and the hole, eis the mean permittivity of the material, a
is the lattice constant and bis a positive dimensionless parameter,
which accounts for deviation of the Coulomb potential from the
point charge approximation at short distances and sets the
appropriate initial electron–hole interaction energy.
Both types of charges perform hopping in their respective
domains of the lattice. The hopping is simulated using the
Monte-Carlo algorithm as follows. As the initial configuration
the hole and electron are placed on neighbouring sites in the
interfacial region of the donor and acceptor domains. Only the
nearest neighbour sites are taken into account for the hopping
event. A charge can hop into one of six surrounding sites when
it is far from the interface while hopping possibilities are fewer
in the interfacial region. The hopping rates n
mn
for both the
Paper PCCP
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
2688 |Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 This journal is ©the Owner Societies 201 4
electron and the hole are calculated using the Miller–Abrahams
formula:
29
nmn ¼n0exp 2grmn
ðÞ
exp EnEm
kT

;En4Em
1;EnEm
8
<
:
;(3)
where gis a parameter which characterizes the inverse localiza-
tion length of a charge density, r
mn
is the distance between the
origin site mand the target site n,E
m
and E
n
are their energies
respectively. In the acceptor domain the hopping rate n
0
n
A
is
constant, while in the donor part we assume the value n
0
n
D1
for hopping to a target site located in a straight part of the same
polymer chain as the origin site, n
0
n
D2
for hopping to a target
site located on a folding point (the point where the orientation
of the polymer chain changes) of the same polymer as the
origin site and n
0
n
D3
for hopping to a target site located on a
different polymer chain.
It is assumed that a hole is less likely to hop to a site located
on another polymer chain, thus the corresponding hopping
rate prefactor n
D3
is smaller than both n
D1
and n
D2
. We also
assume that a hole avoids folding points where holes move
slower than in straight sections of the polymer chain, thus
n
D2
on
D1
. It should be noted that a simple isotropic medium
model was unable to reproduce the carrier drift kinetics during
initial tens of ps therefore this more complex model, previously
suggested to simulate carrier motion in the pure polymer,
28
was used.
When all rates of possible hopping events (including holes
and electrons) have been evaluated, the rates are being translated
into hopping probabilities according to:
pmn ¼nmn
P
k
nk
;(4)
where the summation is performed over all calculated rates of
both the hole and the electron. These probabilities are then
used to determine the destination site nfor either the hole or
the electron, chosen by a linearly distributed random number.
The charge configuration is then switched to the one that has
been determined and the rates of the next hopping events
are recalculated.
For the simulation a 100 400 400 lattice was used. This
lattice simulates the actual structure of the blend, motivating
that no cyclic boundary conditions are introduced. Initially,
charges were created at a random location at the interface
between the donor and the acceptor regions and due to the
external electric field they drifted apart in opposite directions.
While charges moved through the lattice, the distance between
them projected in the direction of the external electric field F,
d
k
(t) was recorded and the result was averaged over 5000
realizations.
Only one electron–hole pair was present in the lattice at a
time, thus the model did not account for the nongeminate
charge carrier recombination. The geminate recombination
was also not accounted for assuming it to be much slower than
the examined time domain.
4. Experimental results
Fig. 1 shows the carrier drift dynamics in as-spun and annealed
samples for various applied voltages, calculated by the proce-
dure described in ref. 19 from the experimentally measured
TREFISH kinetics (not shown). Briefly, the electric field kinetics
was reconstructed from the EFISH kinetics by using steady state
EFISH dependence on the electric field strength. Next we
assume that the electric field drop is proportional to the carrier
drift distance and obtain the drift distance kinetics by normal-
izing the time-resolved field drop to the total field drop at long
delay time when all carriers are extracted and, thus, their
average drift distance equals to the half of the film thickness.
The drift distances presented in Fig. 1 are averaged over
electrons and holes and rapidly increase on the tens of ps time
scale in both samples. At long times (>200 ps) the increase
rate gradually slows down to reach a separation distance of
15–30 nm (depending on film treatment) at 2.5 ns. The drift
distances are approximately proportional to the internal
electric field, suggesting that the initial carrier mobility is
independent of the electric field strength. Qualitatively, similar
drift dynamics has been observed for neat polymers
28
and
Fig. 1 (a) Experimental (symbols) and simulated (lines) charge drift
dynamics in the as-spun (a) and annealed (b) samples at various electric
fields strengths.
PCCP Paper
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 | 2689
attributed to carrier relaxation within a distributed density of
states. The drift distances in the as-spun sample are about half
of those in the annealed sample at the same applied voltages.
The electron and the hole drift in opposite directions by
about 2.5 nm during the initial 10 ps at 6.7 10
5
Vcm
1
electric field in the annealed sample. Thus, the electron–hole
separation distance along the electric field is about 5 nm. This
separation distance is around half as large in the as spun
sample as in the annealed sample and is approximately propor-
tional to the applied field.
5. MC calculation results
Monte Carlo simulations by the procedure described above
have been performed to model the carrier drift dynamics and
to gain insight into the microscopic properties responsible for
the observed differences in drift dynamics of annealed and
as-spun material. The modelling of the hole motion dynamics
accounts for the hole relaxation within the density of states
(DOS), different hole hopping rates within a conjugated segment
(n
D1
), between segments (n
D2
) and between polymer chains (n
D3
).
The electron motion is simpler – the model accounts for the
electron relaxation within the DOS and electron motion inside
PCBM domains is characterised by a single electron hopping rate
prefactor, n
A
. Both electron and hole motions are also affected
by the domain structure of the blend; reaching the domain
boundaries carriers are forced to search for alternative pathways
to continue their motions – this process results in a domain-size
dependence of carrier mobility. The drift kinetics at different
voltages were simulated with the same model parameters, only
varying the internal field strength.
Carrier drift kinetics in as-spun and annealed samples have
been modelled by using exactly the same motion parameters
except for polymer and PCBM domain sizes. The best agree-
ment was obtained with an average acceptor domain diameter
of 7.5 nm for the as-spun sample and 33 nm for the annealed
sample. As a result of fullerene aggregation the polymer domain
dimensions were accordingly larger for annealed samples as
well, but because of nonregular shapes their quantitative char-
acterization, is more difficult. Fig. 2 illustrates the corresponding
material morphologies and Fig. 1 shows the simulated carrier
drift dynamics. The quite good agreement with experimental
results obtained for all curves with only one free variable, the
domain size, validates the simulation results. The obtained
domain dimensions of the annealed samples are somewhat
larger than the B10 nm domains estimated in similar samples
from experimental results.
30
On the other hand, quite similar
domain sizes of 10 to 30 nm were estimated by MC modelling of
carrier recombination in a P3HT:PCBM blend.
23
The MC simula-
tions do not perfectly reproduce the carrier drift kinetics in annealed
samples at high applied electric fields (6.7 10
5
Vcm
1
)attimes
longer than 1 ns. This is not very surprising taking into account the
relatively simple blend structure used in calculations.
We proceed to infer effective charge carrier mobilities from
the data for separation as a function of time. Note that these are
not mobilities as usually defined, describing drift of relaxed
populations of charges in the steady state, but instantaneous
mobilities describing the instantaneous separation velocity of
unrelaxed charge carrier populations. Since the experimental
data gives us information on the sum of electron and hole drift
distances, the actual electron and hole mobilities remain
undisclosed, the ratio between electron and hole hopping rates
being a free parameter. We have chosen the electron hopping
rate on the basis of additional available information on the
ultrafast time-resolved electron mobility and on the basis of the
best agreement between experimental and calculated carrier
drift kinetics. By means of time-resolved microwave conductivity,
Savenije et al.
31
obtained the electron mobility inside PCBM
nanocrystals of 8 10
2
cm
2
V
1
s
1
and a similar mobility of
about 0.1 cm
2
V
1
s
1
was also obtained on a subpicosecond-
several ps time-scale in PCBM film by dynamic Stark effect
measurements.
32
Thus, we have chosen an electron hopping
rate prefactor n
A
to give an electron mobility of 0.1 cm
2
V
1
s
1
at 0.3 ps, while its subsequent evolution was obtained from the
best fitting with experimental data. Similar information on the
initial hole mobility in P3HT is not available and therefore it
was obtained from the modelling of the carrier drift kinetics.
The best agreement was obtained with about ten times lower
hole mobility than that of electrons. The simulation parameters
used to obtain the best agreement between calculated and
measured drift kinetics (see Fig. 1) are presented in Table 1.
A lower initial hole mobility in comparison with the electron
mobility was also concluded for a polyfluorene/fullerene blend.
33
On the other hand, mobilities obtained from time resolved THz
measurements on another polyfluorene low-bandgap polymer/
fullerene blend (APFO3/PCBM) show that picosecond time scale
hole mobility is higher than the electron mobility by approxi-
mately a factor of five.
34
The reason for this difference in the
relative mobility of holes and electrons is probably a result of
different sensitivity to intra- and inter-chain hole transport of
the experimental methods. Fitting of the simulation and
experimental results allows significant freedom of correlated
variation of hopping rates of electrons and holes in different
directions, thus the distinction of electron and hole mobilities
is not reliable. Therefore we present, in Fig. 3, the carrier
mobility averaged over electrons and holes, obtained directly
from the experimental data. The short time carrier mobility is
almost two times larger for the annealed sample. Carrier mobilities
Fig. 2 Cross section of typical simulated structures of as-spun (left) and
annealed (right) samples. Dark areas denote acceptor regions (PCBM) and
white areas denote donor regions (P3HT). The red line represents the
length of 50 nm.
Paper PCCP
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
2690 |Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 This journal is ©the Owner Societies 2 01 4
in both samples drop down several tens of times during 1 ns.
The TREFISH mobilities at t> 1 ns approach literature data for
steady state mobility,
5,18,19
indicating that carrier populations have
almost relaxed into trap states during this time. Qualitatively
similar mobility dynamics was observed in pure polymer films,
27,28
showing that both inherent polymer and PCBM properties,
as well as nanostructured blend morphology, are responsible
for the mobility dynamics.
Our experimental data give information on the carrier drift
distance, while the absolute carrier separation distance is deter-
mined by carrier diffusion as well as drift. These two processes are
interrelated through the Einstein relation D=mk
B
T/q,whereDis the
diffusion coefficient, mis the carrier mobility, k
B
is the Boltzmann
coefficient, Tis the temperature and qis the electron charge. In our
previouspaperwehaveshownthatthediusiondistanceonaps
time scale significantly exceeds the drift distance at low fields and is
responsible for the weakly field dependent carrier separation yield.
7
MC simulation is a convenient approach to obtain average
absolute carrier separation distances caused by both carrier
drift and diffusion from the carrier drift kinetics. Fig. 4 shows a
comparison of the absolute carrier separation distances in
as-spun and annealed samples at different electric field strengths.
At zero electric field, only the diffusion drives the carrier
motion, thus curves at zero field represent diffusion driven
charge separation dynamics. At 0 and 1.7 10
5
Vcm
1
electric
fields the separation distances on a tens of ps time scale are
almost independent of the sample annealing. The difference
appears on a ns time scale, when electrons approach the
boundaries of small PCBM domains in the as-spun sample,
while in the annealed sample with larger PCBM domains, they
continue an unrestricted motion. At higher electric field, when
the carrier drift contributes more to their motion, charge
Fig. 3 Carrier mobility averaged over electrons and holes for as spun
(closed circles) and annealed (open circles) samples at 4.7 10
5
Vcm
1
field strength.
Table 1 Numerical values of the parameters of the model
Lattice dimension
in the xdirection (nm)
Lattice dimension
in the ydirection (nm)
Lattice dimension
in the zdirection (nm)
Lattice constant
a(nm)
Average size of the
acceptor ellipsoid M(nm)
100 400 400 1 As-spun: 220 annealed: 19 800
Average length of the
donor chain (nm)
Hopping rate prefactor
in the acceptor n
A
(s
1
)
Hopping rate prefactor
in the donor n
D1
(s
1
)
Hopping rate prefactor
in the donor n
D2
(s
1
)
Hopping rate prefactor
in the donor n
D3
(s
1
)
6 2.8 10
16
210
15
110
15
510
14
Parameter g(nm
1
) Disorder in the
acceptor s
A
(meV)
Disorder in the
donor s
D
(meV)
Temperature T(K) Mean dielectric permittivity e
5 70 80 293 3
Correction parameter bof the initial electron–hole interaction energy Fraction of exponential distribution exp(E/s)
in the modified Gaussian distribution
2 0.19
Fig. 4 Calculated absolute charge carrier separation distances in as spun
(dotted lines) and annealed (solid lines) samples at different electric field
strengths obtained by Monte Carlo simulation using a model fitted to the
drift distance data in Fig. 1. The curves at higher electric field strengths are
vertically shifted.
PCCP Paper
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
This journal is ©the Owner Societies 2014 Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 | 2691
carriers move faster and reach domain boundaries in the as
spun sample already on a ps time scale, thus the difference in
separation distances appears already during tens of ps. Strongly
restricted carrier motion in the as-spun sample with smaller
PCBM and polymer domains prevents carrier escape from the
Coulomb attraction. In devices such restricted carrier motion
leads to enhanced charge carrier recombination, which is
apparently one of the major factors limiting the carrier genera-
tion yield and performance efficiency of non-annealed P3HT/
PCBM solar cells.
23
Our MC simulations have been performed assuming that
only nearest neighbor e–h pairs are created by exciton splitting
at the donor–acceptor interface as was suggested in ref. 7
and 33. However, there are publications
35–37
arguing that charge
carrier separation at much longer distances takes place on a
femtosecond time scale and it helps for final separation of e–h
pairs into free charges. Since this is still an open question,
which could be also related to the blend annealing, we have
also performed additional calculations directed towards evalua-
tion of the role of the initial carrier separation distance in the
charge separation process. Fig. 5 shows the calculated absolute
charge carrier separation distances at zero applied field with
the model parameters obtained from the above described
simulations. Diffusion driven separation at long times is large
with larger initial separation, but the influence of the initial
separation gradually decreases with time and after several ns
the separation distance is almost independent of the initial
ultrafast separation if this separation is significantly smaller
than 8 nm. Thus, initial carrier separation only weakly influences
the final carrier separation process (at several ns when charges
have reached a distance where the electrostatic attraction energy
is similar to kT), unless the initial separation is comparable with
the Coulomb capture radius. On the other hand, as we have
discussed in ref. 7, the large distance carrier separation is hardly
compatible with our experimental carrier drift data showing no
quasi-instantaneous carrier drift component.
6. Conclusions
In conclusion, our experimental investigations of the initial
carrier motion in as-spun and annealed P3HT:PCBM blends
together with Monte Carlo simulations of the carrier drift
dynamics suggest a mechanism for the improved performance
of annealed solar cells. The initial carrier drift rates, on a
subnanosecond–nanosecond time scale are about two times larger
in annealed samples. Monte Carlo simulations of the motion
dynamics suggest that the increase in the carrier separation rate
caused by blend annealing is related to the increased polymer and
PCBM domain sizes enabling longer distance carrier separation on
a ps time scale, which reduces the probability of their geminate
recombination and thus increases the free charge carrier genera-
tion yield in annealed samples. On the other hand, the role of
other material properties such as the presence of energy traps, or
formation of semicrystalline polymer domains, which change as a
result of annealing, cannot be completely ruled out.
Additional MC simulations directed towards evaluation of
theroleoftheinitialcarrierseparation distance showed that the
more efficient carrier separation in annealed samples can be
hardly related to increased initial carrier separation distance.
The initial separation distance only weakly influences the carrier
separation efficiency at times and distances where free charges
are formed if it is shorter than about 8 nm, while longer distance
separation is non-compatible with our experimental data.
Acknowledgements
This research was funded by the European Social Fund under
the Global Grant measure, by the Swedish and European
Research Councils (ERC 226136-VISCHEM), by the Swedish
Energy Agency and the Knut & Alice Wallenberg Foundation
and by Laser Lab Europe (project ID LLC001578, framework of the
Initiative of Infrastructures Programme), by the UK Engineering
and Physical Sciences Research Council via the Supergen
programme and by the Royal Society.
Notes and references
1 J. Nelson, Curr. Opin. Solid State Mater. Sci., 2002, 6, 87–95.
2 G. Yu, J. Gao, J. C. Hummelen, F. Wudl and A. J. Heeger,
Science, 1995, 270, 1789–1791.
3 Z. He, C. Zhong, S. Su, M. Xu, H. Wu and Y. Cao, Nat.
Photonics, 2012, 6, 591–595.
4 J. Guo, H. Ohkita, H. Benten and S. Ito, J. Am. Chem. Soc.,
2010, 132, 6154–6164.
5 V. D. Mihailetchi, H. Xie, B. de Boer, L. Jan Anton Koster and
P. W. M. Blom, Adv. Funct. Mater., 2006, 16, 699–708.
6 S. De, T. Pascher, M. Maiti, K. G. Jespersen, T. Kesti,
F. Zhang, O. Ingana
¨s, A. Yartsev and V. Sundstro
¨m, J. Am.
Chem. Soc., 2007, 129, 8466–8472.
7 D. Amarasinghe Vithanage, A. Devizˇis, V. Abramavic
ˇius,
Y. Infahsaeng, D. Abramavic
ˇius, R. C. I. MacKenzie,
P. E. Keivanidis, A. Yartsev, D. Hertel, J. Nelson,
V. Sundstro
¨m and V. Gulbinas, Nat. Commun., 2013, 4, 2334.
Fig. 5 Calculated time dependence of the absolute carrier separation
distance at zero electric field and at various initial separation distances.
Paper PCCP
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
2692 |Phys. Chem. Chem. Phys., 2014, 16, 2686--2692 This journal is ©the Owner Societies 2014
8 C. Brabec, V. Dyakonov and U. Scherf, IEEE J. Sel. Top.
Quantum Electron., 2013, 16, 1517.
9 R. A. Marsh, J. M. Hodgkiss, S. Albert-Seifried and
R. H. Friend, Nano Lett., 2010, 10, 923–930.
10 F. Zhang, K. G. Jespersen, C. Bjo
¨rstro
¨m, M. Svensson,
M. R. Andersson, V. Sundstro
¨m, K. Magnusson, E. Moons,
A. Yartsev and O. Ingana
¨s, Adv. Funct. Mater., 2006, 16, 667–674.
11 M. Reyes-Reyes, K. Kim and D. L. Carroll, Appl. Phys. Lett.,
2005, 87, 083506.
12A.M.Ballantyne,T.A.M.Ferenczi,M.Campoy-Quiles,
T.M.Clarke,A.Maurano,K.H.Wong,W.Zhang,N.Stingelin-
Stutzmann,J.-S.Kim,D.D.C.Bradley,J.R.Durrant,
I. McCulloch, M. Heeney, J. Nelson, S. Tierney, W. Duffy,
C. Mueller and P. Smith, Macromolecules, 2010, 43, 1169–1174.
13 Y. Kim, S. A. Choulis, J. Nelson, D. D. C. Bradley, S. Cook
and J. R. Durrant, Appl. Phys. Lett., 2005, 86, 063502.
14 C. H. Woo, B. C. Thompson, B. J. Kim, M. F. Toney and
J. M. J. Frechet, J. Am. Chem. Soc., 2008, 130, 16324–16329.
15 T. Agostinelli, S. Lilliu, J. G. Labram, M. Campoy-Quiles,
M. Hampton, E. Pires, J. Rawle, O. Bikondoa,
D. D. C. Bradley, T. D. Anthopoulos, J. Nelson and
J. E. Macdonald, Adv. Funct. Mater., 2011, 21, 1701–1708.
16 E. Verploegen, R. Mondal, C. J. Bettinger, S. Sok, M. F. Toney
and Z. Bao, Adv. Funct. Mater., 2010, 20, 3519–3529.
17 X. Yang, J. Loos, S. C. Veenstra, W. J. H. Verhees,
M. M. Wienk, J. M. Kroon, M. A. J. Michels and
R. A. J. Janssen, Nano Lett., 2005, 5, 579–583.
18 J. Huang, G. Li and Y. Yang, Appl. Phys. Lett., 2005, 87, 112105.
19 C. Nam, D. Su and C. T. Black, Adv. Funct. Mater., 2009, 19,
3552–3559.
20 M.Campoy-Quiles,T.Ferenczi,T.Agostinelli,P.G.Etchegoin,
Y. Kim, T. D. Anthopoulos, P. N. Stavrinou, D. D. C. Bradley
and J. Nelson, Nat. Mater., 2008, 7, 158–164.
21 D.Veldman,O.Ipek,S.C.J.Meskers,J.Sweelssen,M.M.Koetse,
S.C.Veenstra,J.M.Kroon,S.S.vanBavel,J.Loosand
R. A. J. Janssen, J. Am. Chem. Soc., 2008, 130, 7221–7235.
22 P. E. Keivanidis, T. M. Clarke, S. Lilliu, T. Agostinelli,
J. E. Macdonald, J. R. Durrant, D. D. C. Bradley and
J. Nelson, J. Phys. Chem. Lett., 2010, 1, 734–738.
23 R. Hamilton, C. G. Shuttle, B. O’Regan, T. C. Hammant,
J. Nelson and J. R. Durrant, J. Phys. Chem. Lett., 2010, 1,
1432–1436.
24 T. G. J. van der Hofstad, D. Di Nuzzo, M. van den Berg,
R. A. J. Jensses and S. C. J. Meskens, Adv. Energy Mater.,
2012, 2, 1095–1099.
25 C. Mu
¨ller, T. A. M. Ferenczi, M. Campoy-Quiles, J. M. Frost,
D. D. C. Bradley, P. Smith, N. Stingelin-Stutzmann and
J. Nelson, Adv. Mater., 2008, 18, 3510–3515.
26 T. Agostinelli, S. Lilliu, J. G. Labram, M. Campoy-Quiles,
M. Hampton, E. Pires, J. Rawle, O. Bikondoa,
D. D. C. Bradley, T. D. Anthopoulos, J. Nelson and
J. E. Macdonald, Adv. Funct. Mater., 2011, 21, 1701–1708.
27 A. Devizis, A. Serbenta, K. Meerholz, D. Hertel and
V. Gulbinas, Phys. Rev. Lett., 2009, 103, 027404.
28 A. Devizis, K. Meerholz, D. Hertel and V. Gulbinas, Chem.
Phys. Lett., 2010, 498, 302–306.
29 A. Miller and E. Abrahams, Phys. Rev., 1960, 120, 745–755.
30 W. L. Ma, C. Y. Yang and A. J. Heeger, Adv. Mater., 2007, 19,
1387–1390.
31 T. J. Savenije, J. E. Kroeze, M. M. Wienk, J. M. Kroon and
J. M. Warman, Phys. Rev. B: Condens. Matter Mater. Phys.,
2004, 69, 155205.
32 J. Cabanillas-Gonzalez, T. Virgili, A. Gambetta, G. Lanzani,
T. Anthopoulos and D. De Leeuw, Phys. Rev. Lett., 2006,
96, 106601.
33 D. Veldman, O. Ipek, S. C. J. Meskers, J. Sweelssen,
M. M. Koetse, S. C. van Bavel, J. Loos and R. A. J. Janssen,
J. Am. Chem. Soc., 2008, 130, 7721–7735.
34 C. S. Ponseca, H. Nemec, N. Vukmirovic, S. Fusco, E. Wang,
M. R. Andersson, P. Chabera, A. Yartsev and V. Sundstrom,
J. Phys. Chem. Lett., 2012, 3, 2442–2446.
35 A. A. Bakulin, A. Rao, V. G. Pavelyev, P. H. M. van. Loos-
drecht, M. S. Pshenichnikov, D. Niedzialek, J. Cornil,
D. Beljonne and R. H. Friend, Science, 2012, 335, 1340–1344.
36 I. A. Howard, R. Mauer, M. Meister and F. Laquai, J. Am.
Chem. Soc., 2010, 132, 14866–14876.
37 C. Deibel, T. Storbel and V. Dyakonov, Phys. Rev. Lett., 2009,
103, 036402.
PCCP Paper
Published on 05 December 2013. Downloaded by Nanyang Technological University on 08/10/2014 03:18:58.
View Article Online
... Barroso et al. carried out preliminary studies of the effects of external electrical biases [34] and electrocatalysts [48] on photocurrent generation and lifetimes of photo-generated holes in Fe2O3-water systems. Abramavičius et al. made pivotal early progress into understanding charge-carrier hopping in external electric fields in P3HT:PCBM films for photovoltaics (PV) by optical probing [57], which served as an early "proof-of-concept". Most promisingly, Pegg and Hatton modeled geometric enhancement of intrinsic electric fields in organic PV materials using a gross, finite element approach, which omits, unfortunately, atomistic and electronic details [58]. ...
... Barroso et al. carried out preliminary studies of the effects of external electrical biases [34] and electrocatalysts [48] on photocurrent generation and lifetimes of photogenerated holes in Fe 2 O 3 -water systems. Abramavičius et al. made pivotal early progress into understanding charge-carrier hopping in external electric fields in P3HT:PCBM films for photovoltaics (PV) by optical probing [57], which served as an early "proof-of-concept". Most promisingly, Pegg and Hatton modeled geometric enhancement of intrinsic electric fields in organic PV materials using a gross, finite element approach, which omits, unfortunately, atomistic and electronic details [58]. ...
Article
Full-text available
The grand challenges in renewable energy lie in our ability to comprehend efficient energy conversion systems, together with dealing with the problem of intermittency via scalable energy storage systems. Relatively little progress has been made on this at grid scale and two overriding challenges still need to be addressed: (i) limiting damage to the environment and (ii) the question of environmentally friendly energy conversion. The present review focuses on a novel route for producing hydrogen, the ultimate clean fuel, from the Sun, and renewable energy source. Hydrogen can be produced by light-driven photoelectrochemical (PEC) water splitting, but it is very inefficient; rather, we focus here on how electric fields can be applied to metal oxide/water systems in tailoring the interplay with their intrinsic electric fields, and in how this can alter and boost PEC activity, drawing both on experiment and non-equilibrium molecular simulation.
... Typical examples of time dependent properties are spectral diffusion of excitons [33][34][35] or time-dependent charge carrier mobilities [36][37][38][39]. In accordance with this picture, recent studies on prototypical polymer:fullerene blend devices revealed significant thermalization effects [39][40][41][42][43][44]. One example is PCDTBT blended with PCBM. ...
... The situation is very different for a well-crystallized thermally annealed P3HT:PCBM sample (see Figure 5.6e, data from [181]), where incremental recombination data for a wide range of fluences and delays lay on the same line, implying a time-independent recombination mechanism at all times studied here. This is in accordance with insignificant spectral relaxation in TAS experiments on P3HT:PCBM blends [100] and the nearly absence of time dependent mobility in the nanosecond range obtained with TREFISH (time resolved electric field induced second harmonic generation) experiments [42]. Therefore, the pronounced slow-down of recombination in the PCDTBT:PCBM sample is attributed to the amorphous nature of this blend. ...
Thesis
Organic semiconductors are a promising class of materials. Their special properties are the particularly good absorption, low weight and easy processing into thin films. Therefore, intense research has been devoted to the realization of thin film organic solar cells (OPVs). Because of the low dielectric constant of organic semiconductors, primary excitations (excitons) are strongly bound and a type II heterojunction needs to be introduced to split these excitations into free charges. Therefore, most organic solar cells consist of at least an electron donor and electron acceptor material. For such donor acceptor systems mainly three states are relevant; the photoexcited exciton on the donor or acceptor material, the charge transfer state at the donor-acceptor interface and the charge separated state of a free electron and hole. The interplay between these states significantly determines the efficiency of organic solar cells. Due to the high absorption and the low charge carrier mobilities, the active layers are usually thin but also, exciton dissociation and free charge formation proceeds rapidely, which makes the study of carrier dynamics highly challenging. Therefore, the focus of this work was first to install new experimental setups for the investigation of the charge carrier dynamics in complete devices with superior sensitivity and time resolution and, second, to apply these methods to prototypical photovoltaic materials to address specific questions in the field of organic and hybrid photovoltaics. Regarding the first goal, a new setup combining transient absorption spectroscopy (TAS) and time delayed collection field (TDCF) was designed and installed in Potsdam. An important part of this work concerned the improvement of the electronic components with respect to time resolution and sensitivity. To this end, a highly sensitive amplifier for driving and detecting the device response in TDCF was developed. This system was then applied to selected organic and hybrid model systems with a particular focus on the understanding of the loss mechanisms that limit the fill factor and short circuit current of organic solar cells. The first model system was a hybrid photovoltaic material comprising inorganic quantum dots decorated with organic ligands. Measurements with TDCF revealed fast free carrier recombination, in part assisted by traps, while bias-assisted charge extraction measurements showed high mobility. The measured parameters then served as input for a successful description of the device performance with an analytical model. With a further improvement of the instrumentation, a second topic was the detailed analysis of non-geminate recombination in a disordered polymer:fullerene blend where an important question was the effect of disorder on the carrier dynamics. The measurements revealed that early time highly mobile charges undergo fast non-geminate recombination at the contacts, causing an apparent field dependence of free charge generation in TDCF experiments if not conducted properly. On the other hand, recombination the later time scale was determined by dispersive recombination in the bulk of the active layer, showing the characteristics of carrier dynamics in an exponential density of state distribution. Importantly, the comparison with steady state recombination data suggested a very weak impact of non-thermalized carriers on the recombination properties of the solar cells under application relevant illumination conditions. Finally, temperature and field dependent studies of free charge generation were performed on three donor-acceptor combinations, with two donor polymers of the same material family blended with two different fullerene acceptor molecules. These particular material combinations were chosen to analyze the influence of the energetic and morphology of the blend on the efficiency of charge generation. To this end, activation energies for photocurrent generation were accurately determined for a wide range of excitation energies. The results prove that the formation of free charge is via thermalized charge transfer states and does not involve hot exciton splitting. Surprisingly, activation energies were of the order of thermal energy at room temperature. This led to the important conclusion that organic solar cells perform well not because of predominate high energy pathways but because the thermalized CT states are weakly bound. In addition, a model is introduced to interconnect the dissociation efficiency of the charge transfer state with its recombination observable with photoluminescence, which rules out a previously proposed two-pool model for free charge formation and recombination. Finally, based on the results, proposals for the further development of organic solar cells are formulated.
... Nonetheless, upon application of an electric field, the hole and electron drift in opposite directions as a result of their opposing charges as seen clearly in Figure 1 (wherein the applied static electric field are respectively seen to be pointing along the +z axis for both supercells), where evidently the electric field breaks symmetrically 33 in order to pave the way for the enhancement of electron−hole transport. 8,33,34 In fact, the enhanced drift combined with increased diffusion due to the oscillating nuclei in the external field substantially attenuates the obvious problem of electron−hole recombination and, essentially, helps to enhance the contact of the holes with water at the surface, thus facilitating water breakup. Moreover, external electric field water molecule dipoles are inclined to align partially with the applied field, in which for alternating (e/m) fields the water dipoles rotate back and forth with the applied field as it changes direction, 35 thus leading to substantial intramolecular strain and actually further enhancing the rate of the water splitting. ...
Article
Full-text available
In the exploration of the optimal material for achieving the photoelectrochemical dissociation of water into hydrogen, hematite (α-Fe2O3) emerges as a highly promising candidate for proof-of-concept demonstrations. Recent studies suggest that the concurrent application of external electric fields could enhance the photoelectrochemical (PEC) process. To delve into this, we conducted nonequilibrium ab initio molecular dynamics (NE-AIMD) simulations in this study, focusing on hematite–water interfaces at room temperature under progressively stronger electric fields. Our findings reveal intriguing evidence of water molecule adsorption and dissociation, as evidenced by an analysis of the structural properties of the hydrated layered surface of the hematite–water interface. Additionally, we scrutinized intermolecular structures using radial distribution functions (RDFs) to explore the interaction between the hematite slab and water. Notably, the presence of a Grotthuss hopping mechanism became apparent as the electric field strength increased. A comprehensive discussion based on intramolecular geometry highlighted aspects such as hydrogen-bond lengths, H-bond angles, average H-bond numbers, and the observed correlation existing among the hydrogen-bond strength, bond-dissociation energy, and H-bond lifetime. Furthermore, we assessed the impact of electric fields on the librational, bending, and stretching modes of hydrogen atoms in water by calculating the vibrational density of states (VDOS). This analysis revealed distinct field effects for the three characteristic band modes, both in the bulk region and at the hematite–water interface. We also evaluated the charge density of active elements at the aqueous hematite surface, delving into field-induced electronic charge-density variations through the Hirshfeld charge density analysis of atomic elements. Throughout this work, we drew clear distinctions between parallel and antiparallel field alignments at the hematite–water interface, aiming to elucidate crucial differences in local behavior for each surface direction of the hematite–water interface.
... Generally, if the separation distance is shorter than the Coulomb attraction radius, charge carriers, upon losing their excess energy, tend to localize back to the countercharge. Monte Carlo simulations have shown that the initial charge carrier separation of the order of several nanometres helps a little for the final electron-hole separation to free charge carriers, unless this distance is comparable with the Coulomb attraction radius [74]. In contrary, other investigations reported charge carrier generation efficiency basically independent of the excess energy suggesting that the relaxed CT excitons are the major precursors of free charge carriers [75][76][77]. ...
Article
Full-text available
Charge carrier mobility in organic semiconductors is not a constant value unambigously characterizing some particular material, but depends on the electric field, temperature and even on time after it was generated or injected. The time dependence is particularly important for the thin-film devices where charge carriers pass the organic layer before mobility reaching its stationary value. Here we give a review of experimental techniques with ultrafast timeresolution enabling one to address the mobility kinetics and analyse properties of the time-dependent mobility in conjugated polymers and organic solar cells. We analyse kinetics during the charge carrier generation and extraction of free charge carriers. The mobility typically decreases by several orders of magnitude on a picosecond-nanosecond time scale; however, its kinetics also depends on the investigation technique. The mobility kinetics in blends for bulk heterojunction solar cells strongly depends on the stoichiometric ratio of donor and acceptor materials.
... 21 Thermal treatment of the BHJ exceeding the glass transition temperature (T g ) of conjugated polymers 22 results in a higher charge mobility 23−25 due to a reorganization of the materials in the polymer BHJ. 15,26,27 Further benefits observed are the increase of polymer crystallinity, 1,15,16,22 resulting in the enhancement of the optical response and charge mobility. Overall, it has been reported that the spectral response and the power conversion efficiency (PCE) of the devices can increase upon annealing on BHJ. ...
Article
In this work we focus on P3HT:PC61BM bulk heterojunction (BHJ) devices with MoO3 at the hole extraction side of the BHJ which relies on the formation of a strong dipole at the BHJ/MoO3 interface, as a reference system that has been extensively studied. We have observed depending on when the annealing is performed during device fabrication, device performance either increased or decreased due to formation of a sharp or relatively diffuse interface respectively due to diffusion of MoOx into the BHJ. The measured strength of the dipole at this interface following thermal annealing correlated well with the width of the interface and device performance, with the sharper interface resulting in a stronger dipole and in improved device performance. This is expected to be a general phenomenon for evaporated coatings onto polymeric BHJ, regardless of the polymers involved.
Article
Exciton dissociation at the interface between a conjugated polymer as an electron donor (D) and a fullerene derivative as an acceptor (A) is often considered in the dipole model that suggests existence of the dipolar layer between D and A materials. In the current article, we calculate the dissociation probability of excitons in the dipole model assuming a two-dimensional arrangement of the interfacial dipoles at a D-A interface. We find that the dipolar layer between D-A materials can radically alter the energy of the Coulomb attraction between an optically generated electron and its geminate hole. We show that the calculated dissociation probability of electron-hole pairs is in agreement with experimental field dependences of photocurrents reported previously for a bilayer solar cell consisting of C60 as an electron acceptor and a well ordered conjugated polymer as an electron donor.
Article
Given that conversion efficiencies of incident solar radiation to liquid fuels, e.g., H2, are of the order of a few percent or less, as quantified by ‘solar to hydrogen’ (STH), economically inexpensive and operationally straightforward ways to boost photo-electrochemcial (PEC) H2 production from solar-driven water splitting are important. In this work, externally-applied static electric fields have led to enhanced H2 production in an energy-efficient manner, with up to ∼30-40% increase in H2 (bearing in mind field-input energy) in a prototype, open-type solar cell featuring rutile/titania and hematite/iron-oxide (Fe2O3), respectively, in contact with an alkaline aqueous medium (corresponding to respective relative increases of STH by ∼12 and 16%). We have also performed non-equilibrium ab-initio molecular dynamics in both static electric and electromagnetic (e/m) fields, for water in contact with a hematite/iron-oxide (001) surface, observing enhanced break-up of water molecules, by up to ∼70% in the linear-response régime. We discuss the microscopic origin of such enhanced water-splitting, based on experimental and simulation-based insights. In particular, we external-field direction at the hematite surfaces, and scrutinise properties of the adsorbed water molecules and OH⁻ and H3O⁺ species, e.g., hydrogen bonds between water-protons and the hematite surfaces’ bridging oxygen atoms, as well as interactions between oxygen atoms in adsorbed water molecules and underlying iron atoms.
Article
Full-text available
Hole polaron delocalization on polymer chains helps charge separation by lowering the free energy of the spatially separated charge pair.
Article
The molecular-level arrangement of the donor and acceptor (i.e. the local morphology) in organic solar cells governs charge separation and charge transport via its effect on the mobility of charges. However, the nanometre-scale mobility in such systems, which can be measured using terahertz (THz) spectroscopy, has been little investigated at relevant low excitation densities, due to extremely weak signals. Here, we study the mobility over short distances and at ultrashort times using time-resolved optical-pump-THz-probe (OPTP) spectroscopy on pBTTT:PCBM blends. This complements our previous results obtained with transient absorption (TA) and electro-modulated differential absorption (EDA) techniques. In the pBTTT:PCBM system, the co-existence of fullerene-/polymer-rich (‘neat’) and co-crystalline (‘intermixed’) regions can be controlled through choice of composition (weight ratio of the two components, use of processing additive). We demonstrate high short-range mobilities that help charges separate, and we show how this mobility of photogenerated charges develops in time, in particular as the charges move between different phase regions of the blend. By reducing the pump fluence below the threshold for nonlinear recombination mechanisms, we access these properties at solar cell operating conditions. Overall, we explain the necessity of different local phases through their influence on charge lifetime and mobility.
Article
Full-text available
Polymer-fullerene bulk heterojunction solar cells (PSCs) are currently attracting a great deal of attention and gaining increasing importance, having already shown great promise as renewable, lightweight and low-cost energy sources. Recently, the power-conversion efficiency of state-of-the-art PSCs has exceeded 8% in the scientific literature. However, to find viable applications for this emerging photovoltaic technology, further enhancements in the efficiency towards 10% (the threshold for commercial applications) are urgently required. Here, we demonstrate highly efficient PSCs with a certified efficiency of 9.2% using an inverted structure, which simultaneously offers ohmic contact for photogenerated charge-carrier collection and allows optimum photon harvest in the device. Because of the ease of use and drastic boost in efficiency provided by this device structure, this discovery could find use in fully exploiting the potential of various material systems, and also open up new opportunities to improve PSCs with a view to achieving an efficiency of 10%.
Article
Full-text available
Solar cells based on conjugated polymer and fullerene blends have been developed as a low-cost alternative to silicon. For efficient solar cells, electron-hole pairs must separate into free mobile charges that can be extracted in high yield. We still lack good understanding of how, why and when carriers separate against the Coulomb attraction. Here we visualize the charge separation process in bulk heterojunction solar cells by directly measuring charge carrier drift in a polymer:fullerene blend with ultrafast time resolution. We show that initially only closely separated (<1 nm) charge pairs are created and they separate by several nanometres during the first several picoseconds. Charge pairs overcome Coulomb attraction and form free carriers on a subnanosecond time scale. Numerical simulations complementing the experimental data show that fast three-dimensional charge diffusion within an energetically disordered medium, increasing the entropy of the system, is sufficient to drive the charge separation process.
Article
Time-resolved terahertz spectroscopy was employed for the investigation of charge-transport dynamics in benzothiadiazolo-dithiophene polyfluorene ([2,7-(9,9-dioctyl-fluorene)-alt-5,5-(4′,7′-di-2-thienyl-2′,1′,3′-benzothiadiazole)]) (APFO-3) polymers with various chain lengths and in its monomer form, all blended with an electron acceptor ([6,6]-phenyl-C61-butyric acid methyl ester, PCBM). Upon photoexcitation, charged polaron pairs are created, negative charges are transferred to fullerenes, while positive polarons remain on polymers/monomers. Vastly different hole mobility in polymer and monomer blends allows us to distinguish the hole and electron contributions to the carrier mobility.
Article
An important question in research on polymer solar cells is the mechanism of free charge carrier generation. In organic bulk heterojunction solar cells, photoinduced electron transfer between the donor and acceptor molecules leads to population of a charge transfer (CT) state at the donor-acceptor interface. The residual electrostatic binding energy between electron and hole in the CT state may be considerable; experimental studies indicate a binding energy of 0.1-0.2 eV which is much higher than the thermal energy ( k b T = 0.025 eV). [ 1-4 ] Given the low dielectric constant of organic materials ( ε r = 3-4) charges need to be separated over a large distance (10-15 nm) before their electrostatic interaction energy becomes comparable to k b T . Even though the lifetime of the CT state can be as short as only a few nanoseconds, [ 5 , 6 ] the effi ciency of free charge carrier generation can be very high. In optimized polymer solar cells, charge carrier generation occurs with an effi ciency close to unity. [ 7 , 8 ] The mechanism by which the large energy barrier for charge separation from the CT state is overcome within the short lifetime is not known. Several explanations have been proposed, but no consensus on this issue has yet been reached. [ 9 , 10 ]
Article
Motion of photogenerated charge carriers in π-conjugated polymer films with different disorder and chain orientation and also in a blend of conducting and insulating polymers has been investigated by means of time-resolved electric field-induced second harmonic generation technique. Experimental results and Monte Carlo simulation enabled us to distinguish three time domains of charge transport. The charge carriers become separated by about 10 nm in a strong electric field within 1 ps. Subsequently, carriers drift another 10–15 nm on a picosecond time scale with a high mobility. The third and the slowest carrier motion phase is well described by the stochastic drift and determines the macroscopic equilibrium mobility. We attribute the two ultrafast drift phases to carrier motion inside a conjugated segment and along a single polymer chain, respectively, whilst the slow motion phase involves interchain jumps.
Article
The conductivity of an n-type semiconductor has been calculated in the region of low-temperature T and low impurity concentration nD. The model is that of phonon-induced electron hopping from donor site to donor site where a fraction K of the sites is vacant due to compensation. To first order in the electric field, the solution to the steady-state and current equations is shown to be equivalent to the solution of a linear resistance network. The network resistance is evaluated and the result shows that the T dependence of the resistivity is ρ∝exp(ε3/kT). For small K, ε3=(e2/κ0)(4πnD/3)1/3(1-1.35K1/3), where κ0 is the dielectric constant. At higher K, ε3 and ρ attain a minimum near K=0.5. The dependence on nD is extracted; the agreement of the latter and of ε3 with experiment is satisfactory. The magnitude of ρ is in fair agreement with experiment. The influence of excited donor states on ρ is discussed.
Article
Herein we address the factors controlling photocurrent generation in P3HT:PCBM blend films as a function of blend composition and annealing treatment. Absorption, photoluminescence, and transient absorption spectroscopy are used to distinguish the role of exciton dissociation, charge pair separation, and charge collection. Variations in blend film microstructure with composition and annealing treatment are studied using X-ray diffraction. While the trend in photocurrent generation with composition and annealing [Muller, et al., Adv. Mater. 2008, 20, 3510] does not follow the trend in exciton dissociation, it closely follows the trend in charge pair generation. Moreover, charge pair generation efficiency is positively correlated to the degree of polymer crystallization and the appearance of large domains of both polymer and fullerene phases. We argue that larger domains assist charge pair separation by increasing the probability of escape from the P3HT:PCBM interface, thus reducing geminate charge recombination.Keywords (keywords): charge generation; recombination; microstructure; solar cell; fullerenes; P3HT; transient absorption
Article
Transient photovoltage and differential charging have been used to measure the charge density and recombination rate in polymer solar cells consisting of regioregular poly(3-hexythiophene) (P3HT) blended with 1-(3-methoxycarbonyl)propyl-1-phenyl-[6,6]-methano fullerene (PCBM). Charge dynamics were found to be over an order of magnitude faster in nonannealed cells as compared to annealed cells. Numerical modeling demonstrated that physically reasonable changes in the domain size and phase segregation reproduced the change seen in the experimental results and thus suggests that measurements of recombination rate can be used as an indicator of the degree of intermixing of donor and acceptor in bulk heterojunction solar cells. Through calculation of the recombination flux at both open circuit and short circuit conditions, we determine that the increase in rate constant observed for the nonannealed devices is sufficient to explain the differences in device performance.
Article
The carrier collection efficiency (ηc) and energy conversion efficiency (ηe) of polymer photovoltaic cells were improved by blending of the semiconducting polymer with C60 or its functionalized derivatives. Composite films of poly(2-methoxy-5-(2′-ethyl-hexyloxy)-1,4-phenylene vinylene) (MEH-PPV) and fullerenes exhibit ηc of about 29 percent of electrons per photon and ηe of about 2.9 percent, efficiencies that are better by more than two orders of magnitude than those that have been achieved with devices made with pure MEH-PPV. The efficient charge separation results from photoinduced electron transfer from the MEH-PPV (as donor) to C60 (as acceptor); the high collection efficiency results from a bicontinuous network of internal donor-acceptor heterojunctions.
Article
The effects of annealing and fullerene loading in regioregular poly(3-hexylthiophene) (P3HT) and 1-(3-methoxycarbonyl)-propyl-1-phenyl-(6,6)C61 (PCBM) based bulk heterojunction photovoltaics have been investigated. Under specific loading and annealing conditions, a combination of morphological and electronic factors can be brought to play to achieve optimal filling factors, open-circuit voltage (Voc), and short-circuit current density (Jsc). We demonstrate that this occurs at surprisingly low loadings of PCBM and annealing temperatures nearing the melting point of the polymer. Further, we report power conversion efficiencies approaching 5% in the P3HT:PCBM system.