scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Polymer solar cells with enhanced open-circuit voltage and efficiency

01 Nov 2009-Nature Photonics (Nature Publishing Group)-Vol. 3, Iss: 11, pp 649-653
TL;DR: In this article, the open-circuit voltage of polymer solar cells constructed based on the structure of a low-bandgap polymer, PBDTTT, can be tuned, step by step, using different functional groups.
Abstract: Following the development of the bulk heterojunction1 structure, recent years have seen a dramatic improvement in the efficiency of polymer solar cells. Maximizing the open-circuit voltage in a low-bandgap polymer is one of the critical factors towards enabling high-efficiency solar cells. Study of the relation between open-circuit voltage and the energy levels of the donor/acceptor2 in bulk heterojunction polymer solar cells has stimulated interest in modifying the open-circuit voltage by tuning the energy levels of polymers3. Here, we show that the open-circuit voltage of polymer solar cells constructed based on the structure of a low-bandgap polymer, PBDTTT4, can be tuned, step by step, using different functional groups, to achieve values as high as 0.76 V. This increased open-circuit voltage combined with a high short-circuit current density results in a polymer solar cell with a power conversion efficiency as high as 6.77%, as certified by the National Renewable Energy Laboratory. Adding electron-withdrawing groups to the backbone of the polymer PBDTTT is shown to increase the open-circuit voltage of photovoltaic cells, resulting in a polymer solar-cell that has a certified power-conversion efficiency of 6.77%.

Summary (1 min read)

Jump to:  and [Summary]

Summary

  • All the polymers were synthesized in their laboratory and PC 70 BM was purchased from Nano-C (used as received).
  • 1,8-Diiodooctane (purchased from Sigma Aldrich, used as received) with 3% volume ratio was then added to the solutions and stirred before use.
  • The solutions were spin-coated on indium tin oxide (ITO)/glass substrates with a pre-coated PEDOT:PSS (poly(ethylenedioxythiophene):polystyrene sulphonate) layer.
  • The device active area was $0.1 cm 2 for all the solar cell devices discussed in this work.
  • Device characterization was carried out in air after encapsulation under simulated AM1.5G irradiation (100 mW cm 22 ) using a xenon-lamp-based solar simulator.
  • EQE measurements of the encapsulated devices were performed in air (PV Measurements, Model QEX7).
  • Some oxidization of the electrodes (calcium/aluminium) was observed when the devices were open to the air.

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

Polymer solar cells with enhanced open-circuit
voltage and efficiency
Hsiang-Yu Chen
1,2
, Jianhui Hou
1
*
, Shaoqing Zhang
1
,YongyeLiang
3
, Guanwen Yang
2
, Yang Yang
2
,
Luping Yu
3
,YueWu
1
*
and Gang Li
1
Following the development of the bulk heterojunction
1
structure,
recent years have seen a dramatic improvement in the efficiency
of polymer solar cells. Maximizing the open-circuit voltage in a
low-bandgap polymer is one of the critical factors towards
enabling high-efficiency solar cells. Study of the relation
between open-circuit voltage and the energy levels of the
donor/acceptor
2
in bulk heterojunction polymer solar cells has
stimulated interest in modifying the open-circuit voltage by
tuning the energy levels of polymers
3
. Here, we show that the
open-circuit voltage of polymer solar cells constructed based on
the structure of a low-bandgap polymer, PBDTTT
4
, can be tuned,
step by step, using different functional groups, to achieve values
as high as 0.76 V. This increased open-circuit voltage combined
with a high short-circuit current density results in a polymer
solar cell with a power conversion efficiency as high as 6.77%,
as certified by the National Renewable Energy Laboratory.
Polymer solar cells (PSCs) have attracted much attention due to
their potential in low-cost solar energy harvesting, as well as appli-
cations in flexible, light-weight, colourful and large-area devices.
With the discovery of efficient photo-induced electron transfer
from a conjugated polymer to fullerene
1
, the bulk heterojunction
(BHJ) PSC has become one of the most successful device structures
developed in the field to date. By simply blending polymers (electron
donors) with fullerene (electron acceptor) in organic solvents, a self-
assembling interpenetrating network can be obtained using various
coating technologies ranging from laboratory-scale spin coating or
spray coating to large-scale fabrication technologies such as inkjet
printing
5,6
,doctorblading
2
,gravure
7
, slot-die coating
8
and flexo-
graphic printing
9
. In the last few years, several effective methods
have been developed to optimize the interpenetrating network
formed by the electron donor and acceptor, including solvent
annealing (or slow-growth)
10
, thermal annealing
11–13
and mor-
phology control using mixed solvent mixtures
14
or additives
15
in
the solutions of donor/acceptor blends. Poly(3-hexylthiophene)
(P3HT) in particular has been subject to increasing interest in the
polymer research community, but significant progress has also been
made in developing new active-layer polymer materials
4,15–22
.Since
around 2008, the efficiency of PSCs has risen to 6% using new con-
jugated polymers as electron donors
19
. Although progress has been
impressive, there is still much to do before the realization of practical
applications of PSCs. Many factors need to be taken into account in
efficiently converting sunlight into electricity. The absorption range,
the photon–electron conversion rate and the carrier mobilities of
the light-harvesting polymers are among the crucial parameters for
achieving high-efficiency solar cells. Furthermore, fabricating large-
area devices without significantly losing efficiency while maintaining
long device lifetimes remains challenging
23
.
In principle, the strategies used to improve BHJ solar cell effi-
ciency include (i) reducing the bandgap of polymers so as to
harvest more sunlight, which leads to higher short-circuit current
density (J
sc
) and (ii) lowering the highest occupied molecular
orbital (HOMO) of the polymers, which increases the open-
circuit voltage (V
oc
). With the rise in interest in using low-
bandgap polymers to harvest more sunlight from longer wave-
lengths, much effort has been made recently in reducing the
bandgap of polymers. Others
4,15,21
have reported PSCs with power
conversion efficiencies (PCE) of over 5% using different low-
bandgap polymers. The extended absorption of sunlight at longer
wavelengths directly reflects on the value of J
sc
, and a current
density of up to 16 mA cm
22
has been achieved
15
. On the other
hand, PSCs with high V
oc
have been realized by other
groups
19,20,22
by using polymers that absorb at shorter wavelengths.
To push the PCE of PSCs towards the predicted theoretical limit-
ation
24
, however, achieving both a high J
sc
and a high V
oc
is critical,
indeed essential. To match the energy level of the commonly used
electron a cceptor [6,6]-phenyl-C
61
-butyric acid methyl ester (PCBM),
both the HOMO and the lowest unoccupied molecular orbital
(LUMO) of the polymer need to be considered while tuning the
bandgap of polymers. It is known that the energy difference
between the LUMOs of donor and acceptor should be larger than
0.3 eV for efficient charge separation
24
, which directly relates to
the J
sc
of solar cells. However, the V
oc
of PSCs is limited by the
difference between the HOMO of the donor and the LUMO of
the acceptor
2
. As a result,narrowingthebandgapofpolymerswithout
sacrificing efficient charge separation as well as high V
oc
becomes a
major hurdle in achieving high-efficiency PSCs. In this work, we
attempt to alter the HOMO of poly[4,8-bis-substituted-benzo
[1,2-b:4,5-b
0
]dithiophene-2,6-diyl-alt-4-substituted-thieno[3, 4-b]thio-
phene-2,6-diyl] (PBDTTT)-derived polymers
4
by adding different
electron-withdrawing functional groups, step by step. We have
shown that the addition of more than one electron-withdrawing
group is effective in further lowering the HOMO of PBDTTT.
As reported, PSCs based on the copolymer of benzo[1,2-b:4,5-
b
0
]dithiophene and thieno[3,4-b]thiophene (ref. 4) (hereafter
referred to as PBDTTT–E, Fig. 1) can yield a J
sc
greater than
15 mA cm
22
with a V
oc
of 0.6 V. We chose this polymer system
and tried to increase PSC performance by increasing the V
oc
through molecular design. Previous studies on thiophene-based
polymers have shown that the alkyloxy chain has a much stronger
electron-donating effect than an alkyl chain
25
. As a result, the
HOMO of poly(3-alkoxythiophene) is higher than that of poly(3-
alkylthiophene). Based on this knowledge, we replaced the alkyloxy
group on the carbonyl of the thieno[3,4-b]thiophene unit with an
alkyl side chain (hereafter referred to as PBDTTT–C). Both
1
Solarmer Energy Inc., El Monte, California 91731, USA,
2
Department of Materials Science and Engineering, University of California, Los Angeles,
Los Angeles, California 90095, USA,
3
Department of Chemistry and the James Franck Institute, The University of Chicago, 929 East 57th Street, Chicago,
Illinois 60637, USA.
*
e-mail: jianhuih@solarmer.com; yuew@solarmer.com
LETTERS
PUBLISHED ONLINE: 25 OCTOBER 2009 | DOI: 10.1038/NPHOTON.2009.192
NATURE PHOTONICS | VOL 3 | NOVEMBER 2009 | www.nature.com/naturephotonics 649
© 2009 Macmillan Publishers Limited. All rights reserved.

PBDTTT–E and PBDTTT–C (Fig. 1a) were thus synthesized and
their HOMO and LUMO levels measured by means of electroche-
mical cyclic voltammetry (CV). Lower HOMO and LUMO levels
were observed for PBDTTT–C than for PBDTTT–E, as shown in
Fig. 1b. Interestingly, the alkyl group not only results in a lower
HOMO but also a lower LUMO; both the HOMO and the
LUMO of PBDTTT–C are 0.1 eV lower than those of
PBDTTT–E. Therefore, the bandgap of PBDTTT–C is approxi-
mately the same as that of PBDTTT–E, which is supported by
their absorption spectra (Fig. 1c).
In addition to grafting side chains, substitution of the carbon
atom in selected locations also affects the energy levels of a
polymer. In the recent work
26
, higher values of V
oc
are observed
when fluorine, an atom of high electron affinity, is introduced to
the thieno[3,4-b]thiophene unit, a PCE of 6.1% having been
demonstrated
26
. To this end, PBDTTT–C was modified with a flu-
orine atom to lower its HOMO level. The structure of the designed
and synthesized PBDTTT–CF is shown in Fig. 1a. The HOMO and
LUMO of PBDTTT–CF were measured and compared with
PBDTTT–C (Fig. 1b). As was observed with the addition of the
alkyl group, the introduction of fluorine further lowered both the
HOMO and LUMO levels. It is rather interesting that the HOMO
and LUMO of this polymer system have this peculiar property. It
is known that the application of a functional group to a polymer
backbone causes either the HOMO or LUMO level to shift.
However, to our knowledge, a change in both the HOMO and
LUMO levels with just one synthetic modification to the polymer
backbone has not been reported in any polymer system. The
reason for this simultaneous change is not clear and is still under
investigation. The absorption spectra of polymer films based on
these three polymers show similar absorption edges (at
770 nm), suggesting similar bandgaps. This is consistent with
the CV results.
PSC devices based on these three polymers were fabricated and
tested under simulated 100 mW cm
22
AM1.5G illumination (see
Methods). The optimized weight ratios of polymer to PC
70
BM for
PBDTTT–E, PBDTTT–C and PBDTTT–CF are 1:1, 1:1.5 and
1:1.5, respectively. Device current density/voltage (JV) character-
istics are shown in Fig. 2a and the parameters listed in Table 1.
More than 200 devices were fabricated and their highest efficiencies
and average values compared to provide numerical data. From the
S
S
O
O
S
O
S
O
n
S
S
O
O
S
O
S
n
S
S
O
O
S
O
S
n
F
PBDTTT–E PBDTTT–CF
−7.0
−6.5
−6.0
−5.5
−5.0
−4.5
−4.0
−3.5
−3.0
−2.5
−2.0
a
b
c
PBDTTT−E
PBDTTT−CF
PBDTTT−C
PCBM
−6.10
−4.30
−5.22
−3.45
−5.12
−3.35
−5.01
−3.24
Energy level (eV)
300 400 500 600 700 800 900
0.0
0.2
0.4
0.6
0.8
1.0
PBDTTT−E
PBDTTT−C
PBDTTT−CF
Normalized absorption (a.u.)
Wavelength (nm)
PBDTTT–C
Figure 1 | Comparison of three different polymers. ac, Chemical structure
(a) , energy levels (b) and absorption spectra (c) of PBDTTT–E, PBDTTT–C
and PBDTTT–CF. Different HOMOs and LUMOs were obtained when
different functional groups were attached to the PBDTTT backbone.
Nevertheless, the absorption edges of these three polymers are about
the same.
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
−20
−16
−12
−8
−4
0
4
a
b
PBDTTT−E
PBDTTT−C
PBDTTT−CF
Current density (mA cm
−2
)
Bias (V)
Voltage (V)
−0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
−0.1
Current (mA)
V
oc
= 0.7631 V
I
sc
= 0.63240 mA
J
sc
= 13.364 mA cm
−2
Fill factor = 66.39%
I
max
= 0.51800 mA
V
max
= 0.6185 V
P
max
= 0.32030 mW
Eciency = 6.77%
Figure 2 | Characterization of devices based on PBDTTT–E, PBDTTT–C and
PBD TTT–CF. a, Current density v ersus voltage (JV) curves obtained from
our laboratory . b, JV curve of a device based on PBDTTT–CF certified by the
NREL. A significant increase in the open-circuit voltage (V
oc
) is clearly seen
between the PBDTTT–E and PBDTTT–CF. A V
oc
of up to 0.76 V was
observed in devices based on PBDTTT–CF.
LETTERS
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2009.192
NATURE PHOTONICS | VOL 3 | NOVEMBER 2009 | www.nature.com/naturephotonics650
© 2009 Macmillan Publishers Limited. All rights reserved.

JV curves, a significant increase in V
oc
is clearly observed from
PBDTTT–E to PBDTTT–CF. A V
oc
as high as 0.76 V was observed
in devices based on PBDTTT–CF. Combined with its high J
sc
and fill
factor (FF), a high PCE of 7.38+0.4% (a 5% device variation),
measured in more than 75 devices, was achieved in the PBDTTT–
CF system, the highest measured PCE being 7.73%. Devices were
then encapsulated and sent to the National Renewable Energy
Laboratory (NREL) for certification. A certified efficiency of
6.77% (the highest efficiency achieved so far for organic solar
cells) was rated, which differs by 8% from the average obtained
efficiency of 7.38%. The current/voltage curve and corresponding
parameters are shown in Fig. 2b. Among the parameters, the drop
in efficiency results predominantly from the lower J
sc
, most prob-
ably due to degradation of the device
27
. Although the devices were
encapsulated with UV epoxy before shipping to NREL, this non-
ideal encapsulation process has been observed to cause an efficiency
drop of 3% over a 10-day storage period in a nitrogen-filled glove-
box (,0.1 ppm O
2
and H
2
O). Furthermore, part of the device metal
finger (non-active area) was exposed to air during wiring for
measurement purposes. Oxidation of the calcium/aluminium elec-
trodes when exposed to air could also contribute to this difference.
To further confirm the accuracy of the measurements, the external
quantum efficiency (EQE) of the devices based on the three polymers
was measured using an EQE system (Model QEX7) purchased from
PV Measurements. The EQE curves are shown in Fig. 3a. All the
devices show rather efficient photoconversion efficiency in the
range 400–700 nm, with EQE values of 50–70%. In the PBDTTT–
CF device, the highest EQE value was 68.7% at 630 nm, which, to
the best of our knowledge, is the highest achieved in a low-
bandgap PSCs system. The J
sc
values were then calculated by
integrating the EQE data with an AM1.5G reference spectrum. The
calculated J
sc
values were 213.0, 214.1 and 215.0 mA cm
22
for
devices based on PBDTTT–E, PBDTTT–C and PBDTTT–CF,
respectively. These values are rather consistent with those (within
4% error, see Table 1) obtained from the JV measurement.
In fact, the J
sc
values for the PBDTTT–CF solar cell obtained using
these two methods are rather close (215.0 and 215.2 mA cm
22
,
1% difference). The absorption curves of polymer:PC
70
BM blend
films are shown in Fig. 3b. Significant absorption increases are
observed in the absorption range 300–600 nm after blending with
PC
70
BM. According to the absorption spectrum, the polymer films
still have 50% transparency (optical density, OD 0.3) under opti-
mized device conditions, suggesting their potential as a material for
tandem or stackable solar cells
28
. The internal quantum efficiency
(IQE) of the device based on PBDTTT–CF was obtained from its
absorption spectrum and EQE curve. As shown in Fig. 3c, the
average IQE was higher than 90% in the range 400–700 nm, indicat-
ing a highly efficient overall photoconversion process in the cell. To
reach such a high IQE, it is necessary to simultaneously achieve effi-
cient light absorption, exciton diffusion, charge separation and
carrier collection. Physically, this indicates a close to ideal polymer
active layer morphology, with a bicontinuous interpenetrating poly-
mer/fullerene network. Combined with the high V
oc
and FF, a solar
cell with efficiency greater than 7% was achieved. The carrier mobi-
lities of these three devices were measured using the space charge
Ta b le 1 | Solar cell parameters of devices based on the three different polymers.
LUMO HOMO V
oc
(V) J
sc
(mA cm
22
)FF(%) PCE(%)
Best Ave
PBDTTT–E 23.24 25.01 0.62 213.2 63 5.15 4.8
PBDTTT–C 23.35 25.12 0.7 214.7 64.1 6.58 6.3
PBDTTT–CF 23.45 25.22 0.76 215.2 66.9 7.73 7.4
LUMO, lowest unoccupied molecular orbital; HOMO, highest occupied molecular orbital; FF, fill factor, PCE, power conversion efficiency.
300 400 500 600 700 800 900
0
10
20
30
40
50
60
70
80
90
a
b
c
PBDTTT−E
PBDTTT−C
PBDTTT−CF
EQE (%)
Wavelength (nm)
400 500 600 700 800
0.0
0.1
0.2
0.3
0.4
OD
Wavelength (nm)
400 500 600 700 800
Wavelength (nm)
PBDTTT−E
PBDTTT−C
PBDTTT−CF
0.0
0.2
0.4
0.6
0.8
1.0
0
20
40
60
80
100
OD
Quantum eciency (%)
Absorbance
EQE
IQE
PBDTTT−CF
Figure 3 | Photoconversion efficiency study. ac, External quantum
efficiency (EQE) (a) and absorption (OD, optical density) (b) for devices
based on PBDTTT–E, PBDTTT–C and PBDTTT–CF, and internal quantum
efficiency (IQE) for device based on PBDTTT–CF (c). The average IQE is
greater than 90% in the range 400–700 nm. Near 100% IQE was obtained
for the device based on PBDTTT–CF, indicating a highly efficient o v er all
photoconversion process in the cell.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2009.192
LETTERS
NATURE PHOTONICS | VOL 3 | NOVEMBER 2009 | www.nature.com/naturephotonics 651
© 2009 Macmillan Publishers Limited. All rights reserved.

limited current (SCLC) method reported previously
29
. The hole
mobilities obtained using this method were 4 10
24
,2 10
24
and 7 10
24
cm
2
V
21
s
21
for PBDTTT–E, PBDTTT–C and
PBDTTT–CF devices, respectively. The observed variation in the
hole mobilities of these three polymers may be the reason for the
difference in thickness existing in the optimized devices (790,
660 and 970 Å, respectively).
The nanoscale morphologies of the polymer/PC
70
BM films were
studied using tapping-mode atomic force microscopy (AFM).
Surface topography (left) and phase images (right) were taken for
each film and are shown in Fig. 4. Surface roughness values
measured from the topography images were 0.96, 0.92 and
0.84 nm for PBDTTT–E (Fig. 4a), PBDTTT–C (Fig. 4b) and
PBDTTT–CF (Fig. 4c) films, respectively. Very different mor-
phologies were observed for these three polymers in their phase
images (Fig. 4, right panels). As shown in Fig. 4, fibrillar features
are clearly seen in PBDTTT–E (Fig. 4a) and PBDTTT–C (Fig. 4b)
films, whereas domains with different shapes are observed in the
PBDTTT–CF (Fig. 4c) film. Fibrillar features can be clearly seen
even in the topography image of the PBDTTT–C film. The very
different morphology of these three polymer films suggests that
the interactions between the molecules of each polymer may be
different. The fact that the polymer system (PBDTTT–CF) with see-
mingly lowest organization shows the highest performance differs
20.0 nm
10.0 nm
20.0 nm
20.0°
20.0°
20.0°
0 Height 500 nm 0 Phase 500 nm
0 Height 500 nm
0
Phase 500 nm
0 Height 500 nm 0 Phase 500 nm
a
b
c
Figure 4 | Tapping-mode atomic force microscopy images of the three films used in making the devices (under optimized device conditions).
ac, The topography of ea ch film is sho wn in the left panels, and the corr esponding phase images in the right panels. Very different surface morphologies
were obtained for the different polymer films. Fibrillar features are clearly seen in PBDTTT–E (a) and PBDTTT–C (b) films, whereas domains with differ ent
shapes are observed in the PBD TTT–CF (c) film. The brillar features can be clearly seen even in the topography image of the PBD TTT–C film.
LETTERS
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2009.192
NATURE PHOTONICS | VOL 3 | NOVEMBER 2009 | www.nature.com/naturephotonics652
© 2009 Macmillan Publishers Limited. All rights reserved.

greatly from the widely studied benchmark P3HT solar cell
system
30,31
, in which the noodle-like structure typically correlates
very well to the crystallinity
32
as well as the PCE. This indicates
that the correlation of the morphology and PCE in the new
polymer systems differs significantly from conventional under-
standing. Further investigation is under way to provide
more information.
To conclude, tuning the V
oc
of PSCs by means of molecular
design has been realized using a step-by-step approach. Applying
stronger electron-withdrawing groups to the backbone of polymers
has been found to be effective in lowering the HOMO of polymers,
which directly affects the V
oc
of PSCs. Based on the PBDTTT
polymer derivative system, PSCs with a PCE higher than 7% have
been realized by combining the advantages of a low HOMO level
in the polymer (high V
oc
) and long wavelength absorption (high
J
sc
). Regarding the LUMO level of PCBM, there is still much
capacity for increasing the V
oc
by tuning the energy levels of the
polymers. A further improvement in efficiency can be expected if
the HOMO of the polymer can be further lowered without loss of J
sc
.
Methods
All the polymers were synthesized in our laboratory and PC
70
BM was purchased
from Nano-C (used as received). Polymers and PC
70
BM were then dissolved in
chlorobenzene with 1:1 (PBDTTT–E:PC
70
BM ¼ 10 mg ml
21
:10 mg ml
21
) and
1:1.5 (10 mg ml
21
:15 mg ml
21
) weight ratio, respectively. 1,8-Diiodooctane
(purchased from Sigma Aldrich, used as received) with 3% volume ratio was then
added to the solutions and stirred before use. The solutions were spin-coated on
indium tin oxide (ITO)/glass substrates with a pre-coated PEDOT:PSS
(poly(ethylenedioxythiophene):polystyrene sulphonate) layer. The device active area
was 0.1 cm
2
for all the solar cell devices discussed in this work. Device
characterization was carried out in air after encapsulation under simulated AM1.5G
irradiation (100 mW cm
22
) using a xenon-lamp-based solar simulator.
EQE measurements of the encapsulated devices were performed in air
(PV Measurements, Model QEX7). Some oxidization of the electrodes
(calcium/aluminium) was observed when the devices were open to the air.
Received 3 August 2009; accepted 28 September 2009;
published online 25 October 2009
References
1. Sariciftci, N. S., Smilowitz, L., Heeger, A. J. & Wudl, F. Photoinduced electron-
transfer from a conducting polymer to buckminsterfullerene. Science 258,
1474–1476 (1992).
2. Brabec, C. J. et al. Origin of the open circuit voltage of plastic solar cells. Adv.
Funct. Mater. 11, 374–380 (2001).
3. Hou, J. H. et al. Bandgap and molecular energy level control of conjugated
polymer photovoltaic materials based on benzo[1,2-b:4,5-b’]dithiophene.
Macromolecules 41, 6012–6018 (2008).
4. Liang, Y. Y. et al. Development of new semiconducting polymers for high
performance solar cells. J. Am. Chem. Soc. 131, 56–57 (2009).
5. Hoth, C. N., Choulis, S. A., Schilinsky, P. & Brabec, C. J. High photovoltaic
performance of inkjet printed polymer: fullerene blends. Adv. Mater. 19,
3973–3978 (2007).
6. Aernouts, T., Aleksandrov, T., Girotto, C., Genoe, J. & Poortmans, J. Polymer
based organic solar cells using ink-jet printed active layers. Appl. Phys. Lett.
92, 033306 (2008).
7. Pudas, M., Hagberg, J. & Leppavuori, S. Gravure offset printing of polymer inks
for conductors. Progr. Org. Coatings 49, 324–335 (2004).
8. Krebs, F. C., Gevorgyan, S. A. & Alstrup, J. A roll-to-roll process to flexible
polymer solar cells: model studies, manufacture and operational stability studies.
J. Mater. Chem. 19, 5442–5451 (2009).
9. Krebs, F. C. Fabrication and processing of polymer solar cells: a review of
printing and coating techniques. Sol. Eng. Mater. Sol. Cells 93, 394–412 (2009).
10. Li, G. et al. High-efficiency solution processable polymer photovoltaic cells by
self-organization of polymer blends. Nature Mater. 4, 864–868 (2005).
11. Ma, W. L., Yang, C. Y., Gong, X., Lee, K. & Heeger, A. J. Thermally stable,
efficient polymer solar cells with nanoscale control of the interpenetrating
network morphology. Adv. Funct. Mater. 15, 1617–1622 (2005).
12. Padinger, F., Rittberger, R. S. & Sariciftci, N. S. Effects of postproduction
treatment on plastic solar cells. Adv. Funct. Mater. 13, 85–88 (2003).
13. Li, G., Shrotriya, V., Yao, Y. & Yang, Y. Investigation of annealing effects and film
thickness dependence of polymer solar cells based on poly(3-hexylthiophene).
J. Appl. Phys. 98, 043704 (2005).
14. Zhang, F. L. et al. Influence of solvent mixing on the morphology and
performance of solar cells based on polyfluorene copolymer/fullerene blends.
Adv. Funct. Mater. 16, 667–674 (2006).
15. Peet, J. et al. Efficiency enhancement in low-bandgap polymer solar cells by
processing with alkane dithiols. Nature Mater. 6, 497–500 (2007).
16. Chen, C. P., Chan, S. H., Chao, T. C., Ting, C. & Ko, B. T. Low-bandgap
poly(thiophene–phenylene–thiophene) derivatives with broaden absorption
spectra for use in high-performance bulk-heterojunction polymer solar cells.
J. Am. Chem. Soc. 130, 12828–12833 (2008).
17. Blouin, N., Michaud, A. & Leclerc, M. A low-bandgap poly(2,7-carbazole)
derivative for use in high-performance solar cells. Adv. Mater. 19,
2295–2300 (2007).
18. Gadisa, A. et al. A new donor–acceptor–donor polyfluorene copolymer
with balanced electron and hole mobility. Adv. Funct. Mater. 17,
3836–3842 (2007).
19. Park, S. H. et al. Bulk heterojunction solar cells with internal quantum efficiency
approaching 100%. Nature Photon. 3, 297–303 (2009).
20. Chen, M. H. et al. Efficient polymer solar cells with thin active layers based on
alternating polyfluorene copolymer/fullerene bulk heterojunctions. Adv. Mater.
21, 1–5 (2009).
21. Hou, J., Chen, H.-Y., Zhang, S., Li, G. & Yang, Y. Synthesis, characterization and
photovoltaic properties of a low bandgap polymer based on silole-containing
polythiophenes and benzo[c][1,2,5]thiadiazole. J. Am. Chem. Soc. 130,
16144–16145 (2008).
22. Wang, E. et al. High-performance polymer heterojunction solar cells of a
polysilafluorene derivative. Appl. Phys. Lett. 92, 033307 (2008).
23. Krebs, F. C. et al. A complete process for production of flexible large area
polymer solar cells entirely using screen printing—first public demonstration.
Sol. Eng. Mater. Sol. Cells 93, 422–441 (2009).
24. Scharber, M. C. et al. Design rules for donors in bulk-heterojunction solar cells—
towards 10% energy-conversion efficiency. Adv. Mater. 18, 789–794 (2006).
25. Shi, C. J., Yao, Y., Yang, Y. & Pei, Q. B. Regioregular copolymers of 3-
alkoxythiophene and their photovoltaic application. J. Am. Chem. Soc. 128,
8980–8986 (2006).
26. Liang, Y. Y. et al. Highly efficient solar cell polymers developed via
fine-tuning of structural and electronic properties. J. Am. Chem. Soc. 131,
7792–7799 (2009).
27. Jorgensen, M., Norrman, K. & Krebs, F. C. Stability/degradation of polymer
solar cells. Sol. Eng. Mater. Sol. Cells 92, 686–714 (2008).
28. Shrotriya, V., Wu, E. H. E., Li, G., Yao, Y. & Yang, Y. Efficient light harvesting
in multiple-device stacked structure for polymer solar cells. Appl. Phys. Lett.
88, 064104 (2006).
29. Shrotriya, V., Yao, Y., Li, G. & Yang, Y. Effect of self-organization in
polymer/fullerene bulk heterojunctions on solar cell performance. Appl. Phys.
Lett. 89, 063505 (2006).
30. Li, G. et al. ‘Solvent annealing’ effect in polymer solar cells based on
poly(3-hexylthiophene) and methanofullerenes. Adv. Funct. Mater. 17,
1636–1644 (2007).
31. Li, G., Shrotriya, V., Yao, Y., Huang, J. S. & Yang, Y. Manipulating regioregular
poly(3-hexylthiophene): [6,6]-phenyl-C-61-butyric acid methyl ester blends—
route towards high efficiency polymer solar cells. J. Mater. Chem. 17,
3126–3140 (2007).
32. Chu, C. W. et al . Control of the nanoscale crystallinity and phase separation in
polymer solar cells. Appl. Phys. Lett. 92,
103306 (2008).
Acknowledgements
The National Renewable Energy Laboratory (NREL) is thanked for conducting the
certification of devices. The authors in particular thank D.C. Olson at NREL for his help in
verifying and certifying the performances of our devices.
Author contributions
H.Y.C. conceived and performed PSC fabrication, measurements and data analysis. Y.L.
and L.Y. contributed to the design of the polymer’s main chain structure. J.H. designed the
polymer structures and synthesis routes. S.Z. synthesized the polymers. G.Y. performed the
AFM image scans. J.H. and Y.W. conceptualized and directed the research project. All
authors discussed the results and commented on the manuscript.
Additional information
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/. Correspondence and requests for materials should be addressed to
J.H. and Y.W.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2009.192
LETTERS
NATURE PHOTONICS | VOL 3 | NOVEMBER 2009 | www.nature.com/naturephotonics 653
© 2009 Macmillan Publishers Limited. All rights reserved.
Citations
More filters
Journal ArticleDOI
TL;DR: In this article, a review summarizes recent progress in the development of polymer solar cells and provides a synopsis of major achievements in the field over the past few years, while potential future developments and the applications of this technology are also briefly discussed.
Abstract: This Review summarizes recent progress in the development of polymer solar cells. It covers the scientific origins and basic properties of polymer solar cell technology, material requirements and device operation mechanisms, while also providing a synopsis of major achievements in the field over the past few years. Potential future developments and the applications of this technology are also briefly discussed.

3,832 citations

Journal ArticleDOI
TL;DR: In this article, the authors showed that PFN can be incorporated into polymer light-emitting devices (PLEDs) to enhance electron injection from high-work-function metals such as aluminium (work function w of 4.3 eV) and gold (w ¼ 5.2 eV).
Abstract: typically based on n-type metal oxides, our device is solutionprocessed at room temperature, enabling easy processibility over a large area. Accordingly, the approach is fully amenable to highthroughput roll-to-roll manufacturing techniques, may be used to fabricate vacuum-deposition-free PSCs of large area, and find practical applications in future mass production. Moreover, our discovery overturns a well-accepted belief (the inferior performance of inverted PSCs) and clearly shows that the characteristics of high performance, improved stability and ease of use can be integrated into a single device, as long as the devices are optimized, both optically and electrically, by means of a meticulously designed device structure. We also anticipate that our findings will catalyse the development of new device structures and may move the efficiency of devices towards the goal of 10% for various material systems. Previously, we reported that PFN can be incorporated into polymer light-emitting devices (PLEDs) to enhance electron injection from high-work-function metals such as aluminium (work function w of 4.3 eV) 22,23 and has thus been used to realize high-efficiency, air-stable PLEDs 24 . Furthermore, we also found that efficient electron injection can be obtained even in the most noble metals with extremely high work functions, such as gold (w ¼ 5.2 eV), by lowering the effective work function (for example lowering w in gold by 1.0 eV), which has previously been ascribed to the formation of a strong interface dipole 25 .

3,651 citations

Journal ArticleDOI
TL;DR: The uncovered aggregation and design rules yield three high-efficiency (>10%) donor polymers and will allow further synthetic advances and matching of both the polymer and fullerene materials, potentially leading to significantly improved performance and increased design flexibility.
Abstract: Although the field of polymer solar cell has seen much progress in device performance in the past few years, several limitations are holding back its further development For instance, current high-efficiency (>90%) cells are restricted to material combinations that are based on limited donor polymers and only one specific fullerene acceptor Here we report the achievement of high-performance (efficiencies up to 108%, fill factors up to 77%) thick-film polymer solar cells for multiple polymer:fullerene combinations via the formation of a near-ideal polymer:fullerene morphology that contains highly crystalline yet reasonably small polymer domains This morphology is controlled by the temperature-dependent aggregation behaviour of the donor polymers and is insensitive to the choice of fullerenes The uncovered aggregation and design rules yield three high-efficiency (>10%) donor polymers and will allow further synthetic advances and matching of both the polymer and fullerene materials, potentially leading to significantly improved performance and increased design flexibility

2,839 citations


Additional excerpts

  • ...F F PffBT4T-2DT (FBT-Th4[1,4]) PffBT4T-1ON...

    [...]

Journal ArticleDOI
TL;DR: The development of a high-performance low bandgap polymer that enables a solution processed tandem solar cell with certified 10.6% power conversion efficiency under standard reporting conditions, which is the first certified polymer solar cell efficiency over 10%.
Abstract: An effective way to improve polymer solar cell efficiency is to use a tandem structure, as a broader part of the spectrum of solar radiation is used and the thermalization loss of photon energy is minimized. In the past, the lack of high-performance low-bandgap polymers was the major limiting factor for achieving high-performance tandem solar cell. Here we report the development of a high-performance low bandgap polymer (bandgap 60% and spectral response that extends to 900 nm, with a power conversion efficiency of 7.9%. The polymer enables a solution processed tandem solar cell with certified 10.6% power conversion efficiency under standard reporting conditions (25 °C, 1,000 Wm(-2), IEC 60904-3 global), which is the first certified polymer solar cell efficiency over 10%.

2,708 citations

Journal ArticleDOI
TL;DR: This Account discusses the basic requirements and scientific issues in the molecular design of high efficiency photovoltaic molecules, and summarizes recent progress in electronic energy level engineering and absorption spectral broadening of the donor and acceptor photvoltaic materials by my research group and others.
Abstract: Bulk heterojunction (BHJ) polymer solar cells (PSCs) sandwich a blend layer of conjugated polymer donor and fullerene derivative acceptor between a transparent ITO positive electrode and a low work function metal negative electrode. In comparison with traditional inorganic semiconductor solar cells, PSCs offer a simpler device structure, easier fabrication, lower cost, and lighter weight, and these structures can be fabricated into flexible devices. But currently the power conversion efficiency (PCE) of the PSCs is not sufficient for future commercialization. The polymer donors and fullerene derivative acceptors are the key photovoltaic materials that will need to be optimized for high-performance PSCs.In this Account, I discuss the basic requirements and scientific issues in the molecular design of high efficiency photovoltaic molecules. I also summarize recent progress in electronic energy level engineering and absorption spectral broadening of the donor and acceptor photovoltaic materials by my researc...

2,527 citations

References
More filters
Journal ArticleDOI
TL;DR: In this article, the authors report highly efficient polymer solar cells based on a bulk heterojunction of polymer poly(3-hexylthiophene) and methanofullerene.
Abstract: Converting solar energy into electricity provides a much-needed solution to the energy crisis the world is facing today. Polymer solar cells have shown potential to harness solar energy in a cost-effective way. Significant efforts are underway to improve their efficiency to the level of practical applications. Here, we report highly efficient polymer solar cells based on a bulk heterojunction of polymer poly(3-hexylthiophene) and methanofullerene. Controlling the active layer growth rate results in an increased hole mobility and balanced charge transport. Together with increased absorption in the active layer, this results in much-improved device performance, particularly in external quantum efficiency. The power-conversion efficiency of 4.4% achieved here is the highest published so far for polymer-based solar cells. The solution process involved ensures that the fabrication cost remains low and the processing is simple. The high efficiency achieved in this work brings these devices one step closer to commercialization.

5,431 citations


"Polymer solar cells with enhanced o..." refers methods in this paper

  • ...In the last few years, several effective methods have been developed to optimize the interpenetrating network formed by the electron donor and acceptor, including solvent annealing (or slow-growth...

    [...]

Journal ArticleDOI
TL;DR: In this article, the authors presented a review of several organic photovoltaics (OPV) technologies, including conjugated polymers with high-electron-affinity molecules like C60 (as in the bulk-heterojunction solar cell).
Abstract: There has been an intensive search for cost-effective photovoltaics since the development of the first solar cells in the 1950s. [1–3] Among all alternative technologies to silicon-based pn-junction solar cells, organic solar cells could lead the most significant cost reduction. [4] The field of organic photovoltaics (OPVs) comprises organic/inorganic nanostructures like dyesensitized solar cells, multilayers of small organic molecules, and phase-separated mixtures of organic materials (the bulkheterojunction solar cell). A review of several OPV technologies has been presented recently. [5] Light absorption in organic solar cells leads to the generation of excited, bound electron– hole pairs (often called excitons). To achieve substantial energy-conversion efficiencies, these excited electron–hole pairs need to be dissociated into free charge carriers with a high yield. Excitons can be dissociated at interfaces of materials with different electron affinities or by electric fields, or the dissociation can be trap or impurity assisted. Blending conjugated polymers with high-electron-affinity molecules like C60 (as in the bulk-heterojunction solar cell) has proven to be an efficient way for rapid exciton dissociation. Conjugated polymer–C60 interpenetrating networks exhibit ultrafast charge transfer (∼40 fs). [6,7] As there is no competing decay process of the optically excited electron–hole pair located on the polymer in this time regime, an optimized mixture with C60 converts absorbed photons to electrons with an efficiency close to 100%. [8] The associated bicontinuous interpenetrating network enables efficient collection of the separated charges at the electrodes. The bulk-heterojunction solar cell has attracted a lot of attention because of its potential to be a true low-cost photovoltaic technology. A simple coating or printing process would enable roll-to-roll manufacturing of flexible, low-weight PV modules, which should permit cost-efficient production and the development of products for new markets, e.g., in the field of portable electronics. One major obstacle for the commercialization of bulk-heterojunction solar cells is the relatively small device efficiencies that have been demonstrated up to now. [5] The best energy-conversion efficiencies published for small-area devices approach 5%. [9–11] A detailed analysis of state-of-the-art bulk-heterojunction solar cells [8] reveals that the efficiency is limited by the low opencircuit voltage (Voc) delivered by these devices under illumination. Typically, organic semiconductors with a bandgap of about 2 eV are applied as photoactive materials, but the observed open-circuit voltages are only in the range of 0.5–1 V. There has long been a controversy about the origin of the Voc in conjugated polymer–fullerene solar cells. Following the classical thin-film solar-cell concept, the metal–insulator–metal (MIM) model was applied to bulk-heterojunction devices. In the MIM picture, Voc is simply equal to the work-function difference of the two metal electrodes. The model had to be modified after the observation of the strong influence of the reduction potential of the fullerene on the open-circuit volt

4,816 citations


"Polymer solar cells with enhanced o..." refers background in this paper

  • ...To push the PCE of PSCs towards the predicted theoretical limitatio...

    [...]

Journal ArticleDOI
TL;DR: By applying specific fabrication conditions summarized in the Experimental section and post-production annealing at 150°C, polymer solar cells with power-conversion efficiency approaching 5% were demonstrated.
Abstract: By applying the specific fabrication conditions summarized in the Experimental section and post-production annealing at 150 °C, polymer solar cells with power-conversion efficiency approaching 5 % are demonstrated. These devices exhibit remarkable thermal stability. We attribute the improved performance to changes in the bulk heterojunction material induced by thermal annealing. The improved nanoscale morphology, the increased crystallinity of the semiconducting polymer, and the improved contact to the electron-collecting electrode facilitate charge generation, charge transport to, and charge collection at the electrodes, thereby enhancing the device efficiency by lowering the series resistance of the polymer solar cells.

4,513 citations

Journal ArticleDOI
27 Nov 1992-Science
TL;DR: Because the photoluminescence in the conducting polymer is quenched by interaction with C60, the data imply that charge transfer from the excited state occurs on a picosecond time scale.
Abstract: Evidence for photoinduced electron transfer from the excited state of a conducting polymer onto buckminsterfullerene, C(60), is reported. After photo-excitation of the conjugated polymer with light of energy greater than the pi-pi* gap, an electron transfer to the C(60) molecule is initiated. Photoinduced optical absorption studies demonstrate a different excitation spectrum for the composite as compared to the separate components, consistent with photo-excited charge transfer. A photoinduced electron spin resonance signal exhibits signatures of both the conducting polymer cation and the C(60) anion. Because the photoluminescence in the conducting polymer is quenched by interaction with C(60), the data imply that charge transfer from the excited state occurs on a picosecond time scale. The charge-separated state in composite films is metastable at low temperatures.

4,016 citations


"Polymer solar cells with enhanced o..." refers background in this paper

  • ...With the discovery of efficient photo-induced electron transfer from a conjugated polymer to fulleren...

    [...]

Journal ArticleDOI
TL;DR: In this paper, a polymer solar cell based on a bulk hetereojunction design with an internal quantum efficiency of over 90% across the visible spectrum (425 nm to 575 nm) is reported.
Abstract: A polymer solar-cell based on a bulk hetereojunction design with an internal quantum efficiency of over 90% across the visible spectrum (425 nm to 575 nm) is reported. The device exhibits a power-conversion efficiency of 6% under standard air-mass 1.5 global illumination tests.

4,002 citations

Frequently Asked Questions (12)
Q1. What are the key parameters for achieving high-efficiency solar cells?

The absorption range, the photon–electron conversion rate and the carrier mobilities of the light-harvesting polymers are among the crucial parameters for achieving high-efficiency solar cells. 

Here, the authors show that the open-circuit voltage of polymer solar cells constructed based on the structure of a low-bandgap polymer, PBDTTT4, can be tuned, step by step, using different functional groups, to achieve values as high as 0. 76 V. This increased open-circuit voltage combined with a high short-circuit current density results in a polymer solar cell with a power conversion efficiency as high as 6. 77 %, as certified by the National Renewable Energy Laboratory. In this work, the authors attempt to alter the HOMO of poly [ 4,8-bis-substituted-benzo [ 1,2-b:4,5-b0 ] dithiophene-2,6-diyl-alt-4-substituted-thieno [ 3, 4-b ] thiophene-2,6-diyl ] ( PBDTTT ) -derived polymers4 by adding different electron-withdrawing functional groups, step by step. The authors have shown that the addition of more than one electron-withdrawing group is effective in further lowering the HOMO of PBDTTT. 6 V. the authors chose this polymer system and tried to increase PSC performance by increasing the Voc through molecular design. Polymer solar cells ( PSCs ) have attracted much attention due to their potential in low-cost solar energy harvesting, as well as applications in flexible, light-weight, colourful and large-area devices. Furthermore, fabricating largearea devices without significantly losing efficiency while maintaining long device lifetimes remains challenging23. 

In addition to grafting side chains, substitution of the carbon atom in selected locations also affects the energy levels of a polymer. 

In principle, the strategies used to improve BHJ solar cell efficiency include (i) reducing the bandgap of polymers so as to harvest more sunlight, which leads to higher short-circuit current density (Jsc) and (ii) lowering the highest occupied molecular orbital (HOMO) of the polymers, which increases the opencircuit voltage (Voc). 

The extended absorption of sunlight at longer wavelengths directly reflects on the value of Jsc , and a current density of up to 16 mA cm22 has been achieved15. 

In the last few years, several effective methods have been developed to optimize the interpenetrating network formed by the electron donor and acceptor, including solvent annealing (or slow-growth)10, thermal annealing11–13 and morphology control using mixed solvent mixtures14 or additives15 in the solutions of donor/acceptor blends. 

With the rise in interest in using lowbandgap polymers to harvest more sunlight from longer wavelengths, much effort has been made recently in reducing the bandgap of polymers. 

Based on the PBDTTT polymer derivative system, PSCs with a PCE higher than 7% have been realized by combining the advantages of a low HOMO level in the polymer (high Voc) and long wavelength absorption (high Jsc). 

Although the devices were encapsulated with UV epoxy before shipping to NREL, this nonideal encapsulation process has been observed to cause an efficiency drop of 3% over a 10-day storage period in a nitrogen-filled glovebox (,0.1 ppm O2 and H2O). 

In the PBDTTT– CF device, the highest EQE value was 68.7% at 630 nm, which, to the best of their knowledge, is the highest achieved in a lowbandgap PSCs system. 

the authors show that the open-circuit voltage of polymer solar cells constructed based on the structure of a low-bandgap polymer, PBDTTT4, can be tuned, step by step, using different functional groups, to achieve values as high as 0.76 V. 

As a result, narrowing the bandgap of polymers without sacrificing efficient charge separation as well as high Voc becomes a major hurdle in achieving high-efficiency PSCs.