Sie sind auf Seite 1von 9

www.advenergymat.

de
www.MaterialsViews.com

Ming Cheng, Bo Xu, Cheng Chen, Xichuan Yang, Fuguo Zhang, Qin Tan, Yong Hua,
Lars Kloo, and Licheng Sun*
The development of cost-effective, high-efficiency solar cells
that can meet the ever increasing demand of sustainable
energy is one of the great challenges in the 21st century. Over
the past two years, the emerging family of metal-halide-based
perovskites with the general formula (RNH3)MX3 (R = alkyl,
M = Pb, X = I, Br, or Cl) have been demonstrated to constitute
highly promising photovoltaic materials to fulfill this target.
The perovskite materials possess a unique combination of
intrinsic properties, such as the approproate direct band gap
(1.55 eV), high absorption coefficient, excellent carrier mobility,
very low exciton binding energy, long holeelectron diffusion
lengths, simple (often solution-based) fabrication processes all
together generating high power conversion efficiency (PCE)
of solar cells containing such materials.[125] Typical perovskite
solar cell (PSC) devices consist of several layers starting with a
layer of a nanocrystalline metal oxide, such as a mesoporous
TiO2,[1,15,24] an Al2O3 scaffold,[11] or a ZnO thin film,[12] covered
with the photoactive perovskite (RNH3)MX3 layer, a hole-transport material (HTM) layer, and a metal counter electrode. To
date, extraordinary PCEs of over 19% have been reported for
cells based on a lead-containingperovskite as light-absorbing
material and 2,2,7,7-tetrakis(N,N-di-p-methoxyphenyl-amine)9,9-Spirobifluorene (Spiro-OMeTAD) as the HTM.[24] However, the high cost of Spiro-OMeTAD impedes the growth and
advancement of highly efficient and cost-effective PSCs.[2628]
Moreover, Spiro-OMeTAD suffers from low hole mobility and
conductivity in its pristine form; therefore, p-type dopants are
necessary in order to promote high device performance.[1,12,24]
However, some of the p-type dopants used have been observed
to accelerate the cell degradation due to their deliquescent properties.[29] In addition, the p-type doping strategy increases the
overall cost and also requires a strict optimization of the doping
Dr. M. Cheng, B. Xu, Dr. Y. Hua, Prof. L. Kloo,
Prof. L. Sun
Department of Chemistry
KTH Royal Institute of Technology
SE-10044 Stockholm, Sweden
E-mail: lichengs@kth.se
Dr. C. Chen, Prof. X. Yang, F. Zhang, Q. Tan, Prof. L. Sun
State Key Laboratory of Fine Chemicals
Institute of Artificial Photosynthesis
DUT-KTH Joint Education and
Research Centre on Molecular Devices
Dalian University of Technology (DUT)
116024 Dalian, China

DOI: 10.1002/aenm.201401720

Adv. Energy Mater. 2015, 5, 1401720

conditions, such as the solvent and dopants used, as well as the


doping concentrations.[5] Therefore, it is imperative to develop
low-cost, dopant-free HTMs for efficient PSCs. Most recently,
impressive photovoltaic performance has been reported for
different HTMs, such as small molecule-based HTMs,[2936]
polymer-based HTMs,[3740] and inorganic hole conductors.[41,42]
However, the vast majority of these reported HTMs require
doping and the highest efficiency reported in PSCs so far is less
than 13%. Some novel HTMs, besides good hole mobilities,
simultaneously show strong light absorbtion in the visible and
near-infrared region.[34,35] Thus, these materials also have the
potential to act as photoactive donor materials (PDMs) in bulk
heterojunction (BHJ) organic solar cells (OSCs).[43,44]
Herein, we report the design and synthesis of an acceptor
donoracceptor (ADA) structured small-molecule material
M1 based on a rigid benzo[1,2b:4,5b]-dithiophene (BDT) central building block and phenoxazine (POZ) linkers. With its
unique properties, such as suitable energy levels, strong optical
absorption in the visible region, and high hole mobility, M1 has
been used both as HTM in (CH3NH3)PbI3-perovskite-based
solar cells and as PDM in BHJ OSCs. Excellent PCE of 13.2%
under 100 mW cm2 irradiation was achieved using M1 as
dopant-free HTM in PSCs, as compared to 12.4% obtained for
devices containing the p-type doped Spiro-OMeTAD as HTM
and 7.14% for the devices not containing any HTM. When
applied in BHJ OSCs, a PCE of 6.91% was achieved using an
optimized device structure of ITO/ZnO/polyethylenimine, 80%
ethoxylated (PEIE)/M1:PC70BM (1:1.2)/MoO3/Ag. To the best
of our knowledge, this is the first case where a small-molecule
organic material has been shown to act both as an HTM in
PSCs and as a PDM in BHJ OSCs and showing excellent conversion efficiencies in both types of solar cell systems.
The structure and route of synthesis of the small-molecule
material M1 are shown in Figure 1 and Scheme 1, respectively.
The BDT unit is rigid and flat and is widely used as central
building block in the construction of small-molecule donor
materials for the application in BHJ OSCs, and this class of
materials tend to exhibit high hole mobility and overall conductivity.[4549] POZ is a well conjugated heterocyclic ring system
with a butterfly conformation and it is widely used in dyesensitized solar cells, possessing unique electronic and optical
properties.[5053] The BDT core unit was functionalized with (Z)3-(3-ethylrhodanine-5-alkenyl)-10-(2-ethylhexyl)-phenoxazine
groups in order to generate a low band gap material. The symmetrical structure is expected to enhance stacking interactions, which could be beneficial for promoting a high hole

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(1 of 9) 1401720

COMMUNICATION

Phenoxazine-Based Small Molecule Material for Efficient


Perovskite Solar Cells and Bulk Heterojunction Organic
Solar Cells

www.advenergymat.de

COMMUNICATION

www.MaterialsViews.com

Figure 1. Structure of the small-molecule material M1.

mobility and also to enhance the lifetime of a charge-separated


excited state.[54,55] The 3-ethylrhodanine unit was introduced for
adjusting the highest occupied molecular orbital (HOMO) and
lowest unoccupied molecular orbital (LUMO) energy levels of
M1. The aim of the introduction of the long alkyl chains is to
increase the solubility of M1 in organic solvent. M1 was synthesized from relatively inexpensive commercial materials in
five steps with an overall yield of 52%, which is much higher
than that of Spiro-OMeTAD.[2628] The detailed experiments are
shown in the Supporting Information.

The UVvis absorption spectrum of M1 is shown in


Figure 2 and the corresponding data are summarized in
Table 1. M1 shows a broad and strong optical absorption in the
visible region, both in dichloromethane (CH2Cl2) solution and
as a thin film on a glass substrate. In CH2Cl2, the absorption
peaks are observed at 429 nm (molar extinction coefficients
= 23 600 L mol1 cm1) and 521 nm ( = 37 100 L mol1 cm1).
Cast on a thin film, an obvious redshift of 36 nm (from 521 nm
in solution to 557 nm in the form of a film) is detected. The significant redshift indicates a strong intermolecular stacking
of M1 molecules in the presumably sterically more restricted
film. Again, such effects are expected to promote a higher hole
mobility and thus lead to better photovoltaic performance.[54,55]
Cyclic voltammetry (CV) was performed in order to evaluate
if M1 has suitable energy levels for the use in PSCs and BHJ
OSCs (see Figure S1, Supporting Information). The energy
level diagram of the components of the (CH3NH3)PbI3-based
perovskite solar cells and BHJ OSCs are shown in Figure 3. The
oxidation (Eox) and reduction (Ered) potentials of M1 are 0.79
and 1.05 V versus the normal hydrogen electrode, respectively.
Correspondingly, the HOMO and the LUMO energy levels are
estimated to be 5.29 and 3.45 eV according to the formula
of EHOMO = (Eox + 4.50) (eV) and ELUMO = (Ered + 4.50) (eV),
respectively. The reported HOMO energy level for (CH3NH3)
PbI3 is 5.44 eV,[8] which indicates that M1 is energetically

Scheme 1. Route of synthesis for the small-molecule material M1.

1401720 (2 of 9)

wileyonlinelibrary.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1401720

www.advenergymat.de
www.MaterialsViews.com

Figure 2. UVvis absorption spectra of M1 in CH2Cl2 solution and in the


form of a thin film on a glass substrate.

favorable for hole transfer. In comparison to Spiro-OMeTAD


(HOMO, 5.22 eV),[32] the HOMO energy level of M1 is a bit
more negative. Normally, for PSCs, the open-circuit voltage
(Voc) depends on the difference between the HOMO level of
the HTM and the quasi-Fermi level of the metal-oxide thin film
(here ZnO), and consequently a higher Voc could be expected for
PSCs containing M1 as the HTM, as compared to that obtained
for a device based on Spiro-OMeTAD. In addition, M1 possesses fairly appropriate energy levels for BHJ OSCs. For BHJ
OSCs, the Voc is usually determined by the gap between the
LUMO energy level of the acceptor and the HOMO energy level
of the donor, as well as the work function difference between
the anode and the cathode.[56] The low HOMO energy level of
M1 conveys that a high Voc of BHJ OSCs based on M1 blended
with PC70BM could be expected. The LUMO energy level difference (0.46 eV) between the donor (M1) and acceptor (PC70BM)
is large enough for an efficient separation of the excitons. The
band gap of M1 is 1.84 eV as determined by CV, which is similar to the optical band gap of 1.88 eV obtained from the onset
of the UVvis absorption spectrum in a thin film.
In order to gain some insights into the geometrical configuration and electron distribution of the small-molecule material M1, quantum chemical calculations based on the density
functional theory method were performed by employing the
Gaussian 09 software. From the calculated results (see Figure 4),
we can note that the electron density of HOMO-1 is mainly
delocalized over the central BDT unit, the electron density of
the HOMO is slightly shifted and mainly delocalized over the
POZ linker, and the electron density of the LUMO is mainly

Table 1. Optical and electrochemical data of the small-molecule material M1.

M1

max solutiona)
[nm]

[M1 cm1]

max filmb)
[nm]

E00c)
[eV]

HOMOc)
[eV]

LUMOc)
[eV]

Hole mobility
[cm2 V1 S1]

Conductivity
[S cm1]

429

23 600

557

1.84

5.29

3.45

2.71 104

1.16 103

521

37 100

spectra were recorded in CH2Cl2 solution (2 105 M); b)Films were prepared by spin-coating an o-dichlorobenzene solution (20 mg mL1) of M1 onto glass
slides at a spin speed of 1500 rpm; c)CV measurements were carried out in CH2Cl2 solutions with TBAPF6 (0.1 M) as supporting electrolyte, ferrocene/ferrocenium (Fc/Fc+)
as internal reference, and converted to the vacuum scale according to the formula of EHOMO = (Eox + 4.50) (eV) and ELUMO = (Ered + 4.50) (eV).

a)Absorption

Adv. Energy Mater. 2015, 5, 1401720

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(3 of 9) 1401720

COMMUNICATION

concentrated on the 3-ethylrhodanine groups, demonstrating


that M1 can realize the expected effective charge separation
aimed at in the molecule design. This is expected to facilitate
the separation of strongly bound Frenkel excitons at the donor
acceptor interface.
Also the thermal stability of M1 was investigated through
thermal gravimetric analysis (TGA) (see Figure S2, Supporting
Information). The TGA test results show that M1 is quite stable
and starts degrading only above 350 C.
Hole mobility and conductivity are important parameters
for both HTMs and PDMs. The hole mobility of M1 and
Spiro-OMeTAD were estimated by using the spacechargelimited current method with the device structure of ITO/
MoO3/M1 or Spiro-OMeTAD/MoO3/Ag and the test results
are shown in Figure 5.[57,58] The conductivity of M1 and SpiroOMeTAD were determined by using a two-contact electrical
conductivity setup.[36] The extracted hole mobility of M1 is
2.71 104 cm2 V1 s1, which is considerably higher than that
of Spiro-OMeTAD (1.32 104 cm2 V1 s1). Also, M1 shows
much higher conductivity (1.16 103 S cm1) than that of
lithium-bis(trifluoromethanesulfonyl)imide
(LiTFSI)-doped
Spiro-OMeTAD (1.57 104 S cm1). Therefore, it is a realistic
hypothesis that much better photovoltaic performance should
be obtained using M1 as HTM in PSCs, as will be discussed in
detail later.
First, we investigated the photovoltaic properties of M1 as
HTM in PSCs. The scheme of an optimized device structure
(ITO/ZnO/PC70BM/CH3NH3PbI3/HTMs/Ag) is shown in
Figure 3. The devices were prepared according to reported literature with some modification.[1,12,59] The thin ZnO film was
used as the electron collecting layer. The ZnO nanoparticles
were prepared by the hydrolysis of zinc acetate in methanol. A
relatively compact ZnO layer can be obtained by spin-coating
the prepared ZnO nanoparticles onto an ITO substrate; no calcination or sintering step is required.[12] A PC70BM layer on top
of the ZnO film was used to enhance the electronic selection
efficiency.[59,60] A two-step process was subsequently used
to grow the (CH3NH3)PbI3 crystal on the top of the PC70BM
layer.[1,12] The M1-based HTM layer was then deposited by spincoating an o-dichlorobenzene solution and Ag was finally evaporated as the cathode.
The current densityvoltage (JV) characteristics and the
incident-photon-to-current conversion efficiency (IPCE) spectra
of (CH3NH3)PbI3-based solar cells employing M1 or SpiroOMeTAD as HTMs, together with the devices without any
HTM applied at all, are shown in Figure 6. The related photovoltaic data are collected in Table 2. The best performing device
without any HTM displayed a short-circuit current density (Jsc)

www.advenergymat.de

a)

M1

-4.6

Ag

- 5.1

CH3NH3PbI3

- 4.5
PC70BM

Ag
HTM
Perovskite
PC70BM
ZnO
ITO
Glass

-3.45
-3.91
-3.91 -3

ZnO

Energy versus vacuum (eV)

b)

ITO

-5.29

-5.43
- 5.87

e)

d)

0.5 eV

-5.29

-4.6
MoOx

Ag

PC70BM

M1

ZnO

-5.29 -5.6

-5.1
PEIE

Ag

ZnO

MoOx

-3.45
-3.91
-4.0

ITO

- 5.1

-4.6

M1

- 4.5
PC70BM

Ag
MoO3
Active Layer
PEIE
ZnO
ITO
Glass

-3.45
- 3.91

ITO

Energy versus vacuum (eV)

c)

Energy versus vacuum (eV)

COMMUNICATION

www.MaterialsViews.com

-5.6
-5.87

- 5.87
With PEIE layer

Without PEIE layer

Figure 3. a) Device structure of the (CH3NH3)PbI3-based solar cells. Inset shows the molecular structure of PC70BM. b) Energy level diagram of the
components of the (CH3NH3)PbI3-based solar cells under flat band conditions. c) Device structure of the inverted BHJ OSCs based on M1 blended
with PC70BM. Inset shows the molecular structures of PC70BM and PEIE. d) Energy level diagram of the components of the inverted device under
flat band conditions. e) Illustration of the vacuum level shift and reduced work function of ZnO and ITO electrode after deposition of the PEIE layer.

of 15.2 mA cm2, a Voc of 795 mV, and a fill factor (FF) of 59.1%,
resulting in a PCE of 7.14%. The average PCE of ten identical
devices without HTM is (6.73 0.42)% (Table S3, Supporting
Information). The introduction of dopant-free Spiro-OMeTAD,
doped Spiro-OMeTAD or M1 as HTM substantially increased
the best PCEs to 8.9%, 12.4%, and 13.2%, respectively, under
standard global AM 1.5G illumination. There are two main
reasons for this improvement. One is the HTM layer could
play as an electron blocking layer in the solar cell devices, thus
the charge extraction in these devices are more efficient. Also,
HTM might prevent Ag electrode to directly contact the perovskite layer and penetrate into the layer, resulting in reducing
charge recombination sites in these devices. This result indicates that an efficient HTM indeed is a crucial component in
(CH3NH3)PbI3-based perovskite solar cells. However, the efficiency of device based on dopant-free spiro-OMeTAD is still
very low, just about 8.91%, which is mainly due to its low FF
value. The low FF of PSCs based on dopant-free spiro-OMeTAD
can partly be ascribed to the low conductivity of spiro-OMeTAD
in their pristine form.
The statistical data based on ten identical devices containing
M1 or doped Spiro-OMeTAD as the HTM are collected in
Tables S1 and S2, Supporting Information, giving average PCEs
of (12.6 0.5)% and (12.1 0.3)%, respectively. Comparable
and even a little higher efficiencies can be obtained when M1 is
used as HTM without any doping. Comparing the devices containing M1 and Siro-OMeTAD, the following observations can
be made. Considering the only slightly lower Jsc, observed for
1401720 (4 of 9)

wileyonlinelibrary.com

devices containing M1 the slightly higher PCEs, as compared


to devices containing Spiro-OMeTAD, can mainly be linked to
the higher Voc, an effect expected because of the calculationally
indicated lower HOMO energy level of M1. The improvement
in FF can mainly attributed to much higher conductivity of
M1. It has been widely demonstrated that LiTFSI doping and
4-tert-butylpyridine (TBP) are essential for good performance of
Spiro-OMeTAD-based PSCs, and therefore devices with the two
additives and M1 were prepared. In contrast to Spiro-OMeTADbased PSCs, no obvious enhancement in PCE can be detected
by doping or including the additive into M1 (see Figure S4,
Supporting Information) can be noted. From IPCE spectra we
can deduce that all PSCs display a wide response to the solar
spectrum with a long wavelength limit at around 800 nm. As
compared to the devices without any HTM, the use of M1 or
Spiro-OMeTAD as HTMs gave a remarkable improvement of the
IPCE over the whole region between 350 and 800 nm, most likely
because of a more efficient charge extraction. Most importantly,
the similar shapes of the IPCEs for all devices containing M1
or Spiro-OMeTAD suggest that HTM absorption has negligible
effects on the device performance. The integrated current densities from the IPCE spectra for devices containing M1 or SpiroOMeTAD without HTM were 18.2, 18.4, and 13.9 mA cm2,
respectively, all matching the measured Jscs.
In addition, no obvious hysteresis of photocurrent was
observed by changing the direction in the devices (see
Figure S3, Supporting Information). We think that both HTM
and PC70BM/ZnO play important roles for this phenomenon.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1401720

www.advenergymat.de
www.MaterialsViews.com

COMMUNICATION

Figure 4. Geometrical configuration and electron distribution of M1.

For PSCs, the generated charge (electrons or holes) may persist in the perovskite owing to the slow charge collection in the
thick perovskite layer and require significant time for charge
distribution. This would impede a quick response to the rapid
scanning of the applied voltage for the JV measurements.
Therefore, the slow charge collection via the perovskite material itself must be improved to eliminate the large discrepancy
with scan direction. PC70BM is widely used in BHJ solar cells
as acceptor (A) material and can form large interfacial donor/
acceptor (D/A) contacting area with donor (D) material, which
allows the efficient collection and dissociation of a larger
number of excitons. Also, previous research has shown that the

charge transfer from perovskite to PC70BM is very fast. Thus, in


this type device, the electrical buffer layers (M1 as hole-transport layer and PC70BM/ZnO as electron transport layer) can
efficiently extract free holes and dissociate excitons into free
electrons, resulting in negligible hysteresis of photocurrent.
Subsequently, we also applied M1 to BHJ OSCs; now as
PDM. The scheme of the optimized device structure (ITO/ZnO/
polyethylenimine, 80% ethoxylated (PEIE)/M1:PC70BM/MoO3/
Ag) is shown in Figure 3c. The JV characteristics of inverted
BHJ OSCs with various interfacial layers (ITO only, ZnO only,
PEIE only, and the combination of ZnO and PEIE) under AM
1.5G illumination (100 mW cm2) are shown in Figure 7. The

Figure 5. a) JV plots of the hole-only devices based on M1 and Spiro-OMeTAD and b) JV plots of electron-only devices based on M1 and LiTFSIdoped Spiro-OMeTAD.

Adv. Energy Mater. 2015, 5, 1401720

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(5 of 9) 1401720

www.advenergymat.de

COMMUNICATION

www.MaterialsViews.com

Figure 6. a) JV characteristics and b) IPCE spectra of the perovskite-based solar cells with M1 or Spiro-OMeTAD as HTM together with a cell without
HTM.

related photovoltaic data are collected in Table 3. The thin


ZnO film was used as the electron collecting layer. The PEIE
layer on the top of ZnO was used to enhance the efficiency of
the cell by effectively lowering the work function of ZnO (see
Figure 3d,e).[56] From Figure 7 and Table 3, we can observe
that the presence of an electron-collecting layer is crucial in
order to obtain a high cell efficiency. For the ITO-only devices
(ITO/M1:PC70BM (1:1.2)/MoO3/Ag), the Jsc, Voc, and FF are all
very low due to high recombination losses at the ITO surface.[12]
By incorporation of even a very thin PEIE layer (10 nm) or ZnO
layer into the device the photovoltaic parameters Jsc, Voc, and
FF were significantly improved and correspondingly much
higher efficiencies were achieved. When PEIE alone was used
as an interfacial layer, the Voc, Jsc, and FF of the devices were
improved to 878 mV, 9.46 mA cm2, and 44.8%, respectively.
However, PEIE is a wide-band gap material and serves only as
a surface modifier limiting the recombination losses and not
as an efficient electron-extraction layer or hole-blocking layer,
thus, the Jsc and FF are not significantly affected. The thickness
of the ZnO layer also has a big effect on the performance of
the OSCs. The highest PCE of 6.25% was achieved for devices
containing ZnO only; a 30 nm ZnO film as electron-collecting
layer. A further increase of the thickness of the ZnO film will
result in a decrease in device performance. The combination of
a ZnO and a PEIE layer improves both the Jsc and Voc, leading
to a higher PCE of 6.91%. Considering that the same donor
material and acceptor material was used in these devices, the
increase in Voc can be attributed to the reduced work function

of the cathode after the PEIE deposition onto the ZnO film surface (4.0 eV for PEIE-coated ZnO vs 4.5 eV for ZnO only).[56]
The IPCE of the best-performing devices processed with and
without the PEIE layer were investigated. As shown in Figure 7,
both the devices showed a very wide response to the solar
spectrum in the region of 350750 nm and the IPCE values
increased in the region 450750 nm when the PEIE layer was
used; correspondingly, a higher Jsc was obtained. The integrated
current densities calculated from the IPCE spectra are 11.3 and
12.3 mA cm2, respectively, all matching the measured Jscs considering the experimental error.
Going back to the PSC devices, Electrochemical Impedance
Spectroscopy (EIS) was performed under one sun illumination
to the devices containing M1 or Spiro-OMeTAD as HTMs in
order to gain insight into the charge/hole transport in these
devices. The results were fitted by following the equivalentcircuit model reported in previous literature and shown in
Figure 8.[61,62] The main arc located at lower frequencies is identified with the recombination loss process and the arc at higher
frequencies of the impedance spectrum is typically attributed
to the hole transport of and hole extraction from the HTM.
From the results we can note that the devices containing M1
present slightly higher recombination resistances than those of
Spiro-OMeTAD-based devices. Additionally, the HOMO energy
level of M1 is 0.07 eV lower than that of Spiro-OMeTAD. Both
these factors the higher Voc observed for the devices containing
M1. At the same time, the devices containing M1 display lower
HTM resistance than those of Spiro-OMeTAD based ones, but

Table 2. Photovoltaic parameters of (CH3NH3)PbI3-based solar cells with and without HTM.
HTM

Voc
[mV]

Jsc
[mA cm2]

FF
[%]

PCE
[%]

Integrated Jsc
[mA cm2]

M1a)

1016

19.1

67.8

13.2

18.2

M1b)

1030

18.8

68.0

13.1

18.2

a)

970

18.3

50.2

8.9

Spiro-OMeTADb)

981

19.4

65.2

12.4

18.4

No HTM

795

15.2

59.1

7.14

13.9

Spiro-OMeTAD

a)Without

any additives; b)Addition of TBP and LiTFSI.

1401720 (6 of 9)

wileyonlinelibrary.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1401720

www.advenergymat.de
www.MaterialsViews.com

COMMUNICATION

Figure 7. a) JV characteristics and b) IPCE spectra of BHJ OSCs employing M1 as donor material.

the differences are not very obvious in high voltage region,


which may explain the similar FF recorded for both of the
devices containing M1 or Spiro-OMeTAD as HTM.
In summary, we have designed and synthesized an ADA
structured small-molecule material M1, containing an electronrich BDT unit as core building block, flanked by POZ units,
and end-capped with electron-withdrawing 3-ethylrodanine
units. The unique characteristics of M1, such as suitable energy
levels, strong optical absorption in the visible region and high
hole mobility, prompted us to use it both as HTM in (CH3NH3)PbI3-based perovskite solar cells and as PDM in BHJ OSCs.
For PSCs, in the presence of M1 as HTM, a PCE of 13.2% was

achieved, which is comparable and even a little higher than that


of devices based on the well-known HTM Spiro-OMeTAD. In
contrast to the state-of-the-art HTM Spiro-MeOTAD, the M1
material can be used without any p-type doping or other additives. Applied in BHJ OSCs, through optimization, a PCE of
6.9% can be obtained using a device structure of ITO/ZnO/
PEIE/M1:PC70BM (1:1.2)/MoO3/Ag. Moreover, this novel and
high-performing material is synthesized from relatively inexpensive and commercially available reactants and in only five
synthetic steps with a high overall yield (52%). The present
results offer a new design strategy towards the development of
highly efficient solid state photovoltaic devices.

Table 3. Device performance parameters for BHJ OSCs based on M1:PC70BM (1:1.2).
Thickness of ZnO film
[nm]

Thickness of PEIE layer


[nm]

Voc
[mV]

Jsc
[mA cm2]

FF
[%]

PCE
[%]

460 6.5

5.74 0.11

33.1 0.3

0.87 0.02
3.72 0.04

10

878 6.3

9.46 0.04

44.8 0.3

10

805 4.28

10.9 0.09

52.5 0.4

4.60 0.19

25

850 5.1

11.4 0.3

59.8 2.0

5.97 0.22

30

855 3.8

11.9 0.3

61.5 2.1

6.25 0.22

30

10

897 6.6

12.5 0.4

61.6 1.4

6.91 0.13

50

835 3.2

11.1 0.3

56.2 1.9

5.21 0.18

Adv. Energy Mater. 2015, 5, 1401720

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(7 of 9) 1401720

www.advenergymat.de

COMMUNICATION

www.MaterialsViews.com

Figure 8. a) Recombination resistance and b) transport resistance of HTMs in the PSC devices containing M1 or Spiro-OMeTAD as HTMs under one
sun illumination obtained from a fitted electrochemical model to EIS data.

Experimental Section
PSC Fabrication: The devices were fabricated based on an optimized
structure of ITO/ZnO/PC70BM/CH3NH3PbI3/HTM/Ag using a
solution-based process. The ITO-coated glass substrates were cleaned
by ultrasonic treatment in detergent, deionized water, acetone, and
isopropyl alcohol under ultrasonication for 15 min each, followed
by a plasma treatment for 15 min. After treatment, a thin layer of
ZnO was spin-coated onto the ITO surface at 3000 rpm for 30 s. This
procedure repeated three times in order to obtain a continuous and
smooth film with a thickness of about 25 nm. PC70BM was dissolved in
o-dichlorobenzene at a concentration of 15 mg mL1. The solution was
spin cast on the ZnO film at a speed of 1500 rpm for 1 min and then
annealed at 120 C for 10 min in ambient air. The thickness of the PC70BM
is about 10 nm. A PbI2 solution (dissolved in N,N-dimethylformamide at
a concentration of 460 mg mL1) was then spin coated on top of the
PC70BM layer at 3000 rpm for 15 s. After drying several minutes in air,
the substrate was dipped into a solution of (CH3NH3)I in 2-propanol
(10 mg mL1) for 1 h, then dried under a flow of dry N2. The thickness
of the perovskite capping layer is about 350 nm. Subsequently, the HTM
layer (about 40 nm) was produced by spin-coating the M1 (dissolved in
o-dichlorobenzene at a concentration of 30 mg mL1) solution on top
of the perovskite layer with a spin speed of 2000 rpm. For the SpiroOMeTAD-based devices, the Spiro-OMeTAD solution (80 mg SpiroOMeTAD, 30 L TBP and 20 L Li-TFSI solution (520 mg Li-TFSI in 1 mL
acetonitrile), all dissolved in 1 mL chlorobenzene) was deposited by spin
coating at 4000 rpm for 30 s. Finally, a layer of 200 nm Ag was deposited
sequentially under high vacuum (<4 104 Pa) by thermal evaporation
through a shadow mask to form an active area of 16 mm2.
BHJ Solar Sell Fabrication: The devices were fabricated with an
optimized inverted structure of ITO/ZnO/PEIE/M1:PC70BM/MoO3/Ag
using a solution-based process. The ITO-coated glass substrates were
cleaned by ultrasonic treatment in detergent, deionized water, acetone,
and isopropyl alcohol under ultrasonication for 15 min each, followed by
a plasma treatment for 15 min. After treatment, a thin layer of ZnO was
spin-coated onto the ITO surface at 3000 rpm for 30 s. The procedure
can be repeated a different number of times in order to generate a
continuous and smooth film with different thicknesses (this ZnO
coating procedure was skipped for ITO-only and PEIE-only devices).
PEIE (Mw = 70 000 g/mol) was dissolved in H2O at a concentration of
3540 wt%. It was further diluted using 2-methoxyethanol to a weight
concentration of 0.4%. The solution was spin cast onto the ZnO film
(for combined ZnO and PEIE devices) or onto the ITO (for PEIE-only
devices) at a speed of 5000 rpm for 1 min. The approximate thickness
of the film obtained was 10 nm. Spin-coated PEIE films were annealed
at 120 C for 10 min in ambient air (this PEIE coating procedure
was skipped for ZnO-only devices). The active layers (120 nm)

1401720 (8 of 9)

wileyonlinelibrary.com

were spin-cast from a solution of M1:PC70BM (1:1.2 weight ratio) in


o-dichlorobenzene with 0.1 v/v% DIO, with an overall concentration of
33 mg mL1. The prepared solutions were stirred at 90 C before film
casting for 2 h. The coated films were annealed at 90 C for 15 min in
order to obtain the optimum result. Finally, 5 nm MoO3 and 100 nm Ag
layers were deposited sequentially under high vacuum (<4 104 Pa) by
thermal evaporation through a shadow mask to form an active area of
16 mm2.

Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.

Acknowledgements
This work was financially supported by the Swedish Energy Agency, the
Knut and Alice Wallenberg Foundation, the National Natural Science
Foundation of China (21120102036, 91233201), and the National Basic
Research Program of China (973 program, 2014CB239402).
Received: September 27, 2014
Revised: November 8, 2014
Published online: January 7, 2015

[1] J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao,


M. K. Nazeeruddin, M. Grtzel, Nature (London) 2013, 499, 316.
[2] F. De Angelis, Acc. Chem. Res. 2014, 47, 3349.
[3] P. Docampo, J. M. Ball, M. Darwich, G. E. Eperon, H. J. Snaith, Nat.
Commun. 2013, 4, 2761.
[4] G. Hodes, D. Cahen, Nat. Photonics 2014, 8, 87.
[5] N. J. Jeon, H. G. Lee, Y. C. Kim, J. Seo, J. H. Noh, J. Lee, S. I. Seok,
J. Am. Chem. Soc. 2014, 136, 7837.
[6] N. J. Jeon, J. H. Noh, Y. C. Kim, W. S. Yang, S. Ryu, S. I. Seok, Nat.
Mater. 2014, 13, 897.
[7] S. Kazim, M. K. Nazeeruddin, M. Grtzel, S. Ahmad, Angew. Chem.
Int. Ed. 2014, 53, 2812.
[8] H. S. Kim, C. R. Lee, J. H. Im, K. B. Lee, T. Moehl, A. Marchioro,
S. J. Moon, R. Humphry-Baker, J. H. Yum, J. E. Moser, M. Grtzel,
N. G. Park, Sci. Rep. 2012, 2, 591.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1401720

www.advenergymat.de
www.MaterialsViews.com

Adv. Energy Mater. 2015, 5, 1401720

[37] Y. Guo, C. Liu, K. Inoue, K. Harano, H. Tanaka, E. Nakamura,


J. Mater. Chem. A 2014, 2, 13827.
[38] J. H. Heo, S. H. Im, J. H. Noh, T. N. Mandal, C.-S. Lim, J. A. Chang,
Y. H. Lee, H.-j. Kim, A. Sarkar, K. NazeeruddinMd, M. Grtzel,
S. I. Seok, Nat. Photonics 2013, 7, 486.
[39] Y. S. Kwon, J. Lim, H.-J. Yun, Y.-H. Kim, T. Park, Energy Environ. Sci.
2014, 7, 1454.
[40] S. Ryu, J. H. Noh, N. J. Jeon, Y. Chan Kim, W. S. Yang, J. Seo,
S. I. Seok, Energy Environ. Sci. 2014, 7, 2614.
[41] S. Chavhan, O. Miguel, H.-J. Grande, V. Gonzalez-Pedro,
R. S. Sanchez, E. M. Barea, I. Mora-Sero, R. Tena-Zaera, J. Mater.
Chem. A 2014, 2, 12754.
[42] P. Qin, S. Tanaka, S. Ito, N. Tetreault, K. Manabe, H. Nishino,
M. K. Nazeeruddin, M. Grtzel, Nat. Commun. 2014, 5,
3834.
[43] Y. Chen, X. Wan, G. Long, Acc. Chem. Res. 2013, 46, 2645.
[44] Y. Li, Q. Guo, Z. Li, J. Pei, W. Tian, Energy Environ. Sci. 2010, 3,
1427.
[45] Y. Liu, C. C. Chen, Z. Hong, J. Gao, Y. M. Yang, H. Zhou, L. Dou,
G. Li, Y. Yang, Sci. Rep. 2013, 3, 3356.
[46] Y. Liu, X. Wan, F. Wang, J. Zhou, G. Long, J. Tian, Y. Chen, Adv.
Mater. 2011, 23, 5387.
[47] S. Shen, P. Jiang, C. He, J. Zhang, P. Shen, Y. Zhang, Y. Yi, Z. Zhang,
Z. Li, Y. Li, Chem. Mater. 2013, 25, 2274.
[48] J. Zhou, X. Wan, Y. Liu, Y. Zuo, Z. Li, G. He, G. Long, W. Ni, C. Li,
X. Su, Y. Chen, J. Am. Chem. Soc. 2012, 134, 16345.
[49] J. Zhou, Y. Zuo, X. Wan, G. Long, Q. Zhang, W. Ni, Y. Liu, Z. Li,
G. He, C. Li, B. Kan, M. Li, Y. Chen, J. Am. Chem. Soc. 2013, 135,
8484.
[50] M. Cheng, X. Yang, C. Chen, Q. Tan, L. Sun, J. Mater. Chem. A 2014,
2, 10465.
[51] K. M. Karlsson, X. Jiang, S. K. Eriksson, E. Gabrielsson, H. Rensmo,
A. Hagfeldt, L. Sun, Chem. Eur. J. 2011, 17, 6415.
[52] H. Tian, I. Bora, X. Jiang, E. Gabrielsson, K. M. Karlsson,
A. Hagfeldt, L. Sun, J. Mater. Chem. 2011, 21, 12462.
[53] H. Tian, X. Yang, R. Chen, A. Hagfeldt, L. Sun, Energy Environ. Sci.
2009, 2, 674.
[54] Y. Sun, G. C. Welch, W. L. Leong, C. J. Takacs, G. C. Bazan,
A. J. Heeger, Nat. Mater. 2012, 11, 44.
[55] W. Wu, Y. Liu, D. Zhu, Chem. Soc. Rev. 2010, 39, 1489.
[56] A. K. Kyaw, D. H. Wang, V. Gupta, J. Zhang, S. Chand, G. C. Bazan,
A. J. Heeger, Adv. Mater. 2013, 25, 2397.
[57] M. A. Khan, W. Xu, H. Khizar ul, Y. Bai, X. Y. Jiang, Z. L. Zhang,
W. Q. Zhu, J. Appl. Phys. 2008, 103, 014509.
[58] T. Yasuda, Y. Yamaguchi, D.-C. Zou, T. Tsutsui, Jpn. J. Appl. Phys.
2002, 41, 5626.
[59] Q. Wang, Y. Shao, Q. Dong, Z. Xiao, Y. Yuan, J. Huang, Energy
Environ. Sci. 2014, 7, 2359.
[60] J. Seo, S. Park, Y. Chan Kim, N. J. Jeon, J. H. Noh, S. C. Yoon,
S. I. Seok, Energy Environ. Sci. 2014, 7, 2642.
[61] V. Gonzalez-Pedro, E. J. Juarez-Perez, W. S. Arsyad, E. M. Barea,
F. Fabregat-Santiago, I. Mora-Sero, J. Bisquert, Nano Lett. 2014, 14,
888.
[62] H. S. Kim, J. W. Lee, N. Yantara, P. P. Boix, S. A. Kulkarni,
S. Mhaisalkar, M. Grtzel, N. G. Park, Nano Lett. 2013, 13,
2412.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(9 of 9) 1401720

COMMUNICATION

[9] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem. Soc.


2009, 131, 6050.
[10] J. W. Lee, D. J. Seol, A. N. Cho, N. G. Park, Adv. Mater. 2014, 26,
4991.
[11] M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith,
Science 2012, 338, 643.
[12] D. Liu, T. L. Kelly, Nat. Photonics 2014, 8, 133.
[13] M. Liu, M. B. Johnston, H. J. Snaith, Nature (London) 2013, 501,
395.
[14] O. Malinkiewicz, A. Yella, Y. H. Lee, G. M. Espallargas, M. Grtzel,
M. K. Nazeeruddin, H. J. Bolink, Nat. Photonics 2014, 8, 128.
[15] A. Mei, X. Li, L. Liu, Z. Ku, T. Liu, Y. Rong, M. Xu, M. Hu, J. Chen,
Y. Yang, M. Gratzel, H. Han, Science 2014, 345, 295.
[16] J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal, S. I. Seok, Nano Lett.
2013, 13, 1764.
[17] S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. Alcocer,
T. Leijtens, L. M. Herz, A. Petrozza, H. J. Snaith, Science 2013, 342,
341.
[18] C. Wehrenfennig, G. E. Eperon, M. B. Johnston, H. J. Snaith,
L. M. Herz, Adv. Mater. 2014, 26, 1584.
[19] C. Wehrenfennig, M. Liu, H. J. Snaith, M. B. Johnston, L. M. Herz,
Energy Environ. Sci. 2014, 7, 2269.
[20] M. Xiao, F. Huang, W. Huang, Y. Dkhissi, Y. Zhu, J. Etheridge,
A. Gray-Weale, U. Bach, Y. B. Cheng, L. Spiccia, Angew. Chem. Int.
Ed. 2014, 53, 9898.
[21] Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang, Y. Gao,
J. Huang, Energy Environ. Sci. 2014, 7, 2619.
[22] G. Xing, N. Mathews, S. Sun, S. S. Lim, Y. M. Lam, M. Grtzel,
S. Mhaisalkar, T. C. Sum, Science 2013, 342, 344.
[23] W. J. Yin, T. Shi, Y. Yan, Adv. Mater. 2014, 26, 4653.
[24] H. Zhou, Q. Chen, G. Li, S. Luo, T. B. Song, H. S. Duan, Z. Hong,
J. You, Y. Liu, Y. Yang, Science 2014, 345, 542.
[25] Z. Zhu, Y. Bai, T. Zhang, Z. Liu, X. Long, Z. Wei, Z. Wang, L. Zhang,
J. Wang, F. Yan, S. Yang, Angew. Chem. Int. Ed. 2014, 53, 12571.
[26] U. Bach, K. De Cloedt, H. Spreitzer, M. Grtzel, Adv. Mater. 2000,
12, 1060.
[27] T. P. I. Saragi, T. Spehr, A. Siebert, T. Fuhrmann-Lieker, J. Salbeck,
Chem. Rev. 2007, 107, 1011.
[28] B. Xu, H. Tian, D. Bi, E. Gabrielsson, E. M. J. Johansson,
G. Boschloo, A. Hagfeldt, L. Sun, J. Mater. Chem. A 2013, 1, 14467.
[29] J. Liu, W. Yongzhen, C. Qin, X. Yang, T. Yasuda, A. Islam, K. Zhang,
W. Peng, L. Han, W. Chen, Energy Environ. Sci. 2014, 7, 2963.
[30] H. Choi, S. Paek, N. Lim, Y. Lee, M. K. Nazeeruddin, J. Ko, Chem.
Eur. J. 2014, 20, 10894.
[31] K. Do, H. Choi, K. Lim, H. Jo, J. W. Cho, M. K. Nazeeruddin, J. Ko,
Chem. Commun. 2014, 50, 10971.
[32] A. Krishna, D. Sabba, H. Li, J. Yin, P. P. Boix, C. Soci,
S. G. Mhaisalkar, A. C. Grimsdale, Chem. Sci. 2014, 5, 2702.
[33] H. Li, K. Fu, A. Hagfeldt, M. Grtzel, S. G. Mhaisalkar,
A. C. Grimsdale, Angew. Chem. Int. Ed. 2014, 53, 4085.
[34] P. Qin, H. Kast, M. K. Nazeeruddin, S. M. Zakeeruddin, A. Mishra,
P. Baeuerle, M. Grtzel, Energy Environ. Sci. 2014, 7, 2981.
[35] P. Qin, S. Paek, M. I. Dar, N. Pellet, J. Ko, M. Grtzel,
M. K. Nazeeruddin, J. Am. Chem. Soc. 2014, 136, 8516.
[36] B. Xu, E. Sheibani, P. Liu, J. Zhang, H. Tian, N. Vlachopoulos,
G. Boschloo, L. Kloo, A. Hagfeldt, L. Sun, Adv. Mater. 2014, 26, 6629.

Das könnte Ihnen auch gefallen