Vous êtes sur la page 1sur 7

www.advenergymat.

de
www.MaterialsViews.com

Chang Liu, Kai Wang, Pengcheng Du, Chao Yi, Tianyu Meng, and Xiong Gong*
than that of the holes in CH3NH3PbCl3xIx
(Leff, e/Leff, h+ < 1), measured by the
electron-beam induced current (EBIC)
method,[23] a mesoporous electron transporting scaffold is indispensable to effectively extract and transport the electrons.
Consequently, high PCEs over 15% from
meso-superstructured pero-HSCs (MS
pero-HSCs) have been achieved by applying
mesoporous TiO2 and Al2O3 as the electron
extraction layer (EEL) and insulating scaffold, respectively.[2428] In addition, one
outstanding advantage of MS pero-HSCs
over that of PHJ pero-HSCs is the higher
consistence of efficiency under different
scan rates and scan directions, showing
less hysteresis and transient behavior in
currentvoltage measurement,[29] which
is critically important for high throughput
large-scale manufacturing pero-HSCs
products. Nevertheless, the high temperature sintering process for mesoporous metal oxides (TiO2/Al2O3)
and the homogeneous infiltration of perovskite materials into
the mesoporous electron extraction material[2428] are especially
adverse for large-scale commercialization of pero-HSCs.
To address these problems, we fabricate conventional bulk heterojunction (BHJ) pero-HSCs by mixing
PC61BM ([6,6]-phenyl-C61-butyric acid methyl ester) with
CH3NH3PbI3, which aims to circumvent issues existed in the
PHJ and MS pero-HSCs, while inheriting advantages from
both constructions. In the conventional BHJ pero-HSCs,
PC61BM has twofold contributions. First, PC61BM can effectively extract the electrons from the PC61BM/CH3NH3PbI3
interfaces,[30,31] which is originated from the high electron
affinity (4.3 eV) of PC61BM[32] and a 0.4 eV energy level
offset between the lowest unoccupied molecular orbit (LUMO)
energy level of CH3NH3PbI3 (3.9 eV) and that of PC61BM
(4.3 eV). In this way, PC61BM acts as the electron acceptor
in the BHJ composite of CH3NH3PbI3:PC61BM, which would
address the issues induced by the shorter diffusion length of
the electrons than that of holes in CH3NH3PbI3.[23] Second,
a retarded reaction between CH3NH3I and PbI2 is expected
with the addition of PC61BM. The interactions between the
ester group of PC61BM and CH3NH3I could form an intermediate phase, which is expected to induce more uniform and
dense perovskite layers (detailed discussion in later context).
The BHJ pero-HSCs show a tremendously enhanced PCE of
12.78%, with an open circuit voltage (VOC) of 0.90 V, a short
circuit current density (JSC) of 26.86 mA cm2, and a fill factor
(FF) of 52.9%. Moreover, through the synergistic functions

Efficient conventional bulk heterojunction (BHJ) perovskite hybrid solar cells


(pero-HSCs) solution-processed from a composite of CH3NH3PbI3 mixed
with PC61BM ([6,6]-phenyl-C61-butyric acid methyl ester), where CH3NH3PbI3
acts as the electron donor and PC61BM acts as the electron acceptor, are
reported for the first time. The efficiency of 12.78% is twofold enhancement in comparison with the conventional planar heterojunction pero-HSCs
(6.90%) fabricated by pristine CH3NH3PbI3. The BHJ pero-HSCs are further
optimized by using PC61BM/TiO2 bi-electron-extraction-layer (EEL), which
are both solution-processed and then followed with low-temperature thermal
annealing. Due to higher electrical conductivity of PC61BM over that of TiO2,
an efficiency of 14.98%, the highest reported efficiency for the pero-HSCs
without incorporating high-temperature-processed mesoporous TiO2 and
Al2O3 as the EEL and insulating scaffold, is observed from PC61BM modified
BHJ pero-HSCs. Thus, the findings provide a simple way to approach high
efficiency low-cost pero-HSCs.

1. Introduction
The past few years have witnessed the mushroom development of organicinorganic hybrid perovskite materials and
their applications in photovoltaics.[13] The most attractive one
is the class of organometal halide perovskite with a structure
of MAX3 (M = CH3NH3, A = Pb, X = Cl, Br, or I), which
forms the MX6 octahedra where M is located at the center of
the octahedra and X lies in the corner around M.[4] The organometal halide perovskite materials are direct bandgap semiconductors, which endow the material with a high absorption
extinction coefficient ranging from visible to near infrared
region.[57] Moreover, ambipolar transport characteristics of the
perovskite materials enable both hole and electron to be transported simultaneously in perovskite hybrid solar cells (peroHSCs). In addition, the long charge carrier diffusion lengths
of perovskite materials (1 m in CH3NH3PbI3xClx, 100 nm
in CH3NH3PbI3)[8] encourage the progressive development of
planar heterojunction (PHJ) pero-HSCs, with power conversion efficiencies (PCEs) ranging from 3% to over 15%.[922]
However, owing to the shorter diffusion length of the electrons

C. Liu, K. Wang, P. Du, C. Yi, T. Meng, Prof. X. Gong


College of Polymer Science and Polymer Engineering
University of Akron
Akron, OH 44325, USA
E-mail: xgong@uakron.edu

DOI: 10.1002/aenm.201402024

Adv. Energy Mater. 2015, 1402024

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

wileyonlinelibrary.com

(1 of 7) 1402024

FULL PAPER

Efficient Solution-Processed Bulk Heterojunction Perovskite


Hybrid Solar Cells

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com

Scheme 1. Pero-HSCs device structures. a) The conventional planar heterojunction pero-HSCs. b) The conventional bulk heterojunction (BHJ) peroHSCs. c) PC61BM modified conventional BHJ pero-HSCs.

1402024 (2 of 7)

wileyonlinelibrary.com

FF (%)

60
55
50

(mA/cm )

45

SC

The pero-HSCs with the device structures of ITO/TiO2/


CH3NH3PbI3/P3HT/MoO3/Ag (represented as conventional
PHJ pero-HSCs), ITO/TiO2/CH3NH3PbI3:PC61BM/P3HT/
MoO3/Ag (represented as conventional BHJ pero-HSCs),
and ITO/TiO2/PC61BM/CH3NH3PbI3:PC61BM/P3HT/MoO3/
Ag (represented as PC61BM modified conventional BHJ peroHSCs), where ITO is indium tin oxide, P3HT is poly(3-hexylthiophene-2,5-diyl), MoO3 is molybdenum trioxide, and Ag is
silver, respectively, are fabricated and characterized. The device
architectures are shown in Scheme 1. Figure 1 displays the
device performance of the conventional BHJ pero-HSCs influenced by different weight ratios of PC61BM in the PbI2 solution. It was found that both JSC and FF, consequently the PCEs
first increase with the increasing weight ratio of PC61BM:PbI2
from 0 to 10:400, then drop with further enhancement of
PC61BM percentage, and finally remain to a constant value.
VOC stays invariant (0.90 V) for all conventional BHJ peroHSCs with different weight ratios of PC61BM:PbI2. The first
rising stage for both JSC and FF would be primarily attributed
to the increased donor/acceptor interfaces for efficient charge
carrier extraction and transportation from BHJ composite.
However, when the PC61BM percentage is further increased,
the inferior electrical conductivity of PC61BM (107 S cm1)
than that of CH3NH3PbI3 (5 103 S cm1) brings about the
descending stage,[34] resulting in dramatically reduced JSC.
Finally, the almost invariant PCEs with further increasing concentrations of PC61BM indicate the contribution of PC61BM
is insufficient due to the saturated solubility of PC61BM in
the CH3NH3PbI3:PC61BM solution. Thus, balancing the
donor/acceptor interfaces and high electrical conductivity of

28
26
24
22

2. Results and Discussion

CH3NH3PbI3 are critical for approaching high PCEs from the


conventional BHJ pero-HSCs.
JV characteristics of pero-HSCs are carried out to study
the role of PC61BM as the electron acceptor in the conventional BHJ pero-HSCs. We note that during the measurement
of all pero-HSCs, a light-soaking effect on enhancing PCEs
was observed.[3538] As compared with the JSC from the peroHSCs without any illumination, we found that the JSC was
enhanced over 30% from the pero-HSCs with a conventional
device structure of ITO/TiO2/CH3NH3PbI3/P3HT/MoO3/
Al (Scheme 1a) after the pero-HSCs was illuminated about
510 min; under the same illumination condition, however, the
JSC kept a constant for the pero-HSCs with a inverted device
structure
of
ITO/PEDOT:PSS/CH3NH3PbI3/PCBM/Al.[39]
Thus, we conduct the JV characteristics of pero-HSCs after the
pero-HSCs are illuminated about 10 min.[3538]

16

PCE (%)

of PC61BM as described above, the BHJ pero-HSCs exhibit


advantages of MS pero-HSCs, demonstrating marginal hysteresis and transient behaviors (this part will be discussed
later), which is critically important for large-scale commercialization. In contrast, the PHJ counterparts not only show
much inferior PCE of 6.90% with a VOC of 0.90 V, a JSC of
20.10 mA cm2, and a FF of 37.0%, but also exhibit strong
dependence of efficiency on different scan directions. The
device performance of BHJ pero-HSCs is further optimized
by incorporating PC61BM as the EEL.[33] PCE of 14.95%, with
slightly raised VOC of 0.95 V, JSC of 27.60 mA cm2, and FF of
57.1% are observed.

14
12
10
8

4
8
12
16
20
24
28
Weight Ratio of PC BM in PbI Solution (/400 %)
61

Figure 1. The influence of PC61BM on the device performance (JSC, FF,


and PCEs) of the conventional BHJ pero-HSCs with the structure of ITO/
TiO2/CH3NH3PbI3:PC61BM/P3HT/MoO3/Ag.

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

Adv. Energy Mater. 2015, 1402024

www.advenergymat.de
www.MaterialsViews.com

Current Density (mA/cm )

-4

Conventional PHJ pero-HSCs


conventional BHJ pero-HSCs
PC BM modified conventional BHJ pero-HSCs

-8

61

-12
-16
-20
-24
-28
-32
0

0.5

Voltage (V)
Figure 2. Currentvoltage characteristics of the conventional planar heterojunction pero-HSCs, the conventional bulk heterojunction (BHJ) peroHSCs, and PC61BM modified conventional BHJ pero-HSCs.

As shown in Figure 2, the conventional PHJ pero-HSCs


show a VOC of 0.90 V, a JSC of 20.0 mA cm2, a FF of 37%, and
a corresponding PCE of 6.90%. The conventional BHJ peroHSCs exhibit a VOC of 0.90 V, a JSC of 26.86 mA cm2, a FF
of 52.9%, and a corresponding PCE of 12.78%. This is one
of the highest PCEs for pero-HSCs without incorporating
mesoporous metal oxide as the EEL or insulating scaffold so
far. The dramatically enhanced JSC and FF confirm the role of
PC61BM as the electron acceptor in the conventional BHJ
pero-HSCs fabricated by CH3NH3PbI3:PC61BM active layer,
which facilitates the exciton dissociation at the manifold interfaces. Moreover, by incorporating with highly electrical conductive PC61BM as the EEL, PC61BM modified conventional BHJ
pero-HSCs exhibit boosted PCE of 14.95%, along with a slightly
enhanced VOC (0.95 V), JSC (27.60 mA cm2), and a greatly
increased FF (57.1%). Originated from much higher electrical
conductivity of PC61BM than that of TiO2 (1011 S cm1), the
0

a
2 V/s

-5

b
Negative to positive scan

Current Density (mA/cm2)

Current Density (mA/cm2)

Negative to positive scan


Positive to negative
2 V/s

-10

-15

-20

-25

-30

2 V/s

-5

Positive to negative scan


10 mV/s

-10

2 V/s

-15

-20

-25

-30
0

0.2

0.4

0.6

0.8

Voltage (V)

0.2

0.4

0.6

0.8

Voltage (V)

Figure 3. Currentvoltage characteristics of a) the conventional planar heterojunction pero-HSCs and b) the conventional bulk heterojunction (BHJ)
pero-HSCs under different scan directions and scan rates.

Adv. Energy Mater. 2015, 1402024

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

wileyonlinelibrary.com

(3 of 7) 1402024

FULL PAPER

electron extraction from CH3NH3PbI3 perovskite layer to the


cathode electrode would be much more effective, generating
the abundantly enhanced FF and marginally increased JSC.
In addition, on account of the extremely low hole mobility of
PC61BM, holes back transfer would also be suppressed from
CH3NH3PbI3 layer to the cathode electrode, which brings about
restrained leakage current and thus results in the slightly raised
VOC. As a result, boosted PECs are observed from the PC61BM
modified conventional BHJ pero-HSCs.
It is noteworthy that the relatively low FF (57%) is aroused
by low temperature annealing treatment for the compact
TiOx layer.[40] Disparate from other reports,[24,27,28] where the
compact TiOx layer was treated with high temperature sintering process, low temperature (90 C) treatment reported in
this study for the TiOx EEL would be more compatible with
large-scale commercialization of pero-HSCs in any flexible
substrates. However, the defects in the TiOx layer would be
inevitably generated under such low-temperature treatment,
bringing about relatively large leakage current, resulting in
low FF.[41] Nevertheless, the ultrahigh JSC observed from the
BHJ pero-HSCs compensate the low FF, and consequently preserve a high PCE that is comparable with those from the MS
pero-HSCs. Such high JSC would be the results of the largely
suppressed recombination losses in the CH3NH3PbI3:PC61BM
active layer upon the incorporation of PC61BM as the electron
acceptor, which significantly contribute to the charge carrier
extraction efficiency.[42]
To study the hysteresis and transient behaviors of pero-HSCs,
JV measurements under different scan rates and scan directions were conducted for both conventional PHJ and BHJ peroHSCs. As shown in Figure 3a, the JV characteristics of the conventional PHJ pero-HSCs are intensively dependent on different
scan directions, which are consistent with other reports.[4345]
However, as shown in Figure 3b, the JV characteristics of the
conventional BHJ pero-HSCs under different scan directions
and scan rates are sharply contrasted to those from the conventional PHJ pero-HSCs. The highly consistent JV characteristics insinuate that marginal hysteresis and transient behaviors
were obtained by the introduction of PC61BM into CH3NH3PbI3
active layer, which indicates the dual effect of PC61BM in the
CH3NH3PbI3:PC61BM active layer: first, its role as the electron

www.advenergymat.de
6000
CH NH PbI
3

CH NH PbI :PC BM
3

PL Intensity

5000

61

PC BM/CH NH PbI :PC BM


61

61

4000

3000

2000

1000
650

700

750

800

850

Wavelength (nm)
Figure 4. Photoluminescence (PL) spectra of the pristine CH3NH3PbI3,
CH3NH3PbI3:PC61BM, and PC61BM/CH3NH3PbI3:PC61BM thin films.

acceptor provides more efficient pathways for electron


extraction from CH3NH3PbI3 to PC61BM; second, crystal growth
and size of CH3NH3PbI3 would be improved by the addition of
PC61BM (see discussion in late context), which are expected to
bring about more favorable polarizability and long-range charge
carrier and ionic transport within the perovskite absorber.[46] This
synergistic functions of PC61BM will be profoundly explored by
investigating the BHJ composites thin film.
In order to understand underlying of the tremendously increased JSC and FF in the conventional BHJ peroHSCs, photoluminescence (PL) properties of the three thin
films, CH3NH3PbI3, CH3NH3PbI3:PC61BM, and PC61BM/
CH3NH3PbI3:PC61BM, are studied. The PL spectra of
CH3NH3PbI3,
CH3NH3PbI3:PC61BM,
and
PC61BM/
CH3NH3PbI3:PC61BM thin films are shown in Figure 4. A more
strikingly quenching effect in CH3NH3PbI3:PC61BM than that
in CH3NH3PbI3 was found, which further confirms the role of
PC61BM as the electron acceptor for efficient exciton dissociation
1.2
CH NH PbI
3

CH NH PbI :PC BM

Absorbance(a.u.)

FULL PAPER

www.MaterialsViews.com

61

PC61BM
0.8

0.6

0.4

0.2

0
400

500

600

700

800

Wavelength (nm)
Figure 5. Absorption spectroscopy of CH3NH3PbI3, CH3NH3PbI3:PC61BM,
and PC61BM thin films.

1402024 (4 of 7)

wileyonlinelibrary.com

at donor/acceptor interfaces and subsequently electron extraction. Light emission from the electronhole pair recombination
was enormously suppressed, resulting in significantly enhanced
JSC and FF in the conventional BHJ pero-HSCs.[47] Even more
impressive PL quenching effect was manifested by PC61BM/
CH3NH3PbI3:PC61BM thin film, which was originated from the
enhanced electron transport at the interface between the EEL
(PC61BM) and the CH3NH3PbI3:PC61BM BHJ active layer.
In order to verify the hypothesis of crystallizations of
CH3NH3PbI3 affected by PC61BM, absorption spectra of
pristine CH3NH3PbI3 thin film, CH3NH3PbI3:PC61BM thin
film and PC61BM thin film were measured. As shown in
Figure 5, a red-shifted absorption spectrum is observed from
CH3NH3PbI3:PC61BM thin film, which indicates more compact crystal ordering of CH3NH3PbI3 upon the introduction
of PC61BM was formed.[48] This compact crystal structure of
CH3NH3PbI3 by PC61BM is certainly favorable for charge carrier being transported from CH3NH3PbI3:PC61BM thin film
to the corresponding electrodes, resulting in enhanced JSC. In
addition, PC61BM contributes to the absorption in the region
from 350 to 450 nm, which is also one of the origins for the
enhanced JSC in the conventional BHJ pero-HSCs.
X-ray diffraction (XRD) was further performed to study how
the crystallization changed with the introduction of PC61BM in
CH3NH3PbI3 layer. Figure 6a,b presents XRD patterns of the
pristine CH3NH3PbI3 thin film and the CH3NH3PbI3:PC61BM
BHJ thin film under different annealing periods, from
0 min to 120 min. From the first annealing stage (0 min)
(Figure 6a), a prompt reaction between PbI2 and CH3NH3I
can be observed from the diffraction peaks at 14.1 and 40.6,
which are assigned to the (110) and (224) planes in perovskite
orthorhombic crystal structure. After annealing for 60 min,
the formation of orthorhombic crystal structure is almost
completed. In clear contrast, Figure 6b reveals an obviously
retarded reaction between PbI2 and CH3NH3I upon the addition of PC61BM, presenting barely minor diffraction peak at
14.1 that belongs to the perovskite orthorhombic crystal structure in the nonannealing stage. Even after 60 min annealing
period, the sharp peaks at 12.3, 25.3, 38.5 belonging to the
PbI2, and 19.8 belonging to the CH3NH3I declare the incomplete reaction between PbI2 and CH3NH3I. The retarded reaction may be attributed to the formation of PC61BM-CH3NH3I
intermediate phase, which is originated from the interactions
between the ester groups in PC61BM with CH3NH3I. This
retarded reaction facilitates the development of more uniform
and dense perovskite layer,[49] yielding greatly improved FF in
CH3NH3PbI3:PC61BM based BHJ pero-HSCs.
Consequently, top view scanning electron microscopy (SEM)
images were conducted to inspect the thin films morphology of
pristine CH3NH3PbI3 and CH3NH3PbI3:PC61BM BHJ composites. As shown in Figure 7, obvious larger perovskite grains in
CH3NH3PbI3:PC61BM thin film were observed, which is consistent with the observations from UVvis absorption spectra
and XRD patterns. Given that the grain boundaries are generally the recombination sites, the larger grains with less grain
boundaries are favorable for the charge carrier transportation,
which consequently inhibits the recombination within the BHJ
composites activelayer.[50] The results decipher the intensively
enhanced FF and the minor raised VOC.[48]

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

Adv. Energy Mater. 2015, 1402024

www.advenergymat.de
www.MaterialsViews.com

FULL PAPER
Figure 6. Wide angle X-ray diffraction (WAXD) spectra of a) pristine CH3NH3PbI3 thin film under different annealing times for 0 min, 60 min, and
120 min; b) CH3NH3PbI3:PC61BM thin film under different annealing times for 0 min, 60 min, and 120 min.

Finally, AC impedance spectroscopy (IS) was performed


to investigate the electrical conductivities, which can provide
detailed electrical properties of pero-HSCs that cannot be determined in direct current measurement.[51] Figure 8 presents the
IS spectra of the conventional PHJ pero-HSCs, conventional
BHJ pero-HSCs, and PC61BM modified conventional BHJ peroHSCs. The internal series resistance (RS) is composed of the
sheet resistance (RSH) of the electrodes, the charge-transfer
resistance (RCT) inside the perovskite thin film and at perovskite material/EEL (HEL) interfaces. Since all pero-HSCs possess the same device structure, the RSH was assumed to be
the same. The only difference is the RCT which arises from
the different electron transport at EEL/perovskite interfaces.
RS of the conventional PHJ pero-HSCs is 983 , which is lowered to 588 for the conventional BHJ pero-HSCs, and further reduced to 500 for PC61BM modified conventional BHJ
pero-HSCs. These results are in good agreement with previous
characterizations: decreased RS for the conventional BHJ peroHSCs compared to the conventional PHJ pero-HSCs is attributed to both increased PC61BM/CH3NH3PbI3 interfaces and the

enhanced domain size of the perovskite crystals, which exert


combined actions to expedite charge carrier transporting within
the BHJ active layer. Further reduced RS for PC61BM modified
conventional BHJ pero-HSCs verifies the function of PC61BM
as more efficient EEL than that of TiO2.

3. Conclusion
In conclusion, the conventional BHJ pero-HSCs fabricated by
CH3NH3PbI3:PC61BM, where CH3NH3PbI3 acts as the electron
donor and PC61BM as the electron acceptor, were reported. In
the conventional BHJ pero-HSCs, PC61BM contributed for both
enhanced the donor/acceptor interfaces and improved crystal
growth of the CH3NH3PbI3, resulting in PCE of 12.78%, which
is twofold enhancement in comparison with the conventional
PHJ pero-HSCs (6.90%). In order to further boost the efficiency, PC61BM was casted on top of solution-processed TiO2
from o-DCB solution to form bilayer electron extraction layer in
the conventional BHJ pero-HSCs. Owing to the tremendously

Figure 7. Top view scanning electron microscope (SEM) images of a) CH3NH3PbI3 thin film and b) CH3NH3PbI3:PC61BM thin film.

Adv. Energy Mater. 2015, 1402024

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

wileyonlinelibrary.com

(5 of 7) 1402024

www.advenergymat.de

1200
R

CT

1000

C
800

-ImZ(ohm)

FULL PAPER

www.MaterialsViews.com

Conventional PHJ pero-HSCs


Conventional BHJ pero-HSCs
PC BM modified conventional BHJ pero-HSCs
61

600

400

200

0
0

200

400

600

800

1000

1200

ReZ(ohm)
Figure 8. Nyquist plots at V VOC for perovskite hybrid solar cells.

higher electrical conductivity of PC61BM over that of TiO2, a


high PCE of 14.98% was achieved. The novel structure successfully provided a solution to address two major problems of
pero-HSCs: the shorter diffusion length of the electrons than
that of the holes in perovskite materials, which limits further
boosting the efficiency of conventional PHJ pero-HSCs; and
the photocurrent hysteresis of pero-HSCs, which limits largescale manufacturing of pero-HSCs. The resultant efficiency
is the highest reported efficiency for the pero-HSCs without
incorporated with mesoporous TiO2 and Al2O3 as the electron
extraction layer and insulating scaffold, while the simplified
manufacturing process makes it possible for realizing the lowcost and high-throughput manufacturing of pero-HSCs.

on the top of TiO2 layer, followed with thermal annealing at 100 C for
2 h. The thickness of CH3NH3PbI3 is measured to be 280 nm. For the
conventional BHJ pero-HSCs, PC61BM:PbI2 solution (dissolved in DMF
with small amount of dichlorobenzene (o-DCB)) and MAI (dissolved
in ethanol) with concentrations of 400 mg mL1 and 35 mg mL1,
respectively, were spun on the top of either TiO2 layer or PC61BM/TiO2
layer, followed with thermal annealing at 100 C for 2 h. The thickness of
CH3NH3PbI3:PC61BM is 280 nm. The PC61BM/TiO2 layer was fabricated
by depositing 50 nm PC61BM layer on the top of compact TiO2 layer
from o-DCB solution. For all pero-HSCs, 80 nm thick P3HT was spin
coated on the top of either CH3NH3PbI3 or CH3NH3PbI3:PC61BM
layer from o-DCB (2% wt) solution. Afterward, 8 nm thick MoO3 and
100 nm thick Ag were thermal evaporated. The device area is defined
to be 0.045 cm2.
All pero-HSCs are characterized under an AM 1.5 G calibrated
solar simulator (Newport model 91160-1000) with the light intensity
of 100 mW cm2, which was calibrated by utilizing a mono-silicon
detector (with KG-5 visible color filter) of National Renewable Energy
Laboratory (NREL) to reduce spectral mismatch. The current densities
versus voltage (JV) characteristics of pero-HSCs were recorded using
a Keithley 2400 source meter. The incident photon to current efficiency
(IPCE) was measured through the IPCE measurement setup in use
at ESTI for cells and mini-modules. A 300 W steady-state xenon lamp
provides the source light. Up to 64 filters (820 nm width, range from
300 to 1200 nm) are available on four filter-wheels to produce the
monochromatic input, which is chopped at 75 Hz, superimposed on the
bias light, and measured via the usual lock-in technique. After collecting
the IPCE data, the software also integrates the date with the AM1.5G
spectrum and gives the calculated JSC value, which is helpful for checking
the accuracy of the measurement.
The impedance spectroscopy (IS) was obtained using a HP 4194A
impedance/gain-phase analyzer, under the illumination of white light
with the light intensity of 100 mW cm2, with an oscillating voltage of
50 mV and frequency of 5 Hz to 13 MHz.

Acknowledgements
The authors thank NSF for financial support from (ECCS 1351785).
Received: November 12, 2014
Revised: February 10, 2015
Published online:

4. Experimental Section
Materials: TiO2 precursor, tetrabutyl titanate (TBT), and PC61BM were
purchased from Sigma-Aldrich and Nano-C Inc., respectively, used as
received without further purification. Lead iodine (PbI2) was purchased
from Alfa Aesar. Methylammonium iodide (CH3NH3I, MAI) was
synthesized in our lab using the method reported in the literatures.[28]
Thin Film Characterizations: Thicknesses of all thin films were
measured by tapping-mode atomic force microscopy (AFM)
images using a NanoScope NS3A system (Digital Instrument).
Photoluminescence (PL) spectra were obtained with a 532 nm pulsed
laser as excitation source at a frequency of 9.743 MHz. UVvis
absorption spectra of pristine CH3NH3PbI3 and CH3NH3PbI3:PC61BM
thin films were measured using the HP 8453 spectrophotometer. Wide
angle X-ray diffraction (WAXD) spectra were measured by Bruker AXS
Dimension D8 X-Ray System. Scanning electron microscope (SEM)
images were measured by Model JEOL-7401 Japan Electron Optics
Laboratory (JEOL).
pero-HSCs Fabrication and Characterization: The 60 nm thick
compact TiO2 layer was deposited on pre-cleaned ITO substrates from
tetrabutyl titanate (TBT) isoproponal solution (concentration 3 vol%),
followed with thermal annealing at 90 C for 60 min in the ambient
atmosphere. For the conventional PHJ pero-HSCs, PbI2 (dissolved
in dimethylformamide (DMF)) and MAI (dissolved in ethanol) with
concentrations of 400 mg mL1 and 35 mg mL1, respectively, were spun

1402024 (6 of 7)

wileyonlinelibrary.com

[1] N.-G. Park, J. Phys. Chem. Lett. 2013, 4, 2423.


[2] N. Pellet, P. Gao, G. Gregori, T.-Y. Yang, M. K. Nazeeruddin,
J. Maier, M. Grtzel, Angew. Chem. Int. Ed. 2014, 53, 3151.
[3] I. Chung, B. Lee, J. He, R. P. H. Chang, M. G. Kanatzidis, Nature
2012, 485, 486.
[4] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem. Soc.
2009, 131, 6050.
[5] J. H. Im, C. R. Lee, J. W. Lee, S. W. Park, N. G. Park, Nanoscale 2011,
3, 4088.
[6] M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith,
Science 2012, 338, 643.
[7] H. J. Snaith, J. Phys. Chem. Lett. 2013, 4, 3623.
[8] S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou,
M. J. P. Alcocer, T. Leijtens, L. M. Herz, A. Petrozza, H. J. Snaith,
Science 2013, 342, 341.
[9] A. Mei, X. Li, L. Liu, Z. Ku, T. Liu, Y. Rong, M. Xu, M. Hu, J. Chen,
Y. Yang, M. Grtzel, H. Han, Science 2014, 345, 295.
[10] L. Etgar, P. Gao, Z. Xue, Q. Peng, A. K. Chandiran, B. Liu,
M. K. Nazeeruddin, M. Grtzel, J. Am. Chem. Soc. 2012, 134,
17396.

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

Adv. Energy Mater. 2015, 1402024

www.advenergymat.de
www.MaterialsViews.com

Adv. Energy Mater. 2015, 1402024

[31] B. Yang, Y. Yuan, P. Sharma, S. Poddar, R. Korlacki, S. Ducharme,


A. Gruverman, R. Saraf, J. Huang, Adv. Mater. 2012, 24,
1455.
[32] Z. Xu, L.-M. Chen, M.-H. Chen, G. Li, Y. Yang, Appl. Phys. Lett. 2009,
95, 013301.
[33] C. Liu, K. Wang, P. Du, T. Meng, X. Yu, S. Z. D. Cheng, X. Gong,
ACS Appl. Mater. Interfaces 2015, 7, 1153.
[34] F. Brivio, A. B. Walker, A. Walsh, APL Mater. 2013, 1, 042111.
[35] L. Yang, B. Xu, D. Bi, H. Tian, G. Boschloo, L. Sun, A. Hagfeldt,
E. M. J. Johansson, J. Am. Chem. Soc. 2013, 135, 7378.
[36] B. ORegan, D. T. Schwartz, Chem. Mater. 1998, 10, 1501.
[37] P. Tiwana, P. Docampo, M. B. Johnston, L. M. Herz, H. J. Snaith,
Energy Environ. Sci. 2012, 5, 9566.
[38] K. A. Borup, J. de Boor, H. Wang, F. Drymiotis, F. Gascoin, X. Shi,
L. Chen, M. I. Fedorov, E. Muller, B. B. Iversena, G. J. Snyder, Energy
Environ. Sci. 2015, 8, 423.
[39] K. Wang, C. Liu, P. Du, J. Zheng, X. Gong, Energy Environ. Sci. 2015,
DOI: 10.1039/C5EE00222B.
[40] X. Bao, L. Sun, W. Shen, C. Yang, W. Chen, R. Yang, J. Mater. Chem.
A 2014, 2, 1732.
[41] S. R. Raga, M.-C. Jung, M. V. Lee, M. R. Leyden, Y. Kato, Y. Qi,
Chem. Mater. 2015, DOI: 10.1021/cm5041997.
[42] V. D. Mihailetchi, H. Xie, B. de Boer, L. J. A. Koster, P. W. M. Blom,
Adv. Funct. Mater. 2006, 16, 699.
[43] R. S. Sanchez, V. Gonzalez-Pedro, J.-W. Lee, N.-G. Park, Y. S. Kang,
I. Mora-Sero, J. Bisquert, J. Phys. Chem. Lett. 2014, 5, 2357.
[44] H. J. Snaith, A. Abate, J. M. Ball, G. E. Eperon, T. Leijtens,
N. K. Noel, S. D. Stranks, J. T.-W. Wang, K. Wojciechowski,
W. Zhang, J. Phys. Chem. Lett. 2014, 5, 1511.
[45] H.-S. Kim, N.-G. Park, J. Phys. Chem. Lett. 2014, 5, 2927.
[46] Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang,
Y. Gaobc, J. Huang, Energy Environ. Sci. 2014, 7, 2619.
[47] C. Liu, X. Hu, C. Zhong, M. Huang, K. Wang, Z. Zhang, X. Gong,
Y. Cao, A. J. Heeger, Nanoscale 2014, 6, 14297.
[48] P. Gao, M. Gratzel, M. K. Nazeeruddin, Energy Environ. Sci. 2014, 7,
2448.
[49] N. J. Jeon, J. H. Noh, Y. C. Kim, W. S. Yang, S. Ryu, S. Seok, Nat.
Mater. 2014, 13, 897.
[50] D. P. Joshi, Solid-Sate Electron. 1986, 29, 19.
[51] A. Dualeh, T. Moehl, N. Tetreault, J. Teuscher, P. Gao,
M. K. Nazeeruddin, M. Grtzel, ACS Nano 2014, 8, 362.

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

wileyonlinelibrary.com

(7 of 7) 1402024

FULL PAPER

[11] 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, 1.
[12] J. Qiu, Y. Qiu, K. Yan, M. Zhong, C. Mu, H. Yan, S. Yan, Nanoscale
2013, 5, 3245.
[13] M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith,
Science 2012, 338, 643.
[14] 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.
[15] A. Abrusci, S. D. Stranks, P. Docampo, H.-L. Yip, A. K. Y. Jen,
H. J. Snaith, Nano Lett. 2013, 13, 3124.
[16] B. Cai, Y. Xing, Z. Yang, W.-H. Zhang, J. Qiu, Energy Environ. Sci.
2013, 6, 1480.
[17] J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao,
M. K. Nazeeruddin, M. Grtzel, Nature 2013, 499, 316.
[18] M. Liu, M. B. Johnston, H. J. Snaith, Nature 2013, 501, 395.
[19] B. Conings, L. Baeten, C. De Dobbelaere, J. DHaen, J. Manca,
H.-G. Boyen, Adv. Mater. 2014, 26, 2041.
[20] G. E. Eperon, V. M. Burlakov, P. Docampo, A. Goriely, H. J. Snaith,
Adv. Funct. Mater. 2014, 24, 151.
[21] M. Liu, M. B. Johnston, H. J. Snaith, Nature 2013, 501, 395.
[22] D. Liu, T. L. Kelly, Nat. Photonics 2014, 8, 133.
[23] E. Edri, S. Kirmayer, A. Henning, S. Mukhopadhyay, K. Gartsman,
Y. Rosenwaks, G. Hodes, D. Cahen, Nano Lett. 2014, 14, 1000.
[24] J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal, S. I. Seok, Nano Lett.
2013, 13, 1764.
[25] S. Lv, L. Han, J. Xiao, L. Zhu, J. Shi, H. Wei, Y. Xu, J. Dong, X. Xu,
D. Li, S. Wang, Y. Luo, Q. Meng, X. Li, Chem. Commun. 2014, 50,
6931.
[26] J. M. Ball, M. M. Lee, A. Hey, H. J. Snaith, Energy Environ. Sci. 2013,
6, 1739.
[27] A. Abrusci, S. D. Stranks, P. Docampo, H.-L. Yip, A. K.-Y. Jen,
H. J. Snaith, Nano Lett. 2013, 13, 3124.
[28] J. H. Heo, S. H. Im, J. H. Noh, T. N. Mandal, C.-S. Lim, J. A. Chang,
Y. H. Lee, H. Kim, A. Sarkar, M. K. Nazeeruddin, M. Grtzel,
S. I. Seok, Nat. Photonics 2013, 7, 486.
[29] E. L. Unger, E. T. Hoke, C. D. Bailie, W. H. Nguyen, A. R. Bowring,
T. Heumller, M. G. Christoforo, M. D. McGehee, Energy Environ.
Sci. 2014, 7, 3690.
[30] H. Tang, K. Prasad, R. Sanjins, P. E. Schmid, F. Lvy, J. Appl. Phys.
1994, 75, 2042.

Vous aimerez peut-être aussi