Sie sind auf Seite 1von 7

www.advmat.

de

COMMUNICATION

www.MaterialsViews.com

Binary-Metal Perovskites Toward High-Performance


Planar-Heterojunction Hybrid Solar Cells
Fan Zuo, Spencer T. Williams, Po-Wei Liang, Chu-Chen Chueh, Chien-Yi Liao,
and Alex K.-Y. Jen*
Great effort has been made to find promising alternatives for
silicon solar cells to drive the widespread implementation of
photovoltaic energy production forward.[16] Among the innovations in this emerging field, hybrid organolead halide perovskite
solar cells have attracted rapidly increasing attention with power
conversion efficiencies (PCE) rocketing to 15.7% in less than one
year.[717] Hybrid organolead perovskites are a class of semiconductors with the chemical formula AMX3 (A = organic molecule,
M = Pb, X = Cl, Br, and I).[1820] Halide anions form a network of
corner sharing MX6 octahedra with Pb2+ cations located at the
octahedrons center. Organic cations occupy the cuboctahedral
voids between adjacent octahedra (Figure 1). The great interest
in using these hybrid perovskites stems from their unique
combination of properties that are critical for high photovoltaic
performance: 1) direct band gap with gap size ranging from semiconducting to metallic, tunable through choice of metals,[21,22]
halogens,[11] and organic cations;[23] 2) large dielectric coefficient
compared to organic polymers leading to small exciton binding
energy (20 meV),[24] long diffusion lengths (1001000 nm),
and long lifetimes (100 ns);[25,26] 3) low temperature (<100 C)
solution processability due to a self-assembly driven phase transformation;[19,27,28] and 4) prominent and consistent absorption
throughout the visible spectrum.[11]
There are several key challenges that need to be addressed
before this technology becomes feasible for practical applications. The accurate determination of perovskites electron/
hole mobility, chemical structure and its influence on charge
transport behavior, precise function of interfaces in device
architectures, factors influencing stability, and strategies to
replace lead as the divalent metal cation without sacrificing
device performance are the pressing concerns facing the
field.[29] Finding ways to create lead-free perovskites with good

Dr. F. Zuo, S. T. Williams, P.-W. Liang,


Dr. C.-C. Chueh, C.-Y. Liao, and Prof. A. K.-Y. Jen
Department of Materials Science and Engineering
University of Washington
Seattle, WA 98195, USA
E-mail: ajen@u.washington.edu
Prof. A. K.-Y. Jen
Department of Chemistry
University of Washington
Seattle, WA 98195, USA
C.-Y. Liao
Institute of Polymer Science and Engineering
National Taiwan University
Taipei 106, Taiwan

DOI: 10.1002/adma.201401641

6454

wileyonlinelibrary.com

stability is especially critical because Pb toxicity hampers the


practicality of perovskite-based photovoltaics.
Replacing Pb with Sn is promising since both Sn and Pb
belong to the IVA group and have similar ionic radii due to
relativistic effects (Sn2+ 1.35 and Pb2+ 1.49 ), which may
enable substitution without significant lattice perturbation.
The Sn-based perovskites have been reported to possess a narrower band-gap (as small as 1 eV) and much higher charge
mobility (102 103 cm2/Vs) compared to pure Pb perovksites
(10102 cm2/Vs).[21] Although this is encouraging, the PCE of
solution processed planar hetrojunction solar cells fabricated
from pure tin or mixed metal perovskites are typically lower
than those achievable through vapor deposition, primarily
owing to poor film morphology and coverage. This low coverage can cause severe shorting, poor charge transport, and fast
electron/hole recombination; therefore rational control of perovskite nucleation and film growth is critical for improving performance. Directly addressing the difficulties associated with
controlling the crystallization of solution cast perovskite films
is a necessary component of tackling these issues. Relevant
studies on this topic are very rare, and little has been done to
characterize the effect of Sn alloying on the phase transformation of organolead halide perovskites.
Very recently, Sn-based perovskite solar cells have been
reported to demonstrate PCEs around 56%.[3033] However,
all these devices are constructed on a mesoporous TiO2 scaffold, which requires a high temperature (500 C) sintering
process. In order to develop a simple and low temperature
solution processed Sn-based perovskite solar cell, we are interested in exploring the planar heterojunction architecture.
Herein, we report the synthesis of binary Pb-Sn perovskites
(CH3NH3Pb1aSnaI3xClx) and conduct a detailed investigation
of the effect of Sn2+ substitution on perovskite film growth and
optoelectronic properties. By optimizing synthetic conditions,
we demonstrate high quality perovskite films that can be grown
with coverage as high as 97% on PEDOT:PSS coated ITO substrate. Sn substitution is also found to red-shift the absorption
onset from 800 nm for pure Pb perovskite to 900 nm without
significantly affecting absorptivity. This is currently one of
the broadest absorption ranges reported for high-performing
perovskite-based devices.[11,23,3033] Weve achieved an exciting
PCE of 10.1% in a planar heterojunction device based on this
system compared to 7.3% in our purely lead based control
(CH3NH3PbI3xClx) as a result of the broadened absorption
and improved film quality induced by the inclusion of Sn.
The impact of enhanced film quality is evident in the greatly
improved fill factor (0.67 vs. 0.55) and short-circuit current
density (19.5 mA/cm2 vs. 14.5 mA/cm2). To the best of our

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2014, 26, 64546460

www.advmat.de
www.MaterialsViews.com

knowledge, this is the first reported PCE above 10% for solar
cells based on Pb/Sn alloy perovskites.
Unlike recently reported high-performance planar heterojunction perovskite solar cells requiring metal oxide layers and/
or vapor deposition,[13,34,35] our method is a simple solution process. Regardless of the improved performance, development of
this system provides a better understanding of the effect of Pb
substitution in organometal perovskites and it may eventually
lead to completely Pb-free systems to achieve the cost-effective
production of high efficiency and environmentally benign
photovoltaics.
Understanding the effect of Sn on perovskite crystallization is critical for controlling film morphology and thus device
performance. Therefore, we have systematically studied the
structure and morphology of organolead perovskites with differing concentrations of Sn. X-ray diffraction patterns of pure
and mixed Pb/Sn perovskites, CH3NH3Pb1aSnaI3xClx, deposited on PEDOTS:PSS/ITO substrates are shown in Figure S1,
where a = 0, 0.15, 0.5, 0.75, and 1. Calculated space groups and
lattice parameters are summarized in Table 1.
Within the full range of composition, all samples can be
indexed according to the perovskite structure. As a increases
from 0 to 1, corresponding to CH3NH3PbI3xClx (MAPbX)
and CH3NH3SnI3xClx (MASnX), a phase transition is clearly
observed. It has been reported that pure MAPbX and MASnX
have tetragonal I4cm and P4mm symmetry respectively at room
temperature.[21] Figure S2 shows an enlarged view of 2225
2. There are two peaks within this range when a 0.5, which
could be indexed to (211) and (202) planes in the tetragonal
I4cm space group, agreeing with the reported MAPbI3 structure
at room temperature.[36] The two peaks gradually disappear and
a single peak rises up corresponding to the (113) plane in the
P4mm space group when a becomes greater than 0.5 due to

COMMUNICATION

Figure 1. From left to right, SEM image for cross section, device configuration of planar heterojunction solar cell, and the crystalline structure for
MAPb1aSnaX.

the increased symmetry of P4mm. This phenomenon has also


recently been reported by Kanatzidis et al.[21]
There is a gradual and systematic peak shift to higher 2
when a changes from 0 to 1, demonstrating that as Sn2+ concentration increases the unit cell size consistently decreases. From
the calculated lattice parameters presented in Table 1, it is clear
that higher Sn2+ content results in a smaller unit cell, which is
anticipated since Pb2+ (1.49 ) is larger than Sn2+ (1.35 ). It is
worth noting that unit cell volume only changes slightly in this
series, which is consistent with previous reports asserting that
perovskite unit cell dimensions are primarily determined by
the size of halide anion/s, Cl, Br, and/or I.[37] Over the entire
composition range from one end member to the other, Sn and
Pb form a substitutional solid solution without inducing precipitation of a secondary non-perovskite phase. This solubility
is instrumental in the success of our binary system.
The XRD patterns of MAPbX, our control, and
MAPb0.85Sn0.15X, the optimum composition for photovoltaic
performance, both annealed for 120 min at 90 C, are compared
in Figure 2a and b. Upon introducing 15 mole% of Sn2+, peak
intensities significantly increase to 30 times those of MAPbX
suggesting a dramatic increase in crystallinity. More importantly,
unlike the MAPbX pattern which noticeably displays many crystallographic planes including (110), (200), (211), (202), (220),
(310), (312), (224), (314), (404), and (440), the MAPb0.85Sn0.15X
diffraction pattern is strikingly dominated by the (110), (220),
(330), and (440) peaks. This implies that the MAPb0.85Sn0.15X
has significantly greater texture and preferentially grows along
<110>. MAPb0.85Sn0.15X has a tetragonal structure with lattice
parameters a = b = 8.871 and c = 12.434 , close to those
reported for MAPbI3xClx.[38] It is worth noting that there is only
a weak peak at 2 = 12.66 corresponding to the (001) plane of
the PbI2, suggesting high phase purity.

Table 1. Elemental analysis and Structural Analysis for MAPb1aSnaX.


Sample ID
MAPb1Sn0X

Sn at%

Pb at%

Space Group

a ()

b ()

c ()
12.568

I4 cm

8.920

8.920

MAPb0.85Sn0.15X

0.12

0.88

I4 cm/P4mm

8.871

8.871

12.510

MAPb0.5Sn0.5X

0.44

0.56

I4 cm/P4mm

8.838

8.838

12.434

MAPb0.25Sn0.75X

0.75

0.25

P4 mm

6.258

6.258

6.681

P4 mm

6.220

6.220

6.708

MAPb0Sn1X

Adv. Mater. 2014, 26, 64546460

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

6455

www.advmat.de

COMMUNICATION

www.MaterialsViews.com

Figure 2. (a) and (b) XRD spectra for MAPbX and MAPb0.85Sn0.15X films respectively. (c) SEM images for MAPbX and MAPb0.85Sn0.15X films. (d) EDS
elemental maps of Pb and Sn in MAPb0.85Sn0.15X.

Characterization of MAPbX and MAPb0.85Sn0.15X films with


scanning electron microscopy (SEM) supports conclusions
drawn from the XRD data. The two top images in Figure 2c
show the morphology of MAPbX and MAPb0.85Sn0.15X films,
respectively. While the MAPbX film shows decent film
quality with good continuity and high coverage (87%), the
MAPb0.85Sn0.15X film displays far superior quality and coverage
(97%). This markedly increased coverage ensures excellent
connectivity between grains, which is crucial for carrier transport. The virtually pin-hole free film mitigates short-circuiting,
charge leaking, and large series resistance.
In the bottom images of Figure 2c, morphological differences
between MAPbX and MAPb0.85Sn0.15X films are clearly evident.
The MAPbX film is composed of interconnected nanoscale
domains with sizes ranging from 200 nm to 500 nm with no
obvious faceting. In the MAPb0.85Sn0.15X film, however, we
observe strong faceting with comparatively smooth surfaces suggesting enhanced, and potentially more defect free, crystal development. Apparent domain size is significantly larger (500 nm
1 m) than that expressed in the MAPbX film suggesting
an alteration of transformation kinetics. Regular faceting and
orientation of the MAPb0.85Sn0.15X crystallites reveal texture
suggested by the XRD data, demonstrating that {110} planes
are preferentially developed and exposed. Regular faceting surrounding interconnected features suggests that the film is composed of an oriented network of 500 nm to 1 m single crystals.
To elucidate the growth kinetics of two films, we have
studied their morphologies and phase compositions at different
annealing times with SEM (Figure 3) and XRD (Figure S3).
After spin-coating the precursor solutions onto the substrate,

6456

wileyonlinelibrary.com

ions in the precursor solutions immediately begin to selfassemble into the perovskite structure as is evidenced in the
XRD patterns for both MAPb0.85Sn0.15X and MAPbX measured
before annealing (Figure S3). The alteration in transformation kinetics induced by tin inclusion is evident in the massive
increase in the (110) reflection 14.2 2 after being annealed for
30 min in MAPb0.85Sn0.15X relative to MAPbX, which indicates
more rapid transformation. Peaks corresponding to both iodide
and chloride based organometal perovskites (at 2 = 14.2 and
15.9, respectively)[13,38] are evident in MAPb0.85Sn0.15X and
MAPbX at initial stage and after 30 min of annealing while only
the reflection corresponding to the iodide based organometal
perovskite remains after annealing for 120 min.
This indicates that in addition to being exist in the final
iodide based perovskite as substitutional impurities, chloride
ions also alter the transformation pathway during annealing
by introducing the competing chloride based perovskite phase.
This alteration in growth pathway likely influences overall
transformation kinetics in addition to modulating the resulting
morphology as compared to pure lead systems, but as of now
the exact mechanism by which chloride inclusion enhances
ultimate material properties is inadequately understood and an
area of active research in the field.[25,38,39] It is understood from
current literature that chloride ions preferentially occupy axial
positions in the mixed I/Cl lead octahedra within the perovskite
lattice while iodide ions may occupy both equatorial and axial
positions,[13,38] however it is difficult for us to make any conclusion to this end in this work. It seems apparent that tin inclusion suppresses the formation of the chloride based perovskite
lattice by virtue of the decreased peak height at 15.9 2 at

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2014, 26, 64546460

www.advmat.de
www.MaterialsViews.com

COMMUNICATION
Figure 3. Top images show the morphologies of MAPb0.85Sn0.15X films for annealing times of 0 min, 30 min, and 120 min. Bottom images show the
morphologies of MAPbX in the same manner. The scale bars indicate 5 m.

both 0 and 30 minutes in MAPb0.85Sn0.15X relative to MAPbX


(Figure S3).
Even before annealing, MAPb0.85Sn0.15X and MAPbX show
great morphological differences (Figure 3). Holes in both films
may result from the evaporation of dimethylformamide (DMF)
during spin-coating. Compared to the MAPbX sample, contrast
features apparent in the MAPb0.85Sn0.15X film suggest the formation of larger ordered domains, again implying a more rapid
transformation via self-assembly. Even in this as cast state, the
MAPb0.85Sn0.15X film demonstrates greater coverage and continuity, possibly due to more favorable interfacial energy in addition to the faster assembly. After annealing at 90 C for 30 min,
the differences between the systems become more pronounced.
MAPb0.85Sn0.15X crystals begin to show well-defined facets with
smooth surfaces, sharp edges, and preferential orientations not
apparent in the MAPbX sample. After 120 min of annealing,
the two films fully develop. The MAPb0.85Sn0.15X perovskite crystals fuse and grow together to share certain facets
(Figure 2c), significantly depressing crystallite boundaries. On
the contrary, the MAPbX crystals do interconnect but retain
more voids and a greater degree of discontinuity. The rapid perovskite conversion rate for MAPb0.85Sn0.15X relative to MAPbX
is also evidenced by the time-resolved XRD study as discussed
above and shown in Figure S3.
The effect of composition on morphology is shown in
Figure S4. As the mole fraction of Sn increases, domains
become larger. Surfaces of all Sn-containing films are relatively
smooth, but sharp edges and corners disappear along with the
widening of voids at concentrations greater than 15%, possibly
because of increasingly rapid transformation and unfavorable
interfacial energy.
Figure 2d shows an elemental map of MAPb0.85Sn0.15X
obtained via SEM energy-dispersive X-ray spectroscopy (EDS).
To avoid any interference from Sn in the ITO, these films are
spin-coated on bare glass coated with PEDOT:PSS. Sn and Pb
are homogeneously distributed throughout the film with no
apparent phase separation, which further supports the proposed solid solution. EDS analysis shows the real Pb:Sn ratio
to be 88:12 in the MAPb0.85Sn0.15X film, which is close to the

Adv. Mater. 2014, 26, 64546460

theoretical 85:15 ratio given by stoichiometry of the precursor


solutions. The EDS results for all perovskite compositions are
summarized in Table 1. All perovskites exhibit reasonable consistency between real and theoretical Pb:Sn ratios.
Chemical differences between Sn2+ and Pb2+ explain Sns
significant effect on nucleation and growth behavior. Their differences in size results in differing affinities toward halogen
anions, which is reflected by bond lengths and energies listed in
Table S1. Sn-Cl has the largest bond energy (350 kJ/mol), which
suggests that during nucleation Sn-Cl bonds form first followed by Pb-Cl, Sn-I, and Pb-I bonds. Although this is a kinetic
argument derived from thermodynamic data, the hypothesis is
supported by our EDS analysis of MAPb0.85Sn0.15X at differing
annealing times. Before annealing, the Pb:Sn and I:Cl ratios
are 2.3:1 and 1:1, respectively, which are much lower than the
same ratios produced after 2 h of annealing: 7.3:1 and 5.8:1.
After annealing for 30 min, the Pb:Sn and I:Cl ratios reach
intermediate values of 5.6:1 and 3.2:1 respectively, demonstrating a consistent trend.
It should be noted that it may appear from this data as if tin
is being steadily evolved as a volatile species during annealing,
but the initial Pb:Sn ratio noted above is artificially high.
From this EDS data, at the beginning of the annealing process the composition appears to be 30 at.% Sn with respect
to Pb, which is much higher than the stoichiometry of the precursor solution used (15 at.% Sn). The final composition after
annealing for 2h is 12 at.% Sn which is within the reasonable
range compared to the initial solution composition, considering
errors that may be introduced by surface topology and chemical environment in EDS characterization. Additionally, a significant loss of chloride is generally observed by both us and
others because of its relatively low concentration allowed in the
iodide based organolead perovskite lattice (4%).[38] Since we
have observed similar degree of chloride loss in both MAPbX
and MAPb1aSnaX systems, we think it should not be caused
by the presence of tin. With regard to the loss mechanism,
it has been found in the pure organolead iodide perovskite
system that iodide is lost through the decomposition of methylammonium iodide into volatile methylamine and hydroiodic

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

6457

www.advmat.de

COMMUNICATION

www.MaterialsViews.com

acid.[40] We surmise that an analogous loss mechanism is relevant for excess chloride in both MAPbX and MAPb1aSnaX
and we present this in Scheme S1. Thus, we conclude the trend
of morphological changes should not be relevant to the loss
of chloride since it is a common phenomenon observed in all
systems.
To appreciate the implications of these compositional trends,
it is necessary to understand the physical mechanism behind
SEM EDS, which is shown in Figure S5. The generated characteristic X-ray emission enables elemental analysis. The pearshaped character of the detection volume in EDS means that
this technique acquires more signals from a samples bulk
than its surface. Note that the electron accelerating voltage in
our EDS measurement is 15 keV, which means a penetration
depth of few micrometers, much greater than the thickness
of our perovskite films (350450 nm). Therefore, it is reasonable to propose that the elemental ratios reported above and in
Table 1 are most representative of the composition at the base
of the growing perovskite film, an event that is preferentially
initiated at the substrate/perovskite interface. Summarizing the
above data with this in mind, we observe an abnormally high
concentration of Sn and Cl at the substrate interface in the as
cast specimens which we thought it is due to initiation of the
nucleation process, suggesting a Sn-Cl bond generation event
before other bonds formation. This explains why the bulk of
the sample in Figure 3a has a much higher concentration of
Sn2+ and Cl than the theoretical value.
All the above demonstrates the unique influence of Sn2+ on
nucleation and growth behavior of these hybrid perovskites.

It regulates perovskite nucleation rate and ensuing crystal


development. At the optimal Sn2+ concentration a high quality
and almost completely continuous film is achieved.
Earlier, Kanatzidis et al. reported that the introduction of Sn
into organolead halide perovskites narrows the band gap, redshifting the absorption onset.[21] Band-gaps of pure iodide end
members MAPbI and MASnI are 1.6 and 1.2 eV, respectively.
Precisely determining the energy levels of MAPb1aSnaI3 is
still challenging and the reported values so far are quite scattering,[3033] potentially due to the ease of oxidation of the Sncontaining samples. Regardless of this, the band gap of our
resulting perovskites match well with their proposed electronic
structure. Figure 4a compares UV-visible absorption spectra
for the mixed halide MAPbX and MAPb0.85Sn0.15X. Upon
partial substitution with Sn, the absorption edge red-shifts to
900 nm. More importantly, the absorption from 600 nm to
800 nm is greatly enhanced, demonstrating the binary-metal
system's potential for expanding light harvesting capacity of
current perovskites. The absorption spectra for all compositions
are plotted in Figure S6 and a diminishing return is apparent
upon increasing Sn2+ substitution significantly beyond 15%,
resulting in dramatically decreased absorption below 700 nm
due to the low absorption coefficient of Sn.[21]
Photovoltaic performance of devices fabricated with perovskites of all compositions studied is presented in Table 2 and
external quantum efficiency (EQE) spectra for devices fabricated
with MAPbX and MAPb0.85Sn0.15X are shown in Figure 4b.
From 330 nm to 800 nm, MAPbX exhibits an average EQE of
60%. The EQE range of MAPb0.85Sn0.15X extends up to 900 nm,

Figure 4. (a) UV-vis absorption spectra, (b) EQE results and (c) JV curves for MAPbX and MAPb0.85Sn0.15X measured under simulated AM1.5 illumination and in the dark, (d) The relationship between 1/Rs and light intensity for MAPbX and MAPb0.85Sn0.15X devices.

6458

wileyonlinelibrary.com

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2014, 26, 64546460

www.advmat.de
www.MaterialsViews.com

Voc (V)

FF

Jsc (mA/cm2)

PCE (%)

0.91 0.02

0.52 0.025

13.3 1.0

7.01 0.22 (7.33)

15% Sn perovskite*

0.76 0.01

0.66 0.008

19.1 0.2

9.77 0.25 (10.10)

50% Sn perovskite

0.24 0.03

0.53 0.010

0.7 0.1

0.11 0.02

75% Sn perovskite

0.06 0.02

0.25 0.035

3.5 0.8

0.06 0.03

100% Sn perovskite

0.03 0.02

0.25 0.016

5.3 2.0

0.04 0.01

Pure Pb perovskite

*The

number in the parenthesis is the highest PCE.

which is one of the broadest EQE ranges achieved in highperformance hybrid perovsktie solar cell.[11,23,3033] Peak EQE
exceeds 80% near 450 nm, and across the entire range tested
(330 nm to 900 nm) the MAPb0.85Sn0.15X device shows marked
enhancement compared to the MAPbX device.
Figure 4c shows current density-voltage (JV) curves measured under simulated AM1.5 illumination (100 mW/cm2) and
in the dark, respectively. Both perovskites show good diode
behavior with relatively low leakage current in the dark JV
curves. The MAPb0.85Sn0.15X perovskite solar cell exhibits a
short-circuit current dentsity (Jsc) of 19.5 mA/cm2, an opencircuit voltage (Voc) of 0.77 V, and a fill factor (FF) of 0.67,
yielding an overall power conversion efficiency (PCE) of 10.1%.
Relevant results for the MAPbX perovskite solar cell are Jsc =
14.5 mA/cm2, Voc = 0.92 V, and FF = 0.55. The MAPb0.85Sn0.15X
device shows a remarkably improved Jsc, which agrees well with
its broadened absorption and superior EQE. In addition to the
substantial increase in Jsc, Sn substitution enables a significant
FF improvement from 0.55 to 0.67. As both devices are fabricated with the same architecture, electrode and base contact
resistances are consistent. Thus, the increase in FF points to a
decrease in internal and interfacial resistances of the perovskite
absorber in MAPb0.85Sn0.15X due to the greatly improved perovskite film quality and coverage as compared to the pure lead
based control.
As evidenced in the results from XRD and SEM, the
MAPb0.85Sn0.15X film possesses much enhanced crystallinity
and surface coverage compared to those from the pure Pb
sample, suggesting a more ordered arrangement of atoms and
less defects. Shown in Figure 4d is the plot 1/Rs vs. light intensity, where the series resistance (Rs) is estimated from the slope
of the JV curve near Voc. As can be seen, MAPb0.85Sn0.15X has
a steeper slope than MAPbX, suggesting that carrier transport
is significantly improved via partially replacement of Pb2+ with
Sn2+. It can be speculated that the improved film coverage and
homogeneity enable more effective exciton dissociation and
facilitate the charge transport across the interfacial layers to
the electrodes while the poor film formation and coverage will
induce severe charge recombination. When a 0.5, Voc dramatically decreases. Takahashi et al. has attributed this to Sn2+
oxidation and the associated p-type doping.[22] Moreover, poor
fill factors with a 0.5 can be easily explained by consistently
inferior film quality as is evident in Figure S4.
In conclusion, we have demonstrated the success of partial Sn substitution in organolead perovskites to simultaneously reduce Pb content and increase performance through
broadened absorption and improved solution-cast film

Adv. Mater. 2014, 26, 64546460

COMMUNICATION

Table 2. Performance of the studied solar cells under AM 1.5G Illumination (100 mW/cm2).

morphology. Time-resolved SEM study of perovskite growth


during annealing indicates the dynamic morphological modulation offered by Sn2+ incorporation. A high coverage (97%)
film with excellent continuity can be grown because of Sns
effect on nucleation and growth. The markedly improved
morphology with well-ordered and relatively large crystalline
domains suppresses charge recombination and improves transport, resulting in an increased fill factor. UV-visible absorption
measurements demonstrate the 100 nm red-shift of the binarymetal perovskites absorbance onset. The EQE study further
verifies both the red-shifted absorption and efficiency enhancement that spans the entire characterized range from 330 nm to
900 nm. The enhancement in solar light absorption contributes
to the large Jsc as do the dramatic improvements in film quality
uniquely enabled in this binary system. By optimizing Sn concentration in MAPb1aSnaX, a very promising PCE of 10.1% is
achieved. To compliment these findings, the rigorous correlation between domain size and perovskite optoelectronic properties will be elucidated in the near future.
Our results are quite encouraging considering that partial Sn
substitution not only reduces the content of Pb in hybrid perovskite photoactive layers, mitigating their ecological impact,
but also enables higher PCE compared to the purely lead based
perovskite devices, enhancing their economic viability. This preliminary study establishes the foundation for further exploring
the benefits of alloy perovskite systems to replace lead while
simultaneously increasing performance to make these hybrid
photovoltaics competitive with Si-based technologies.

Experimental Section
Materials and Sample Preparation: Methylammonium iodide (MAI)
was synthesized by reacting 24 mL, 0.20 mole methylamine (33 wt% in
absolute ethanol, Aldrich), 10 mL, 0.04 mole hydroiodic acid (57 wt%
in water with 1.5% hypophosphorous acid, Alfa Aesar), and 100 mL
ethanol in a 250 mL round bottom flask under nitrogen protection at
0 C for 2 h with magnetic stirring. After reaction, the white precipitate
of MAI was recovered by rotary evaporating solvents at 40 C then
dissolved in ethanol and sedimentated in diethyl ether by stirring the
solution for 30 min. This step was repeated three times then the MAI
powder was collected and dried at 50 C in a vacuum oven for 24 h.
To prepare perovskite precursor solution, MAI and lead chloride (PbCl2,
98%, Aldrich) powder were mixed in anhydrous dimethylforamide
(DMF, Aldrich) with a molar ratio of 3:1. For the Sn-based perovskite
precursor solution, MAI, PbCl2, and Tin Chloride (SnCl2, 98%, Aldrich)
powder were mixed with desired molar ratio. The solutions (40 wt%)
were stirred overnight at 80 C and filtered through 0.45 m PVDF filters
before device fabrication.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

6459

www.advmat.de

COMMUNICATION

www.MaterialsViews.com
Fabrication of the Thin Film Perovsktie Devices: The thin film perovskite
devices were fabricated with the configuration: indium tin oxide (ITO)/
poly(3,4-ethylenedioxythiophene):poly(p-styrene sulfonate) (PEDOT:PSS,
CLEVIOSTM P VP Al 4083)/CH3NH3Pb1aSnaI3xClx/[6, 6]-phenyl-C61
-butyric acid methyl ester (PCBM, American Dye Source Inc.)/fullerene
surfactant (C60-bis)/Ag. ITO glass substrates (15 ohm/sq) were cleaned
sequentially by sonication in detergent and deionized water, acetone,
and isopropanol for 10 min. After drying under N2 stream, substrates
were further cleaned by exposing to plasma for 10 s. A PEDOT:PSS
(Baytron P VP Al 4083 filtered through a 0.45 m Nylon filter) holetransporting layer with a thickness of 45 nm was spin-coated onto ITO
substrates at 5k rpm for 30 s and annealed at 150 C for 10 min in air.
The substrates were then transferred into a N2-filled glove box, spincoated with prepared precursor solution at 6k rpm for 45 s, and annealed
at 90 C for 2 h. The thickness of the studied perovskite thin films is
around 350450 nm, determined by surface profiler (Dektak 3030).
The electron-transporting layer (PCBM, 15 mg/mL in chloroform) and
C60-bis surfactant (2 mg/mL in isopropyl alcohol) was then sequentially
deposited by spin coating at 1 k rpm for 60 s and 3k rpm for 60 s,
respectively. Silver electrodes with a thickness of 150 nm were deposited
under high vacuum (<2 106 Torr) through a shadow mask, defining
a device area of 3.14 mm2, by thermal evaporation. All JV curves in
this study were recorded using a Keithley 2400 source meter unit. The
device photocurrent was measured under AM1.5 illumination condition
at an intensity of 100 mW/cm2. The illumination intensity of the light
source was accurately calibrated with a standard Si photodiode detector
equipped with a KG-5 filter, which can be traced back to the standard cell
of the National Renewable Energy Laboratory (NREL). The EQE spectra
performed here were obtained from an IPCE setup consisting of a Xenon
lamp (Oriel, 450 W) as the light source, a monochromator, a chopper
with a frequency of 100 Hz, a lock-in amplifier (SR830, Stanford Research
Corp), and a Si-based diode (J1157111-Si detector) for calibration.

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

Acknowledgements
The authors thank the support from the Air Force of Scientific Research
(FA95500910426), the Office of Naval Research (N00014111
0300), the Asian Office of Aerospace R&D (FA23861114072), the
Boeing Foundation, and the National Science Foundation Graduate
Research Fellowship Program under Grant No. DGE-1256082. Part of
this work was conducted at the University of Washington NanoTech User
Facility, a member of the NSF National Nanotechnology Infrastructure
Network (NNIN).
Received: April 11, 2014
Revised: May 30, 2014
Published online: August 14, 2014
[1] M. A. Green, K. Emery, Y. Hishikawa, W. Warta, E. D. Dunlop, Prog.
Photovoltaics 2013, 21, 827.
[2] K. L. Chopra, P. D. Paulson, V. Dutta, Prog. Photovoltaics 2014, 12, 69.
[3] G. Li, R. Zhu, Y. Yang, Nature Photon. 2012, 63, 153.
[4] G. Li, V. Shrotriya, J. S. Huang, Y. Yao, T. Moriarty, K. Emery, Y. Yang,
Nature Mater. 2005, 4, 864.
[5] B. Oregan, M. Gratzel, Nature 1991, 353, 737.
[6] M. Law, L. E. Greene, J. C. Johnson, R. Saykally, P. D. Yang, Nature
Mater. 2005, 4, 455.
[7] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem. Soc.
2009, 131, 6050.
[8] N. G. Park, J. Phys. Chem. Lett. 2013, 4, 2423.

6460

wileyonlinelibrary.com

[9] 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. Gratzel,
N. G. Park, Sci. Rep. 2012, 2, 591.
[10] M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami, H. J. Snaith,
Science 2012, 338, 643.
[11] J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal, S. I. Seok, Nano Lett.
2013, 13, 1764.
[12] L. Etgar, P. Gao, Z. S. Xue, Q. Peng, A. K. Chandiran, B. Liu,
M. K. Nazeeruddin, M. Gratzel, J. Am. Chem. Soc. 2012, 134, 17396.
[13] M. Z. Liu, M. B. Johnston, H. J. Snaith, Nature 2013, 501, 395.
[14] 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, M. K. Nazeeruddin, M. Grtzel,
S. I. Seok, Nature Photon. 2013, 7, 487.
[15] J. M. Ball, M. M. Lee, A. Hey, H. J. Snaith, Energ. Environ. Sci. 2013,
6, 1739.
[16] J. Y. Jeng, Y. F. Chiang, M. H. Lee, S. R. Peng, T. F. Guo, P. Chen,
T. C. Wen, Adv. Mater. 2013, 25, 3727.
[17] D. Liu, T. L. Kelly, Nature Photon. 2014, 8, 133.
[18] D. B. Mitzi, J. Chem. Soc., Dalton Trans. 2001, 1, 1.
[19] Z. Y. Cheng, J. Lin, Crystengcomm 2010, 12, 2646.
[20] B. Cai, Y. D. Xing, Z. Yang, W. H. Zhang, J. S. Qiu, Energ. Environ.
Sci. 2013, 6, 1480.
[21] C. C. Stoumpos, C. D. Malliakas, M. G. Kanatzidis, Inorg. Chem.
2013, 52, 9019.
[22] Y. Takahashi, R. Obara, Z. Z. Lin, T. Naito, T. Inabe, S. Ishibashi,
K. Terakura, Dalton Trans. 2011, 40, 5563.
[23] T. M. Koh, K. W. Fu, Y. N. Fang, S. Chen, T. C. Sum, N. Mathews,
S. G. Mhaisalkar, P. P. Boix, T. Baikie, J. Phys. Chem. C 2014,16458.
[24] S. Sun, T. Salim, N. Mathews, M. Duchamp, C. Boothroyd,
G. C. Xing, T. C. Sum, Y. M. Lam, Energ. Environ. Sci. 2014, 7, 399.
[25] 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.
[26] G. C. Xing, N. Mathews, S. Y. Sun, S. S. Lim, Y. M. Lam, M. Gratzel,
S. Mhaisalkar, T. C. Sum, Science 2013, 342, 344.
[27] D. B. Mitzi, Chem. Mater. 1996, 8, 791.
[28] D. B. Mitzi, J. Mater. Chem. 2004, 14, 2355.
[29] M. A. Loi, J. C. Hummelen, Nature Mater. 2013, 12, 1087.
[30] Y. Ogomi, A. Morita, S. Tsukamoto, T. Saitho, N. Fujikawa, Q. Shen,
T. Toyoda, K. Yoshina, S. S. Pandey, T. L. Ma, S. Hayase, J. Phys.
Chem. Lett. 2014, 5, 1004.
[31] F. Hao, C. C. Stoumpos, D. H. Cao, R. P. H. Chang,
M. G. Kanatzidis, Nature Photon. 2014, 8, 489.
[32] F. Hao, C. C. Stoumpos, R. P. H. Chang, M. G. Kanatzidis, J. Am.
Chem. Soc. 2014, 136, 8094.
[33] N. K. Noel, S. D. Stranks, A. Abate, C. Wehrenfennig, S. Guarnera,
A. Haghighirad, A. Sadhanala, G. E. Eperon, M. B. Johnston,
A. Petrozza, L. M. Herz, H. J. Snaith, Energ. Environ. Sci. 2014,
DOI: 10.1039/C4EE01076K.
[34] Q. Chen, H. P. Zhou, Z. R. Hong, S. Luo, H. S. Duan, H. H. Wang,
Y. S. Liu, G. Li, Y. Yang, J. Am. Chem. Soc. 2014, 136, 622.
[35] B. Conings, L. Baeten, C. D. Dobbelaere, J. DHaen, J. Manca,
H. G. Boyen, Adv. Mater. 2014, 26, 2041.
[36] J. H. Im, C. R. Lee, J. W. Lee, S. W. Park, N. G. Park, Nanoscale 2011,
3, 4088.
[37] S. J. Clark, J. D. Donaldson, J. A. Harvey, J. Mater. Chem. 1995, 5,
1813.
[38] S. Colella, E. Mosconi, P. Fedeli, A. Listorti, F. Gazza, F. Orlandi,
P. Ferro, T. Besagni, A. Rizzo, G. Calestani, G. Gigli, F. De Angelis,
R. Mosca, Chem. Mater. 2013, 25, 4613.
[39] E. Mosconi, A. Amat, M. K. Nazeeruddin, M. Gratzel, F. De Angelis,
J. Phys. Chem. C 2013, 117, 13902.
[40] G. D. Niu, W. Z. Li, F. Q. Meng, L. D. Wang, H. P. Dong, Y. Qiu, J.
Mater. Chem. A 2014, 2, 705.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2014, 26, 64546460

Das könnte Ihnen auch gefallen