Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Applied Materials Today 7 (2017) 112–119

Contents lists available at ScienceDirect

Applied Materials Today


journal homepage: www.elsevier.com/locate/apmt

TiO2 colloid-based compact layers for hybrid lead halide perovskite solar cells
Sateesh Prathapani a,∗ , Venumadhav More b , Siva Bohm a,c , Parag Bhargava a , Aswani Yella a ,
Sudhanshu Mallick a,∗
a
Department of Metallurgical Engineering & Materials Science, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
b
Centre for Research in Nanotechnology & Science, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
c
Tata Steel R&D, Swinden Technology Center, Moorgate, Rotherham, South Yorkshire S60 3AR, UK

a r t i c l e i n f o a b s t r a c t

Article history: In this report, we present a simple method of TiO2 colloidal preparation and its application to form
Received 1 October 2016 compact conformal TiO2 (c-TiO2 ) layers on transparent conducting substrates. The synthesized colloid
Received in revised form 15 January 2017 has ∼5 nm particle size distribution and is stable over a period exceeding one year. The blocking nature
Accepted 27 January 2017
of c-TiO2 is evaluated by the method of cyclic voltammetry using Spiro-OMeTAD as the redox probe.
This method of c-TiO2 preparation is successfully employed in the fabrication of hybrid perovskite solar
Keywords:
cells. The critical requirement of c-TiO2 colloid-based compact layer as hole-blocking material is justified
TiO2 colloids
with the realization of photo-conversion efficiencies of 8.08% for small active areas of 0.07 cm2 and 7.52%
TiO2 nanoparticles
Titania nanoparticles
with relatively large active areas of 0.23 cm2 in one-step method. In two-step sequential deposition
TiO2 compact layer for perovskite solar cells method, a photo-conversion efficiency of 4.61% is achieved over relatively large active area of 0.64 cm2 .
Our observations suggest that c-TiO2 formation by TiO2 colloid can be an alternative potential route to
approach for perovskite solar cell fabrication. All device processing is carried out in ambient air of relative
humidity 40 ± 5% at room temperature.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction extract electrons from it, (iv) hybrid perovskite as light harvester,
(v) hole transport material (HTM) to extract holes from perov-
Organic–inorganic metal halide perovskite materials are the skite and (vi) metal Au as back electrode [7,10,11]. In literature,
new class of hybrid semiconductors with the general formula ABX3 , perovskite processing is predominantly reported by two methods
where A and B are organic and inorganic cations and X (Cl, Br and I) is of depositions, i.e. one-step deposition and two-step sequential
the halide anion, respectively. The exceptional physical properties deposition [7,10–12]. In one-step deposition method, a perovskite
of hybrid perovskite materials like a tunable band gap [1,2], high solution in the solvent dimethylformamide (DMF) is formed by
absorption coefficient [3], and long-range charge transport with mixing the precursors lead chloride (PbCl2 ) and methylammonium
high mobilities [4] have brought about a surge of interest in the iodide (CH3 NH3 I) in 1:3 molar ratio, respectively. As-prepared
optoelectronic device community to seek hybrid perovskite mate- solution is infiltrated into mp-TiO2 substrates and subsequently
rials as potential candidates for solar cell fabrication [5–7]. The subjected to an annealing process (at 100 ◦ C for 40 min) to form
semiconductor research community has put considerable effort perovskite thin film. In two-step sequential deposition method, 1 M
in aspects of compositional tailoring and device engineering, and lead iodide (PbI2 ) solution in DMF is made and coated on to the mp-
hence have improved the photo-conversion efficiencies of 3.81% TiO2 substrates. On annealing for 20 min at 100 ◦ C, solid PbI2 films
from the first report to as high as 19.3% and 22.1% [5,8,9]. are formed. Subsequently, these PbI2 films on mp-TiO2 are dipped
A typical hybrid perovskite solar cell (PSC) device consists of a in MAI solution (8 mg/ml in 2-propanol) to form perovskite layer at
total of the following six layers: (i) FTO (fluorine-doped tin oxide) room temperature. From the literature and with our experimental
as transparent electrode, (ii) c-TiO2 as electron transport material observations, we found that it is a challenging task to form a contin-
(ETM), which can additionally block the hole from reaching FTO, uous pinhole-free perovskite film on mp-TiO2 [13,14]. Moreover, if
(iii) mesoporous TiO2 (mp-TiO2 ) to infiltrate light harvester and to the processing is carried out in ambient air, the moisture in the air
can be the additional factor that can alter the crystallization kinet-
ics of perovskite materials and thus its morphology [13]. One-step
∗ Corresponding authors.
deposition method leaves island kind of discontinued perov-
E-mail addresses: sateesh.prathapani@gmail.com (S. Prathapani),
skite grains, whereas two-step deposition method gives uniform
mallick@iitb.ac.in (S. Mallick). grains with less pinholes. For the potential development of PSC

http://dx.doi.org/10.1016/j.apmt.2017.01.009
2352-9407/© 2017 Elsevier Ltd. All rights reserved.
S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119 113

technology, simplification of processing conditions is needed. the TTIP addition, the temperature of the oil bath was raised to
Therefore, we have processed our PSC devices in the ambient air, 80 ◦ C and kept so for 12 h. In the initial stage with the addition
which may lead to non-uniform perovskite films. of TTIP, the solution turned milky white in color due to hydroly-
When the perovskite films contain pinholes, the problem arises sis, and after few hours with constant stirring, the entire solution
that HTM can leak through these pinholes and reach c-TiO2 via converted into transparent white with complete formation of TiO2
mp-TiO2 . If the c-TiO2 is not conformal on FTO, HTM touches with colloid. After this step, the solution was cooled down to room tem-
it creating shunt paths in device causing recombinations. There- perature and stored. A very slow addition of TTIP into the HNO3
fore, it is critical to have conformal c-TiO2 on FTO. To deal with stock solution is critical to control the particle size of TiO2 colloid.
non-uniformly grown perovskite films and to realize the possi- As-formed colloids are stable over a period exceeding a year at room
ble PSC working devices, we have synthesized TiO2 -based colloidal temperature.
nanoparticles (NP) for the compact layer (c-TiO2 ) preparation and
fabricated the PSCs.
2.3. Synthesis of precursors for perovskite materials
In literature, the use of TiO2 compact layer for dye-sensitized
and perovskite-sensitized solar cells has been emphasized as effec-
The precursor material methylammonium iodide (CH3 NH3 I –
tive hole-blocking material [10,15]. Several methods have been
MAI) was synthesized by reacting equimolar ratio of methyl amine
reported to prepare c-TiO2 , but each of them has its own limitations.
and Hydroiodic acid in a round bottom flask at temperature 5–10 ◦ C
Recently, Kavan et al. evaluated the blocking nature of c-TiO2 grown
in an ice bath. The reacted solution was heated in open air at
by different methods like electrochemical, thermal oxidation, spray
50 ◦ C for 30 min, which yielded a white color solid product. This
pyrolysis and atomic layer deposition [16,17]. Other solution-based
product was washed with anhydrous diethyl ether three times
methods include chemical bath deposition by using TiCl4 and direct
and dissolved in dry ethanol. By using rotary evaporator main-
deposition of titanium (IV) isopropoxide (TTIP) solution in acetic
tained at 50 ◦ C, ethanol was taken out which lead to the formation
acid. These solution-based methods are also not viable, as TiCl4
of white crystals. These white crystals are of MAI as confirmed
vigorously reacts with moisture (while handling in ambient air)
from XRD study (Supplementary Fig. S1). This obtained product
or water and releases toxic hydrochloric acid gas [18]. TTIP-based
MAI was kept in a vacuum oven for 24 h and finally stored in
solutions form agglomerates on substrate during coating process on
nitrogen-filled glove box [5–7]. For one-step deposition method,
reaction with moisture in air. Hence, an alternative simple solution
the perovskite precursor solution was prepared by mixing CH3 NH3 I
route is needed.
and PbCl2 in 3:1 molar ratio in DMF to prepare 40 wt% of perovskite
Here we report one-step direct synthesis of colloidal TiO2
(CH3 NH3 PbI3−x Clx ) and stirred for 10 h at a constant temperature
nanoparticles with some modification to the earlier reported
of 70 ◦ C and obtained a transparent yellow solution [5–8]. For two-
method [19]. The current route of colloidal formation is water-
step sequential deposition method, 1 M PbI2 solution in DMF was
based environmental friendly method, which is non-hazardous,
prepared to coat on to the mesoporous TiO2 layers. The concentra-
amenable for large-scale production, inexpensive and can be han-
tion of MAI in 2-propanol was 8 mg/ml.
dled in ambient condition without creating any agglomerates.
These TiO2 colloids are stable for a period exceeding a year without
forming any agglomerates or precipitates and are suitable for dip-, 2.4. Perovskite solar cells fabrication
spin- and spray-coating techniques to form thin and compact TiO2
films. By using these compact layers, perovskite solar cells have The fabrication procedure for PSC devices involves several steps.
been fabricated in ambient conditions with various active areas First, FTO substrates were cleaned with soap solution, deionized
and realized considerable photo-conversion efficiencies (PCE) with water and ethanol by keeping in ultrasonic water bath for 15 min
better stability. at each step. Then these FTO substrates were heated to 450 ◦ C
for 30 min to remove the residual organics. The cleaned FTO sub-
strates were coated with colloidal TiO2 by spin-coating method and
2. Experimental details
heated to 500 ◦ C for 30 min to form c-TiO2 . For mp-TiO2 forma-
tion, slurry was prepared by using Dyesol 18 NR-T paste mixed
2.1. Materials
in ethanol in 1:4 weight ratios (TiO2 to ethanol). As-formed TiO2
slurry is spin coated on top of c-TiO2 and sintered at 500 ◦ C for
Chemicals titanium (IV) isopropoxide (TTIP) (97%), lead chlo-
20 min. For one-step deposition, at room temperature, the above-
ride (PbCl2 , 98%), lead iodide (PbI2 , 99%), hydroiodic acid
prepared FTO/c-TiO2 /mp-TiO2 substrates were spin coated with
(HI, 57 wt% in water), chlorobenzene (anhydrous 99.8%), N,N-
previously made perovskite solution in ambient air at 40 ± 5% RH
dimethylformamide (DMF, anhydrous 99.8%), and acetonitrile
and annealed at 100 ◦ C in nitrogen environment for perovskite
(anhydrous 99.8%) were bought from Sigma–Aldrich and used
formation. The spin-coating parameters for all above processes
as received. The other chemicals like HNO3 (65% assay) was
used were 5000 rpm for 30 s. For two-step sequential deposition
bought from Merck Millipore, Methylamine (40% in methanol)
method, first 1 M lead iodide solution was coated on to FTO/c-
from TCI chemicals and bis(trifluoromethane)sulfonamide lithium
TiO2 /mp-TiO2 substrates and heated to 100 ◦ C for 20 min and
(LiTFSI, 98%) salt from Alfa Aesar. TiO2 paste 18NR-T grade,
cooled down to room temperature. Then, these substrates were
fluorine-doped tin oxide on glass (FTO) (sheet resistance
dipped in MAI solution at room temperature for 1 min to com-
7 /sq) and 2,2 ,7,7 -tetrakis(N,N-di-p-methoxyphenyl-amine)-
pletely convert PbI2 into perovskite (CH3 NH3 PbI3 ) and annealed
9,9 -spirobifluorene (spiro-MeOTAD) were procured from Dyesol.
for 20 s at 70 ◦ C in ambient conditions. After annealing of perovskite
materials in both the methods, HTM was coated (in ambient condi-
2.2. Synthesis of TiO2 colloids tions) with the same spinning parameters as that of other material.
The formulation for HTM used was 80 mg/ml of Spiro-MeOTAD in
To synthesize TiO2 colloids, first, a stock solution of 0.1 M HNO3 chlorobenzene with the additives 17.5 ␮l of LiTFSI (from the stock
in distilled water was prepared. From the stock solution, 72 g solution of LiTFSI 520 mg/ml in acetonitrile) and 29.5 ␮l of tert-
of 0.1 M HNO3 was taken in a 150 ml round bottom flask and butyl pyridine [15]. Finally, metal Au was evaporated on to HTM
immersed in an oil bath. 12 g of TTIP was added drop by drop to make the back contact; then, I–V measurements were carried
into the stock solution while stirring at room temperature. After out.
114 S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119

Fig. 1. TEM images for TiO2 colloid: (a) high-resolution image, which confirms ∼5 nm size, (b) the uniform size distribution of colloid NP over the TEM grid, (c) XRD
spectrogram showing TiO2 phase information as anatase, and (d) bottle containing synthesized TiO2 colloid.

2.5. Characterization information XRD studies were carried out for the colloids by drop
casting on Plain glass and the results are shown in Fig. 1(c),
TiO2 colloids crystalline phase information was confirmed by which confirm that these colloids are single crystals of anatase
X-ray diffraction with X-ray diffractometer model PANalytical – phase having the diffraction peaks matching with the standard
MRDTM . TiO2 particle size distribution and its conformal thin crystallographic data, with JCPDS card no. 21-1272. The observed
film formation on FTO were studied by using TEM- model JEOL- peaks in the XRD pattern are very broad indicating qualitatively
JEM 2100F and FEG-SEM model JEOL JSM-7600F, respectively. the presence of very small nanoparticles. By considering high-
Absorption/transmission measurements were carried out by using est intensity peak (101), the crystallite size is calculated by using
UV–Visible–NIR spectrophotometer model Jasco V-650; for trans- Debye–Scherrer’s equation (D = 0.9/ˇ cos ; D – particle size,  –
mission measurements, integrating sphere was employed with wavelength of X-rays, ˇ – full width at half maxima of correspond-
BaSO4 as standard reference to account for reflection losses. The ing peak and  – Bragg angle), and it is found that the crystallite
current density–voltage characteristics for the fabricated PSCs were size is of ∼5 nm, which is in agreement with the observed particle
obtained by using standard solar simulator with 150 W Xenon size in TEM results. A photograph of the prepared colloids is shown
arc-lamp model NEW PORT-92251A-1000 attached with keithley- in Fig. 1(d).
2420 source meter under one sun (AM 1.5, 100 mW/cm2 ). Cyclic Size-dependent quantization effects for extremely small TiO2
voltammetry measurements were carried out by using potentio- particles were reported earlier [20–22]. To test the possible quan-
stat/galvanostat model PGSTAT 302 N Metrohm Autolab attached tum confinement effects for these colloids, absorbance studies were
with NOVA software. carried out. Fig. 2 shows the absorbance response of the colloids
with an absorption edge at 340 nm corresponding to band gap of
3.64 eV. However, from the literature it is understood that TiO2 in
3. Results and discussion anatase phase has an indirect band gap of 3.2 eV [23–25]. To con-
firm the actual band gap of these colloids, Tauc plots were made
3.1. TiO2 colloid for direct [(Ah)2 vs. h; A – absorbance, h – plank’s constant and
 – frequency of light] and indirect [(Ah)1/2 vs. h] allowed transi-
Fig. 1(a) shows the high-resolution transmission electron tions as shown in Fig. 2 inset. The abscissa of the intercepts for these
microscopy image confirming the ∼5 nm size of as-prepared two curves on energy (h) axis gives the band gaps for direct and
TiO2 nanoparticles; their uniform size distribution is shown by indirect transitions, respectively. On observing these band gap esti-
low-magnification TEM image in Fig. 1(b). For crystallography mations with the help of these allowed transition plots (as shown in
S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119 115

3.2. TiO2 compact layer from colloid

For solar cell applications, transmittance of substrates is con-


sidered to be a paramount factor to pass light through them and to
check that preliminarily sintered TiO2 compact films of ∼145 nm
and 40 nm average thicknesses were made by dip- and spin-coating
of TiO2 colloids on FTO respectively (Fig. 3(a) and (b), respectively).
In general, the surface roughness of FTO is around 20 nm (Fig. S2);
the prepared TiO2 colloids with ∼5 nm size can effectively fill the
hillocks on FTO and form a continuous dense compact layer with
full coverage as shown in SEM top view images in Fig. 3(c) and (d).
Apparently, it is evident from Fig. 3(c) and (d) that there are no vis-
ible cracks or pinholes in c-TiO2 films. These FTO/c-TiO2 substrates
were subjected to transmittance test over 300–800 nm of incident
light, which is from far ultraviolet to near infrared region as shown
in Fig. 4. In this regime, FTO has a transmittance of 80–85% and for
spin- and dip-coated FTO/c-TiO2 , it is of about 75–80%, which is
considerable to use for solar cell fabrication.

Fig. 2. Absorbance of TiO2 colloids (inset – Tauc plots).


3.3. Cyclic voltammetry

Kavan et al. applied the cyclic voltammetry (CV) method as


Fig. 2 inset), it is clear that in ‘direct transition curve,’ the intercept a potential technique to evaluate the blocking nature of c-TiO2
is at 3.64 eV that corresponds to direct inter band gap, and for ‘indi- [16,17]. On adopting this method the blocking nature of spin-
rect transition curve,’ the intercept is at 3.23 eV that corresponds to coated c-TiO2 films on FTO was evaluated and compared with bare
‘indirect band gap’. These observed energy band gap results are in FTO. As Spiro-OMeTAD is the HTM used in perovskite solar cells,
good agreement with the earlier reports [23,24]. Since the lowest a redox probe based on the same can better justify the blocking
value is considered to be the band gap, it can be concluded that nature of c-TiO2 . Therefore, the composition of electrolyte chosen
for the ∼5 nm particle size TiO2 colloids, the band gap is of indirect was 1 mM Spiro-MeOTAD in Dichloromethane with the addition of
type with 3.23 eV and no quantum confinement effects have been 0.3 M Tetrabutylammonium hexafluorophosphate [16,17]. Three
seen. Therefore, the colloids are considered to be nanoparticles of electrode system was used for CV, where Ag/AgCl was reference, Pt
∼ 5 nm sizes and are not the quantum dots. was auxiliary and c-TiO2 or FTO was working electrode. Fig. 5 shows

Fig. 3. FEG-SEM images: (a) cross-section view showing sintered dense c-TiO2 film on FTO formed by dip-coating with minimum thickness 114 nm and a maximum of
174 nm (average of ∼145 nm), (b) cross-section view of sintered c-TiO2 on FTO formed by spin-coating method with 40 nm thickness, (c) high-resolution top-view showing
full coverage of FTO surface hillocks by c-TiO2 , and (d) large area top-view image showing absence of visible cracks in c-TiO2 .
116 S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119

literature we note that the factors affecting the uniformity of the


perovskite films are wettability and thus nucleation and growth
[13], the presence of humidity [13,29], type of precursors taken
and their concentration [14,30,31], the annealing temperatures and
time of annealing [27,30–32], formation of intermediate products
and their volatilization [30,33], etc. In one-step method, during fast-
spinning process, a fraction of perovskite phase was formed even at
room temperature along with some amount of residual PbI2 , which
is confirmed from XRD studies as shown in Fig. 6(d), and labeled as
MALI-BA (before annealing). After annealing process at 100 ◦ C for
40 min, a complete perovskite phase was achieved; this is indicated
as MALI-AA in Fig. 6(d). The perovskite phase confirmation for the
films made by sequential deposition method is shown in Fig. 6(d)
with the label ‘Two-step MALI’.
The entire device architecture with its energy band diagram is
shown in Fig. 7(a) and (b), respectively. Since, the estimated energy
gap for colloidal TiO2 NP is of 3.2 eV, which is in accordance with
the energy gap for bulk TiO2 [23–25], the same band gap is consid-
Fig. 4. Transmittance of c-TiO2 formed on FTO by dip- and spin-coating method in ered for both c-TiO2 and mp-TiO2 for Fig. 7(b). In a typical working
comparison with FTO.
device, when light shines, electron–hole pairs are generated (elec-
tron being in conduction band and holes being in valence band of
perovskite) in perovskite materials. Instantly, the electrons move
from perovskite conduction band to ETM, i.e. TiO2 conduction band,
and reach FTO. In the same way, holes move from valence band of
perovskite to the highest occupied molecular orbit (HOMO level) of
HTM, i.e. to Spiro-MeOTAD, and then reach back electrode Au com-
pleting the circuit. The entire mechanism is depicted in Fig. 7(b); the
energy-level values for the various materials are considered from
the literature [7,34].
Fig. 8 shows the current density vs. voltage (J–V) characteristics
for the fabricated best PSC devices processed in ambient condi-
tions. The highest PCE () achieved through one-step process was
8.08% for an active area of 0.07 cm2 and the devices with an active
area of 0.23 cm2 achieved a best efficiency of 7.52% as shown in
Fig. 8(a) and (b), respectively. For 0.64 cm2 , the observed efficiency
was very low. A batch of three cells for each active area has been
prepared and their mean PCE values are tabulated along with the
best cell parameters in Table 1. In two-step method, the perov-
Fig. 5. Cyclic voltammograms corresponding to bare FTO and 40-nm thick C- skite morphology and thickness are decided by the PbI2 layer. The
TiO2 on FTO (FTO/c-TiO2 ). The electrolyte composition: 1 mM Spiro-OMeTAD thickness of the PbI2 layer can be varied more efficiently with the
in Dichloromethane + 0.3 M Tetrabutylammonium hexafluorophosphate. Scan rate deposition parameters, and later it can be converted into perov-
50 mV/s.
skite by dipping in MAI solution. PSCs were made with 600-nm
thick perovskite films (Fig. S3) and tested for their intrinsic sta-
the cyclic voltammograms for c-TiO2 and bare FTO. The cracks or bility against ambient moisture by storing them in a shelf in dark
pinholes on sintered c-TiO2 can be monitored by observing the without any encapsulation. These thick PSCs show stable perfor-
occurrence of anodic current due to oxidation of Spiro-OMeTAD mance with only 3% degradation up to 500 h as shown in Fig. 8(c).
at exposed FTO regions. From these voltammograms, it is clear Complete test details with respect to aging time are presented in
that the observed anodic current density is much smaller for FTO/ Table 2. It is to be noted that the PSCs fabricated by both the meth-
c-TiO2 on compared with bare FTO proving the blocking nature of ods have shown considerable hysteresis (Fig. S4). Fig. 8(d) shows
c-TiO2 . the photographs of 0.23 and 0.64 cm2 PSCs.
The reason for observed low efficiencies can be understood in
3.4. PSCs fabrication and their performance evaluation the following way. The ability of charge collection, recombina-
tions and their effect on quasi-Fermi-level movement in ETL are
The critical step in the device fabrication is the formation of considered to be the major reasons for the VOC loss [30,35,36].
crystalline perovskite film with good continuity on to the mp-TiO2 . In one-step method, the poorly covered island kind of perovskite
Therefore, before the device fabrication, the perovskite film for- has large areas of exposed TiO2 in contact with HTM causing the
mation studies were carried out in view of perovskite infiltration recombination of holes in the perovskite with the electrons in TiO2 ,
and surface coverage. Fig. 6(a) shows the cross-section view of thus reducing the electron charge density of TiO2 . This reduction
the infiltrated perovskite onto the mp-TiO2 (by one-step method), in charge density in TiO2 can cause the lowering of quasi-Fermi
which is about 440 nm. From the literature it is understood that energy-level, thus leading to the reduction in the measured VOC .
the perovskite materials with the thickness of 440 nm can have Since in two-step deposition method the perovskite coverage is
good optical density [26]. In accordance with the earlier literature, comparatively better than the one-step, comparatively better VOC
the microstructure of as-formed one-step film was non-uniform is achieved. In one-step method, due to the considerable thickness
with perovskite islands on mp-TiO2 as shown in Fig. 6(b) [13,27,28]. (400 nm) of the perovskite active layer, it can have large optical
But, with the two-step sequential deposited method, we obtained density leading to the higher JSC values (for small areas), whereas
comparatively better uniform films as shown in Fig. 6(c). From the in two-step method, due to large thickness (600 nm), some carriers
S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119 117

Fig. 6. (a and b) FEG-SEM images showing cross-section and surface morphology of infiltrated perovskite into TiO2 at RH 40 ± 5%, respectively via one-step method, (c)
FEG-SEM image showing top view of perovskite film formed by two-step method, and (d) XRD of perovskite films in one and two-step process.

Fig. 7. (a and b) Architecture and energy-level diagram for the various layers for PSC device, respectively.

Table 1
CH3 NH3 PbI3−x Clx PSCs performance parameters form J–V measurements by one-step method.

Cell area (cm2 ) Category (one-step) JSC (mA/cm2 ) VOC (V) FF  (%)

0.07 Best (mean) 18.52 (15.77 ± 2.47) 0.713 (0.703 ± 0.01) 0.61 (0.59 ± 0.05) 8.08 (6.68 ± 1.51)
0.23 Best (mean) 18.07 (15.30 ± 2.49) 0.743 (0.734 ± 0.01) 0.56 (0.55 ± 0.03) 7.52 (6.31 ± 1.33)
0.64 Best (mean) 4.40 (4.19 ± 0.66) 0.589 (0.612 ± 0.04) 0.68 (0.54 ± 0.09) 1.76 (1.38 ± 0.27)

Table 2
CH3 NH3 PbI3 PSC performance parameters prepared by two-step sequential deposition method and performance variation with the aging time (cell active area is 0.64 cm2 ).

Time (h) JSC (mA/cm2 ) VOC (V) FF  (%) Degradation (%)

0 12.4 0.82 0.45 4.61 –


500 10.46 0.85 0.5 4.47 3.00
1200 7.36 0.8 0.47 2.75 40.34
118 S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119

Fig. 8. J–V characteristic curves for the best PSC devices (a) 0.07 cm2 active area with PCE of 8.08% (one-step), (b) 0.23 cm2 active area with PCE of 7.52% (one step), (c)
0.64 cm2 active area with PCE of 4.61% and its performance degradation with aging time (0 h represents immediate measurement after the PSC fabrication within an hour),
and (d) figure showing pictures of fabricated large active area cells by one- and two-step methods.

may be lost in recombination leading to comparatively low JSC . As Acknowledgements


observed from the FEG-SEM images, the grain size of the perov-
skite in sequential deposition method is smaller; therefore, there The authors sincerely acknowledge Devices & Interfaces Lab-
might be grain boundary contributions for recombinations, which oratory, IIT Bombay for metallization and device measurement
may further affect VOC and JSC [37,38]. Performance parameters for facilities, and Dr. Chris Moore – Dyesol UK Limited for provid-
a series of devices fabricated under ambient conditions at room ing Spiro-MeOTAD and TiO2 Materials. The financial and analytical
temperature are given in Table S1. facilities provided by SERIIUS (Solar Energy Research Institute for
India and the United States), NCPRE and SAIF at IIT Bombay and
central funding agencies MNRE and DST are greatly acknowledged.
4. Conclusion
Appendix A. Supplementary data
In summary, TiO2 colloids were synthesized and used to form
c-TiO2 on FTO. The blocking nature of c-TiO2 was confirmed by Supplementary data associated with this article can be found, in
the method of cyclic voltammetry using Spiro-MeOTAD as redox the online version, at doi:10.1016/j.apmt.2017.01.009.
probe. Non-uniform perovskite film morphologies were observed
with ambient processing. The method of TiO2 colloid-based com- References
pact layer formation was emphasized as effective hole-blocking
layers for PSCs processed under ambient conditions with observed [1] B. Suarez, V. Gonzalez-Pedro, T.S. Ripolles, R.S. Sanchez, L. Otero, I. Mora-Sero,
photo-conversion efficiency of 8.08% for small active area 0.07 cm2 Recombination study of combined halides (Cl, Br, I) perovskite solar cells, J.
Phys. Chem. Lett. 5 (2014) 1628–1635.
and 7.52% for relatively large active area of 0.23 cm2 via one- [2] T.B. Song, Q. Chen, H. Zhou, C. Jiang, H.-H. Wang, Y. (Michael) Yang, Y. Liu, J.
step method. In two-step sequential method, comparatively better You, Y. Yang, Perovskite solar cells: film formation and properties, J. Mater.
perovskite coverage was observed, and therefore PSCs with com- Chem. A 3 (2015) 9032–9050.
[3] G. Xing, N. Mathews, S. Sun, S.S. Lim, Y.M. Lam, S. Mhaisalkar, T.C. Sum,
paratively large active areas of 0.64 cm2 show PCE of 4.61% with Long-range balanced electron- and hole-transport lengths in
considerable stability in ambient air. organic–inorganic CH3 NH3 PbI3 , Science 342 (2013) 344–347.
S. Prathapani et al. / Applied Materials Today 7 (2017) 112–119 119

[4] S.D. Stranks, G.E. Eperon, G. Grancini, C. Menelaou, M.J.P. Alcocer, T. Leijtens, [22] L. Kavan, T. Stoto, M. Grätzel, D. Fitzmaurice, V. Shklover, Quantum size
L.M. Herz, A. Petrozza, H.J. Snaith, Electron–hole diffusion lengths exceeding effects in nanocrystalline semiconducting TiO2 layers prepared by anodic
1 micrometer in an organometal trihalide perovskite absorber, Science 342 oxidative hydrolysis of TICl3 , J. Phys. Chem. 97 (1993) 9493–9498.
(2013) 341–344. [23] D.O. Scanlon, C.W. Dunnill, J. Buckeridge, S.A. Shevlin, A.J. Logsdail, S.M.
[5] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, Organometal halide perovskites Woodley, C.R.A. Catlow, M.J. Powell, R.G. Palgrave, I.P. Parkin, G.W. Watson,
as visible-light sensitizers for photovoltaic cells, J. Am. Chem. Soc. 131 (2011) T.W. Keal, P. Sherwood, A. Walsh, A.A. Sokol, Band alignment of rutile and
6050–6051. anatase TiO2 , Nat. Mater. 12 (2013) 798–801.
[6] J.H. Im, C.R. Lee, J.W. Lee, S.W. Park, N.G. Park, 6.5% efficient perovskite [24] N. Serpone, D. Lawless, R. Khairutdinov, Size effects on the photophysical
quantum-dot-sensitized solar cell, Nanoscale 3 (2011) 4088–4093. properties of colloidal anatase TiO2 particles: size quantization or direct
[7] H.S. Kim, C.R. Lee, J.H. Im, K.B. Lee, T. Moehl, A. Marchioro, S.J. Moon, R. transitions in this indirect semiconductor? J. Phys. Chem. 99 (1995)
Humphry-Baker, J.H. Yum, J.E. Moser, M. Grätzel, N.G. Park, Lead iodide 16646–16654.
perovskite sensitized all-solid-state submicron thin film mesoscopic solar cell [25] I. Chung, B. Lee, J. He, R.P.H. Chang, M.G. Kanatzidis, All-solid-state
with efficiency exceeding 9%, Sci. Rep. 2 (2012) 591, 1–7. dye-sensitized solar cells with high efficiency, Nature 485 (2012) 486–489.
[8] H. Zhou, Q. Chen, G. Li, S. Luo, T.B. Song, H.S. Duan, Z. Hong, J. You, Y. Liu, Y. [26] T. Leijtens, B. Lauber, G.E. Eperon, S.D. Stranks, H.J. Snaith, The importance of
Yang, Interface engineering of highly efficient perovskite solar cells, Science perovskite pore filling in organometal mixed halide sensitized TiO2 -based
345 (2014) 542–546. solar cells, J. Phys. Chem. Lett. 5 (2014) 1096–1102.
[9] http://www.nrel.gov/ncpv/images/efficiency chart.jpg (accessed on 03.06.16). [27] B. Conings, A. Babayigit, T. Vangerven, J. D’Haen, J. Manca, H.-G. Boyen, The
[10] A. Yella, L.P. Heiniger, P. Gao, M.K. Nazeeruddin, M. Grätzel, Nanocrystalline impact of precursor water content on solution processed organometal halide
rutile electron extraction layer enables low-temperature solution processed perovskite films and solar cells, J. Mater. Chem. A 3 (2015) 19123–19128.
perovskite photovoltaics with 13.7% efficiency, Nano Lett. 14 (2014) 2591. [28] G.E. Eperon, S.N. Habisreutinger, T. Leijtens, B.J. Bruijnaers, J.J. van Franeker,
[11] J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M.K. Nazeeruddi, D.W. deQuilettes, S. Pathak, R.J. Sutton, G. Grancini, D.S. Ginger, R.A.J. Janssen,
M. Grätzel, Sequential deposition as a route to high-performance A. Petrozza, H.J. Snaith, The importance of moisture in hybrid lead halide
perovskite-sensitized solar cells, Nature 499 (2013) 316–319. perovskite thin film fabrication, ACS Nano 9 (2015) 9380–9393.
[12] M.M. Lee, J. Teuscher, T. Miyasaka, T.N. Murakami, H.J. Snaith, Efficient hybrid [29] J.A. Christians, P.A. Miranda Herrera, P.V. Kamat, Transformation of the excited
solar cells based on meso-superstructured organometal halide perovskites, state and photovoltaic efficiency of CH3 NH3 PbI3 perovskite upon controlled
Science 338 (2012) 643–647. exposure to humidified air, J. Am. Chem. Soc. 137 (2015) 1530–1538.
[13] H. Gao, C. Bao, F. Li, T. Yu, J. Yang, W. Zhu, X. Zhou, G. Fu, Z. Zou, Nucleation [30] F.K. Aldibaja, L. Badia, E. Mas-Marza, R.S. Sanchez, E.M. Barea, I. Mora-Sero,
and crystal growth of organic–inorganic lead halide perovskites under Effect of different lead precursors on perovskite solar cell performance and
different relative humidity, ACS Appl. Mater. Interfaces 7 (2015) 9110–9117. stability, J. Mater. Chem. A 3 (2015) 9194–9200.
[14] B. Philippe, B.-W. Park, R. Lindblad, J. Oscarsson, S. Ahmadi, E.M.J. Johansson, [31] D. Khatiwada, S. Venkatesan, N. Adhikari, A. Dubey, A.F. Mitul, L. Mohammad,
H. Rensmo, Chemical and electronic structure characterization of lead halide A. Iefanova, S.B. Darling, Q. Qiao, Efficient perovskite solar cells by
perovskites and stability behavior under different exposures – a temperature control in single and mixed halide precursor solutions and films,
photoelectron spectroscopy investigation, Chem. Mater. 27 (2015) J. Phys. Chem. C 119 (2015) 25747–25753.
1720–1731. [32] A.T. Mallajosyula, K. Fernando, S. Bhatt, A. Singh, B.W. Alphenaar, J.-C.
[15] A.K. Chandiran, A. Yella, M. Stefik, L.-P. Heiniger, P. Comte, M.K. Nazeeruddin, Blancon, W. Nie, G. Gupta, A.D. Mohite, Large-area hysteresis-free perovskite
M. Grätzel, Low-temperature crystalline titanium dioxide by atomic layer solar cells via temperature controlled doctor blading under ambient
deposition for dye-sensitized solar cells, ACS Appl. Mater. Interfaces 5 (2013) environment, Appl. Mater. Today 3 (2016) 96–102.
3487. [33] B. Wang, K.Y. Wong, X. Xiao, T. Chen, Elucidating the reaction pathways in the
[16] L. Kavan, M. Zukalova, O. Vik, D. Havlicek, Sol–gel titanium dioxide blocking synthesis of organolead trihalide perovskite for high-performance solar cells,
layers for dye-sensitized solar cells: electrochemical characterization, Chem. Sci. Rep. 5 (2015) 1–11, 10557.
Phys. Chem. 15 (2014) 1056–1061. [34] Y. Song, S. Lv, X. Liu, X. Li, S. Wang, H. Wei, D. Li, Y. Xiao, Q. Meng, Energy level
[17] L. Kavan, N. Tétreault, T. Moehl, M. Grätzel, Electrochemical characterization tuning of TPB-based hole-transporting materials for highly efficient
of TiO2 blocking layers for dye-sensitized solar cells, J. Phys. Chem. C 118 perovskite solar cells, Chem. Commun. 50 (2014) 15239–15242.
(2014) 16408–16418. [35] F. Fabregat-Santiago, J. Bisquert, L. Cevey, P. Chen, M. Wang, S.M. Zakeeruddin,
[18] P.K. Roy, A. Bhatt, C. Rajagopal, Quantitative risk assessment for accidental M. Grätzel, Electron transport and recombination in solid-state dye solar cell
release of titanium tetrachloride in a titanium sponge production plant, J. with Spiro-OMeTAD as hole conductor, J. Am. Chem. Soc. 131 (2009) 558–562.
Hazard. Mater. A102 (2003) 167–186. [36] H.J. Snaith, Estimating the maximum attainable efficiency in dye-sensitized
[19] W. Choi, A. Termin, M.R. Hoffmann, The role of metal ion dopants in solar cells, Adv. Funct. Mater. 20 (2010) 13–19.
quantum-sized TiO2 : correlation between photoreactivity and charge carrier [37] R. Long, J. Liu, O.V. Prezhdo, Unravelling the effects of grain boundary and
recombination dynamics, J. Phys. Chem. 98 (1994) 13669–13679. chemical doping on electron–hole recombination in CH3 NH3 PbI3 perovskite
[20] M. Anpo, T. Shima, S. Kodama, Y. Kubokawa, Photocatalytic hydrogenation of by time domain atomistic simulation, J. Am. Chem. Soc. 138 (2016)
CH3 CCH with H2 O on small-particle TiO2 : size quantization effects and 3884–3890.
reaction intermediates, J. Phys. Chem. 91 (1987) 4310–4317. [38] N. De Marco, H. Zhou, Q. Chen, P. Sun, Z. Liu, L. Meng, E.-P. Yao, Y. Liu, A.
[21] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, Preparation and Schiffer, Y. Yang, Guanidinium: a route to enhanced carrier lifetime and
characterization of quantum-size titanium dioxide, J. Phys. Chem. 92 (1988) open-circuit voltage in hybrid perovskite solar cells, Nano Lett. 16 (2016)
5201–5205. 1009–1016.

You might also like