Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

pubs.acs.

org/JPCC Article

Electron-Selective Layers for Dye-Sensitized Solar Cells Based on


TiO2 and SnO2
Ladislav Kavan,* Zuzana Vlckova Zivcova, Magda Zlamalova, Shaik M. Zakeeruddin,
and Michael Grätzel
Cite This: J. Phys. Chem. C 2020, 124, 6512−6521 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Titanium dioxide (anatase, rutile) and quasi-amorphous tin dioxide are
Downloaded via UNIV OF AGRI FAISALABAD on October 22, 2020 at 09:18:23 (UTC).

prepared on F-doped SnO2 in the form of dense thin films, which can serve as electron-
selective layers in perovskite solar cells and dye-sensitized solar cells (DSSCs). The present
study brings new data about electronic and electrochemical properties of these films at the
authentic conditions occurring in a dye-sensitized solar cell (DSSC). Hydrolysis of TiCl4
provides pure rutile TiO2 at low temperatures, but TiO2 (anatase) grows in these layers
upon calcination. In acetonitrile medium, the flat band potential of TiO2 (rutile) is more
negative than that of TiO2 (anatase). This is opposite ordering to that observed in aqueous
media. The energy of conduction band minimum of TiO2 (anatase) equals −4.15 ± 0.07 eV
at the conditions mimicking the DSSC’s environment. Electrochemical reductive doping of
SnO2 provides a material with the most negative flat band potential and the largest overpotential for the reduction of I3−,
Co(bpy)33+, and Cu(tmby)22+. Voltammetric screening of all the electrode materials in six different electrolyte solutions, relevant to
DSSC applications, gives salient information about the mediator type and effects of calcination and the addition of 4-tert-
butylpyridine. These data provide novel inputs for optimization of DSSCs and for perovskite photovoltaics, too.

1. INTRODUCTION Later on, a spray pyrolysis of titanium(IV) diisopropoxide-


The electron-selective layer (ESL; also called electron bis(acetylacetonate) was introduced 14 and further im-
transport, electron extraction, hole-blocking, or buffer layer) proved,15−17 while it became to be the most popular method
is a key component of both the dye-sensitized solar cell for the fabrication of ESL both in DSSCs and PSCs. The next
(DSSC)1−5 and the perovskite solar cell (PSC).6−8 ESL is important method of the ESL growth is the chemical bath
deposited, as the first layer, on the negative electrode terminal deposition from aqueous TiCl4 solutions.18,19 Whereas the
(typically made from F-doped SnO2, FTO, or indium−tin spray pyrolysis provides the anatase form of TiO2,15,17,19 the
oxide, ITO). Its primary role consists of blocking of any growth from TiCl4 can produce rutile18 or anatase18,20
depending on the synthetic details. Sometimes, TiO2 ESL is
electron leakage from FTO to the oxidized form of the
also grown by atomic layer deposition (ALD) providing
electrolyte mediator or holes in the hole-transport medium
perfectly compact films down to several nm thickness21 with
(including the perovskite itself) in the respective devices. ESL
promising performance in DSSC.22
thus enhances the open-circuit voltage, VOC, and fill factor of
The use of ESL is essential in DSSCs with organic dyes,1,4,23
both types of photovoltaics. Additional benefits of ESL follow
novel redox mediators (Co(III/II)19,24 or Cu(II/I)25), and in
from improved physical and electrical contacts between FTO
systems operating at weak illumination, for example, under
and the mesoporous oxide semiconductor (usually TiO2)9 and
ambient lighting.19,26 The ESL must be compact (pinhole-
enhanced optical transmittance.9,10 Specifically, in the I3−/I−-
free), which can be easily tested by cyclic voltammetry of
mediated DSSCs, this layer blocks the reaction of electrons
model redox probes, such as Fe(CN)63−/4−20,21,27 (note
with photo-generated I2−· radical anions, thus enhancing not
however, possible formation of Prussian blue, obscuring this
only the VOC but also the photocurrent.9
test28). Good ESLs must be rectifying, that is, it must exhibit
The traditional DSSCs with Ru sensitizers and the I3−/I−
solely the cathodic wave of ferricyanide reduction at potentials
redox mediator work reasonably well even in the absence of
ESL, which is ascribed to the fact that (i) the Ru dye itself can
effectively screen the electron transfer from FTO to the Received: December 24, 2019
electrolyte and (ii) the I3−/I− exchange current is very small on Revised: February 28, 2020
FTO.1−4 Yet, already the pioneering works11,12 noticed a Published: March 2, 2020
significant performance improvement if few monolayers of
dense TiO2 were electrochemically deposited13 over the FTO/
mesoporous TiO2 in the originally “ESL-free” cell architecture.

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.jpcc.9b11883


6512 J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

negative to the flat band potential but no anodic counterwave. 2.2. Characterization of the Films. The Raman spectra
Lee et al.19 compared the ESL made by spray pyrolysis with were acquired with the 514 nm laser excitation on a Labram
the ESL made from TiCl4. The latter was fabricated by the HR spectrometer (Horiba Jobin-Yvon) interfaced to an
usual synthetic protocol, that is, immersion of FTO into an Olympus microscope (objective 100×). The laser power at
aqueous 40 mM TiCl4 at 70 °C for 30 min.19,29 However, this the sample was 2 mW (the spot area was approx. 1 μm2).
layer provided poor FTO coverage by isolated TiO 2 Scanning electron microscopy (SEM) images were obtained by
particles19,30 translating into large proportion of the electro- a Hitachi S-4800 microscope. X-ray photoelectron spectros-
chemically detectable pinholes. Yet, the performance of this copy (XPS) was performed using an Omicron Nanotechnology
“non-ideal” ESL in a Co-mediated DSSC was still acceptable at instrument equipped with a monochromatized Al Kα source
solar illumination (1 to 0.1 sun) but not at weak ambient (1486.7 eV) and a hemispherical analyzer operating in
lighting (251−1001 lux) when solely the spray pyrolysis-made constant analyzer energy mode with a multichannel detector.
ESL gave efficient DSSCs.19 The binding energies were calibrated by evaporated gold (Au
The second ESL material, which is employed especially in
4f7/2 at 84.0 eV) and were referenced to the specimen Fermi
perovskite photovoltaics, is SnO2.31,32 It is grown usually by
level, which is set as zero. For the work function measurement
atomic layer deposition (ALD),27,32−35 sometimes also by
(by the secondary electron cut-off), the sample was biased at
spin-coating from SnCl2 or SnCl4 solutions.35,36 The spray
pyrolysis from butyltin trichloride was complicated by fluorine −6.3 V. The CasaXPS program was used for spectra analysis.
migration from FTO into the SnO2 at the temperature of The spectra were curve-fitted after subtraction of Shirley
pyrolysis, but this problem could be circumvented by replacing background using the Gaussian−Lorentzian line shape.
FTO with Al-doped ZnO.37 Quantitative surface analysis employed the relative sensitivity
Despite the great efforts invested so far into the ESL factors based on the Scofield’s cross sections.
development, a systematic comparison of various materials in 2.3. Electrochemical and Photoelectrochemical Prop-
different DSSC’s electrolyte solutions is still lacking. Hence, erties of the Films. Electrochemical experiments were carried
the central aim of this study is to acquire accurate data about out in a one-compartment cell using an Autolab Pgstat-30
electrochemical and electronic properties of the TiO2 and equipped with the FRA module (Metrohm). The counter-
SnO2 films, which already proved to be useful for ESL electrode was a platinum mesh, and the reference electrode
applications. These studies are carried out at the authentic was a non-aqueous Ag/AgCl (sat. LiCl in ethanol), which was
conditions occurring in dye-sensitized solar cells using interfaced by a bridge with 0.1 M lithium trifluoromethylsulfo-
acetonitrile electrolyte solutions and three different redox nylimide (LiTFSI) in acetonitrile. The potential was calibrated
mediators. We confirm here the formation of pure rutile in ex- using ferrocene (200 μL of 0.1 M acetonitrile solution per 10
TiCl4 layers, but quite unexpectedly, also the growth of anatase mL of electrolyte solution added and tested using the Pt
in these layers upon calcination. We find serendipitously that a working electrode; the potential of this primary standard (Fc+/
simple electrochemical treatment of ALD-SnO2 can downshift Fc) was fluctuating between 0.499 to 0.506 V vs. the non-
substantially the onset potential of dark current, which is aqueous Ag/AgCl). Cyclic voltammetry (in aqueous medium)
important for solar cells’ optimization. was performed in 0.5 M KCl with 0.5 mM K3[Fe(CN)6] and
0.5 mM K4[Fe(CN)6]; its pH was adjusted by HCl to 2.5. In
2. EXPERIMENTAL SECTION this case, the reference electrode was aqueous Ag/AgCl (sat.
2.1. Preparation of the TiO2 or SnO2 Layers. FTO glass KCl). The electrolyte solution was purged with Ar, and the
(NSG10, Nippon Sheet Glass, 10 ohm/sq) was cleaned measurement was carried out under Ar in a closed electro-
ultrasonically using a detergent solution (Deconex) followed chemical cell. Impedance spectra were measured in the
by water, ethanol, and acetone (10 min treatment in each frequency range from 100 kHz to 0.1 Hz (modulation
bath). Thin-film rutile, denoted TiO2(R), was prepared by a amplitude 10 mV) at varying potentials. Spectra were evaluated
chemical bath deposition method. The FTO substrate was using ZView (Scribner) software by fitting to a Randles-type
treated in 200 mM aqueous solution of TiCl4 at 70 °C for 1 h, circuit (Figure S1, Supporting Info). Here, RCT is the charge
producing an ∼25 nm-thick TiO2 film.18 It was subsequently transfer resistance, which is parallel to the constant phase
washed with water and ethanol and dried at 100 °C. The element (CPE) to account for non-ideal capacitive behav-
anatase thin film (ca. 30 nm), denoted TiO2(A), was deposited ior.21,27 The circuit is completed with a series resistance, RS,
by spray pyrolysis (at 450 °C) using oxygen as the carrier gas characterizing the ohmic resistance of electrodes, electrical
and a solution of 0.2 M titanium diisopropoxide-bis- contacts, and electrolyte solution. The Warburg impedance,
(acetylacetonate) and 0.4 M acetylacetone in ethanol.15,16 ZW, corresponds to the ionic transport in solution.
The quasi-amorphous SnO2 films (3−15 nm) were prepared as The impedance of CPE equals38
in ref 27 by atomic layer deposition (ALD) at 118 °C using
tetrakis-(dimethylamino)tin(IV) (99.99% Sn, Strem Chem- ZCPE = B−1(iω)−β (1)
icals, Inc.) and ozone using Savannah ALD 100 (Cambridge
Nanotech, Inc.) or R200 (Picosun, Finland). The redox with ω being the EIS frequency and B (admittance prefactor)
mediators, Co(bpy) 3 (TFSI) 3 , Co(bpy) 3 (TFSI) 2 , Cu- and β (exponent) are the frequency-independent parameters of
(tmby)2(TFSI)2 and Cu(tmby)2(TFSI) (bpy = 2,2′-bipyr- the CPE (0.8 ≤ β ≤ 1; experimental values were from 0.8 to
idine; tmby = 4,4,6,6-tetramethyl-2,2′-bipyridine; TFSI = 0.9). The capacitance, C, is calculated from38
trifluoromethylsulfonylimide) were purchased from Dyenamo,
Sweden. Table S1 (Supporting Info) summarizes the details (R CTB)1/ β
about electrolyte solutions. Other chemicals were from Aldrich C=
and used as received. R CT (2)

6513 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

3. RESULTS AND DISCUSSION anatase TiO2 (Figure 1; cf. also ref 17), which is insensitive to
3.1. ESL from TiO2. The chemical bath deposition of TiO2 calcination (450 °C, 1 h in air; the standard protocol of the
on FTO by the standard synthetic protocol (treatment in 40 fabrication of a TiO2 photoanode).
mM TiCl4 at 70 °C for 30 min19,29) gave a poorly blocking film The Raman analysis of ex-TiCl4 titania is less easy due to
as illustrated in Figure S2 (Supporting Info). Yella et al.18 low Raman intensities in this particular sample. However, the
reported that a simple enhancement of the TiCl4 concentration presence of rutile is confirmed in a film, which was grown as
(to 200 mM) gave a more uniform film with excellent the TiO2(R) sample, but with deliberately doubled reaction
performance in PSC. This observation is confirmed here by a time (Figure 1). Hence, we conclude, in accordance with Yella
blocking test (Figure S2), which illustrates (i) efficient et al.,18 that our TiO2(R) sample is a rutile thin film with
suppression of the ferrocyanide oxidation and (ii) fast perfect compactness useful for ESL fabrication. Interestingly,
reversible charging/discharging of chemical capacitance with TiO2(R), which was calcined at 450 °C for 1 h, exhibits a
a characteristic prepeak assigned to surface states.39 This film clearly resolved peak at 141 cm−1, which is assignable to the
(denoted TiO2(R)) was therefore chosen for all additional diagnostic line of anatase, (Eg(1); see Figure 1). Obviously, the
studies of the ex-TiCl4 layers. ex-TiCl4 layer is unique not only by the low-temperature
Raman spectroscopy is a useful technique for the phase growth of rutile TiO218 but also by its unexpected trans-
analysis of titania.40,41 Figure 1 shows that a blank FTO has formation to anatase upon calcination. (An alternative
explanation is that anatase crystallizes during calcination
from the originally “amorphous” titania present in the parent
TiO2(R)).
The fast reversible charging/discharging of chemical
capacitance in TiO2(R) (cf. Figure S2) is reproduced in
acetonitrile medium, too. Figure 2 shows the relevant cyclic
voltammograms. Furthermore, Figure 2 illustrates the effect of
4-tert-butylpyridine (TBP) in the electrolyte solution. This
base is ubiquitously used as the standard component of
virtually all electrolyte solutions in DSSCs.43 (However, some
concern exists about the Cu(II/I) mediator, which is affected
by the coordination with TBP44−46). In general, TBP
significantly enhances VOC, which is ascribed to the Fermi
level upshift and longer electron lifetime in the conduction
band of both TiO247 and SnO2.48 Our voltammogram in
Figure 2 confirms that the onset potential of the capacitive
charging is downshifted by ca. 0.2 V in the presence of TBP.
This finding matches qualitatively the observed downshift of
the flat band potential (see section 3.3). Furthermore,
interaction of TBP with titania blocks the surface states
(electron traps) in TiO2 and adsorption sites for Li+ and H+
(e.g., from trace humidity).3,43,47 Hence, the cathodic/anodic
branches of the cyclic voltammogram are more symmetrical in
the presence of TBP because the voltammetric features of
Figure 1. Raman spectra of the FTO-supported TiO2(R) electrode surface states and cation adsorption are effectively blocked by
(blue curve), blank FTO substrate (turquoise), and TiO2(A)
TBP (Figure 2).
electrode (black curve). Also shown is the spectrum of an identically
grown titania film [as for TiO2(R)] but with doubled reaction time (2 3.2. ESL from SnO2. In contrast to TiO2 films, which
h instead of 1 h, labeled “2hr”). The symbol HT denotes the exhibit well-developed reversible charging/discharging of
electrodes that were heat-treated at 450 °C for 1 h. Spectra of chemical capacitance at potentials near the flat band (see
reference rutile powder (blue dashed curve) and anatase powder section 3.1), thin films of ALD-made SnO2 behave differently.
(black dashed curve) are shown, too. The diagnostic lines for rutile Figure 3 shows that a purely capacitive (double-layer) charging
and anatase are labeled by full and dashed arrows, respectively. The occurs in the potential window from 0.5 V to ca. −1 V versus
spectra are offset for clarity, but the intensity scale is identical for all Fc+/Fc. The corresponding Helmholtz capacitance, CH, equals
spectra, except for TiO2(A) and TiO2(A)-HT, which were attenuated
by a factor of 10. JH
CH =
ν (3)
relatively weak Raman features (the main band being A1g mode (JH is the current density and ν is the scan rate). Equation 3
of SnO2 at 635 cm‑1 42). They do not interfere with the provides an estimate of the Helmholtz capacitance of 30 μF/
diagnostic bands of TiO2 neither in anatase nor in rutile forms. cm2. (This value would translate, for our compact 4 nm film,
Hence, the FTO-supported thin TiO 2 films can be into ca. 10 F/g, which is nearly identical to the specific
conveniently analyzed by Raman. This is a significant capacitance of SnO2 quantum dots in aqueous medium42). The
advantage over X-ray diffraction because the strong FTO capacitances are corrected by considering the average rough-
lines dominate the diffractograms of, for example, anatase ness factor of 1.3 for these films.27
made by spray pyrolysis15,19 or by ALD.22 This problem is In addition to capacitive effects, we observe irreversible
further augmented with TiO2 rutile thin films because the FTO cathodic charging at potentials negative to ca. −1 V, which is
(cassiterite structure) has a very similar diffraction pattern. The more pronounced in virgin films (dashed lines in Figure 3). If
TiO2(A) layer expectedly shows the Raman spectrum of pure this film was pretreated at −1.5 V for 2 min, then the net
6514 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 2. Cyclic voltammograms of a TiO2(R) thin-film electrode (made by hydrolysis of TiCl4). Electrolyte solution: (A) 0.1 M LiTFSI +
acetonitrile; (B) 0.1 M LiTFSI +0.5 M 4-tert-butylpyridine in acetonitrile. Scan rates of 100, 50, and 20 mV/s for the red, blue, and black curves,
respectively.

of SnO and SnO2 strongly overlap, and the deconvolution by


fitting of the overlapping lines is not accurate for the Sn2+
concentrations below ∼10%.
The mechanism of Li doping is still unclear. In contrast to
TiO2 (anatase or rutile, which exhibits simple interstitial Li
insertion), SnO2 accommodates lithium electrochemically by
conversion reactions producing various Sn2+- and Sn0-
containing species.51,52 These reactions proceed at potentials
smaller than ca. 1 V versus Li+/Li (which is approximately
−2.6 V vs. Fc+/Fc). At the potentials that are used for the
growth of SnO2(D) (≥ −1.5 V vs. Fc+/Fc), the conversion
reactions hardly occur at significant speed. Instead, milder
effects are at play, similar to those reported during the growth
of Li/SnO2 from SnCl2 and LiTFSI. In this case, Park et al.50
proposed that Li doping consists of the substitution of Sn4+ by
Li+. As the binding energies of our SnO2 and SnO2(D) are not
affected (±0.1 eV) and close to those of Li/SnO2,50 the
structural perturbations of the SnO2 lattice are small (if any) in
Figure 3. Cyclic voltammograms of a SnO2 thin-film (4 nm) electrode both cases of Li doping, electrochemical and chemical.50 In
in 0.1 M LiTFSI + acetonitrile, with a scan rate of 0.1 V/s. Dashed accordance with earlier works,27,32,53 we detect trace amounts
lines are for the virgin electrode, full lines are for the same electrode of non-carbonaceous impurities (N, Si) but no F 1s from the
that was treated at −1.5 V for 120 s. FTO substrate.
3.3. Characterization by Electrochemical Impedance
Spectroscopy. Electrochemical impedance spectra were
cathodic charge of 0.34 mC/cm2 was consumed irreversibly,
acquired to determine the flat band potentials (Efb) and the
and the voltammogram changed as illustrated in Figure 3. We
donor concentrations (ND). The spectra were fitted to an
denote this product as “doped” and symbolize it further
equivalent circuit with the constant phase element, which
SnO2(D).
accounts for non-ideal capacitive behavior of polycrystalline
The found cathodic charge of 0.34 mC/cm2 would
thin films21 (eq 2 and Figure S1). The values of Efb and ND
theoretically correspond to approximately 9.5% conversion of
were extracted from the Mott−Schottky equation21
our SnO2 film into SnO. Additional experiments with varying
ij 2 yzij
= jjj zzjjE − E − kBT yzzz + 1
film thicknesses (from 3 to 10 nm; data not shown) provided
j zzj e z{ C H 2
k eε0εrND {k
roughly similar conversion rates, but they dropped for thicker 1
2 fb
films (up to 50 nm), clearly indicating that the reductive C (4)
doping affects the film down to several nm underneath the
surface. Figures S3−S5 (Supporting Info) show supplementary e is the electron charge, ε0 is the permittivity of free space, εr is
details about reductive doping of SnO2. It can simply be the dielectric constant (εr values are 173, 55, and 10 for TiO2
carried out by repeated voltammetric scanning, too. In thicker (rutile), TiO2 (anatase), and SnO2, respectively21,27,54), E is
films, the doping reaction manifests itself by a peak near −1.2 the applied voltage, Efb is the flat band potential, kB is the
V (Figure S5). Boltzmann’s constant, and T is the temperature.21,27,55 The
X-ray photoelectron spectroscopy shows that doping of value of CH is approximately 30 μF/cm2 for SnO2 (cf. eq 3).
SnO2 causes incorporation of ca. 6 at % Li (Figure S6, The voltammetric determination of CH for TiO2 is less
Supporting Info). The presence of Sn2+ would manifest itself straightforward (cf. Figure 2), but the experimental Mott−
by shifting of the Sn 3d5/2 photoemission line to smaller Schottky plots allow estimation of roughly the same value from
binding energies (by about 0.7 eV).49 Our found binding impedances measured at potential negative to the flat band
energy of Sn 3d5/2 (486.9 eV) is nearly identical to the value (Figure 4).
reported for Li/SnO2 prepared by spin-coating from SnCl2 and Figure 4 shows the Mott−Schottky plots of our four
LiTFSI solution (486.8 eV).50 Nevertheless, the contributions representative electrodes in 0.1 M LiTFSI + acetonitrile
6515 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 4. Mott−Schottky plots for different electrode materials as specified in the annotation. The film thickness of SnO2 and SnO2(D) equals 4
nm. The electrolyte solution is 0.1 M LiTFSI + acetonitrile.

ij 2ε ε yz ij k Ty
W = jjj 0 r zzz jjjE − Efb − B zzzz
j eND z
medium. To explore these plots at conditions mimicking the 1/2 1/2

k { k e {
DSSC’s electrolyte solution, we used two other model media:
0.1 M LiTFSI + 0.5 M TBP and 0.1 M LiTFSI + 0.1 M LiI + (5)
0.05 M I2 + 0.5 M TBP. Figure S7 (Supporting Info) shows the Equation 5 predicts (for the 1 V band-bending and the
comparative data for a TiO2(A) electrode. The found values of experimentally found ND values) the widths (W) of 14, 25, and
Efb and ND are collected in Table S2 (Supporting Info). The 3 nm for TiO2(A), TiO2(R), and SnO2, respectively. However,
flat band potentials are also listed here in the absolute scale, as explained in ref 27, the electrochemically found ND values of
EF(fb). Under certain approximations (neglecting the potential SnO2 (Table S2) are overestimated; the Hall measurements on
drop in the Hemholtz layer, chemisorbed ions, oriented virtually identical materials gave ND values of the order of 1016
dipoles, etc.), the energies EF(fb) are equal to the Fermi level cm−3, which would put the W values of SnO2 to the 102 nm
positions and define the work functions (φ = −EF(fb)). Our level. Consequently, the depletion layer safely resides within
values are roughly comparable to the work functions that were our TiO2 films (∼25−30 nm), but in SnO2, it can extend into
measured by ultraviolet photoelectron spectroscopy (UPS) on the supporting FTO, too. Indeed, the 4 nm SnO2 film exhibits
the second slope (Figure 4) corresponding to ND ≈ 1.2 × 1021
analogously made samples, for example, TiO2(A) (3.65 eV56 or
cm−3, that is, the typical donor density of FTO.21,27
3.75−3.96 eV57), ALD-SnO2 (4.4653 or 4.20 eV32), and ALD- The found flat band potential of TiO2(A) in our “DSSC-
SnO2 calcined at 300 °C (4.52 eV32). Our own data from XPS like” electrolyte solution (containing TBP and I3−/I−) is
secondary electron edge (Figure S9, Supporting Info) are 3.9, comparable to the value reported by Cameron and Peter2 from
3.7, 4.2, and 4.1 eV (± 0.1 eV) for the samples TiO2(A), a Mott−Schottky analysis of their spray-pyrolyzed TiO2
TiO2(R), SnO2, and SnO2(D), respectively. blocking layers in thin-layer cells with a similar DSSC
Interestingly, the addition of I3−/I− to the solution has no electrolyte. Their value of Efb = −0.76 V versus I3−/I−58
effect on the flat band potential and very small effect on the would translate59 into approximately −1 V versus Fc+/Fc,
ND, which is within the experimental error. The addition of which is close to our value of −0.90 ± 0.06 V versus Fc+/Fc
TBP causes downshift of Efb (by ca. 0.1−0.2 V) but again a (Table S2). On the other hand, our flat band potentials are
negligible change of ND. Obviously, there is no reason that the larger than those measured spectroelectrochemically on similar
films in dry acetonitrile (−2.4 to −1.3 V vs. Fc/Fc+ depending
donor density in a semiconductor electrode should depend on
on Li+ concentration).60 The difference is explained (besides
the electrolyte solution, but the expected effect of TBP43−48 is very diverse experimental conditions in the reference
thus confirmed (cf. also Figure 1 and discussion thereof). studies2,60) by the collective influence of Li+ and humidity,
The applicability of the depletion layer model for our thin both upshifting the Efb values significantly.60 Notably, in
films is tested by the estimation of the width of the space- contrast to studies of thin-layer cells,2 our approach provides a
charge layer (W), assuming a parallel-plate model of the well-defined reference electrode and the three-electrode setup
capacitance for accurate electrochemical experiments.
6516 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

The electrochemical potential corresponding to the This result is in remarkable accordance with our observed 0.3
conduction band minimum (ECBM) is V downshift of Efb after heat treatment of ALD-SnO2.
To further advance this discussion, Figure S9 (Supporting
kBT NC* Info) shows the XPS spectra of our ESL materials in the
ECBM = Efb − ln valence band region. The valence band offset (referenced to
e ND (6)
the Fermi level position) equals 3.1, 2.9, 3.8, and 3.7 eV (± 0.1
where NC* is the effective density of conduction band states eV) for the samples TiO2(A), TiO2(R), SnO2, and SnO2(D),

ji 2πme*kBT zyz
respectively. These values are comparable to those reported by

2jjj zz
others32,34,57,65,68,69 (although they are also known to depend
j z
3/2

k {
NC* = strongly on the samples’ preparation history32,57,68). For the
h2 (7) known optical band gaps (Eg) of TiO2 anatase and rutile of 3.2
and 3.0 eV, respectively, and for ALD-SnO2 (Eg = 3.98 eV),32
me* is the effective mass of electron (for TiO2 anatase me* ≈ the corresponding offset of Fermi level from the CBM is quite
1.9m0 with m0 being the electron’s rest mass61). Assuming ND small (from ∼0.1 to 0.3 eV) in all our materials. This confirms
≈ (4.2 ± 1.2) × 1019 cm−3 (Table S2), the calculation gives the their significant n-doping (cf. Table S2 and eqs 6,7). The
energy of CB minimum equal to −4.15 ± 0.07 eV in the valence bands of SnO2 and SnO2(D) have very similar shapes,
absolute scale (with the work function of SHE being 4.44 eV). which is in accordance with the negligible doping-induced
We should note that the CBM position of TiO2 appears changes in the spectra of core levels (cf. Figure S6 and
frequently in the publications discussing energetics of DSSC discussion thereof).
but without any clear reference how this CBM was actually 3.4. Dark Current on ESLs with Different Redox
determined.3 Here, we provide the authentic value for TiO2 Mediators. Finally, we investigated our four generic materials,
(anatase) at conditions occurring in the solar cell. that is, TiO2(A), TiO2(R), SnO2, and SnO2(D), either as
Our Efb for anatase in TiO2(A) is positive by 0.16 ± 0.04 V grown or upon calcination at 450 °C, together with a blank
against the value for rutile in TiO2(R). This is interesting FTO and Pt-disc electrode by cyclic voltammetry in six
because in aqueous media, we observe just an opposite trend: different electrolyte solutions. Their composition is specified in
TiO2(A) (as well as single-crystal anatase) exhibits Efb ≈ −0.15 Table S1 (Supporting Info). They are actually similar or
V versus RHE (reversible hydrogen electrode), whereas rutile identical to the standard electrolyte solutions used in DSSCs.
(in various forms) has its Efb near 0 V versus RHE.62 This However, in contrast to the measurement of dark current in
finding resonates with the theoretical prediction by Deák et real DSSC devices (which are inherently two-electrode, thin-
al.63,64 that the OH−/H+ adsorption is responsible for reversing film cells), we work here with the standard three-electrode
the band alignment between rutile and anatase, referring to setup, providing optimum cell geometry and the well-defined
clean surfaces in vacuum. In other words, the aprotic reference electrode. Figures 5, 6, and 7 summarize the relevant
electrolyte solution provides the same ordering of anatase/
rutile like vacuum techniques (XPS)65 and some photocatalytic
studies.66 This finding could obviously help to address the
long-term controversy in the field.62−66
The observed Efb of SnO2 is similar or more negative than
that of titania in both forms (Table S2), which is again
unexpected because in aqueous media, we observe an opposite
trend.27,62 The electrochemical doping of SnO2 (section 3.2)
expectably causes some enhancement of ND and ca. 0.1−0.15
V downshift of Efb (Table S2). The doping-induced enhance-
ment of ND is about 10%, matching interestingly our estimate
that ∼9.5% SnO2 is affected by doping (Section 3.2), but this
coincidence could be just casual, considering the experimental
errors of the ND measurement (Table S2) and the different
information depth of XPS (compared to W, see eq 5). On the
other hand, the doping-induced downshift of Efb is significant
within the experimental accuracy. Our found downshift of Efb
is nearly equivalent to the recently reported change of the
Fermi level position in Ru/SnO267 but contradicts the opposite
effect in Li/SnO2 made by solution processing.50 We have no Figure 5. Cyclic voltammograms in I3−/I−-containing electrolyte
simple explanation for these discrepancies, but note that the solution with addition of TBP (full curves) and without addition of
mentioned energy shifts were found by UPS and by using TBP (dashed curves). Red curves are for oxide layers calcined at 450
°C.
different preparative protocols for the Li/SnO2.50 Of interest is
the significant upshift (by 0.3 V) of Efb caused by heat
treatment (Figure S8 and Table S2). A similar upshift (by 0.5 data. Important practical information follows from the onset of
V) was observed in aqueous medium27 and was assigned to cathodic (dark) current because it controls the accessible
crystallization of the parent quasi-amorphous SnO2 to voltage of the solar cell.
cassiterite, without damaging the compact nature of this The corresponding cathodic overvoltages (η) for the
film.27 A recent UPS study by Jeong et al.32 reported a 0.32 eV mediator reduction are summarized in Figure S10 (Supporting
downshift of the Fermi level (enhancement of the work Info). The Pt-disc electrode was selected as the general
function) in ALD-made SnO2 upon heat treatment at 300 °C. reference system for estimation of the η values (Figure S10)
6517 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

evidence for an independent mechanism of the TBP action


(beyond coordination) based on adsorption of TBP to oxidic
surfaces, including also FTO (cf. also Figure 2 and discussion
thereof).
Our different electrode materials combined with three
different redox mediators provide a set of 54 values of η
from which the following general trends are deduced: (i) The
overvoltages increase in the series Co(III/II) < I3−/I− <
Cu(II/I). In all media, they are (ii) roughly comparable for
TiO2(A) and TiO2(R); (iii) the smallest overvoltages are
observed for SnO2; (iv) the largest overvoltages are observed
for SnO2(D); and (v) TBP enhances the overvoltages. The
effect of TBP is further illustrated in Figure S11 (Supporting
Info), which correlates the overvoltages for triiodide reduction
with the corresponding flat band potentials. In addition to the
ubiquitous “TBP enhancement” (of both Efb and η), we
observe a roughly linear correlation of Efb and η. This relation
is surely expectable for the ideally rectifying n-semiconductor/
Figure 6. Cyclic voltammograms in Co(III/II)-containing electrolyte electrolyte solution interfaces because the cathodic charge
solution with addition of TBP (full curves) and without addition of transfer cannot proceed at potentials positive to Efb. However,
TBP (dashed curves). Red curves are for oxide layers calcined at 450 simple energetic arguments (the η/Efb correlation) are
°C.
insufficient to describe the interface because the kinetic
(electrocatalytic) effects play a role, too. (The electrocatalytic
activity of the n-semiconductor surface for various cathodic
reactions, e.g., for H+ reduction, is known to change
significantly even for different facets of the same crystal.70)
The interplay of thermodynamic and kinetic factors is
obviously responsible for the spread of data in Figure S11.
Figure S10 further surveys the effects caused by calcination
of the ESL materials. The cathodic overpotential is significantly
decreased by heat treatment in most materials although the
undoped SnO2 shows interesting mediator-specific features:
The drop of overpotential is strong for the I3− reduction but
much smaller (if any) for the Co(III) and Cu(II) reductions.
Most importantly, our doped tin oxide, SnO2(D), exhibits the
largest overpotential for all the three mediators. Hence, it
appears to be the best material for blocking of dark current in
DSSCs of all types. Unfortunately, SnO2(D) is, at the same
time, also very sensitive to calcination, which erases its
beneficial properties. Heat treatment brings the high reductive
overpotential of SnO2(D) back to the values characteristic for
Figure 7. Cyclic voltammograms in Cu(II/I)-containing electrolyte undoped SnO2 in all three redox systems studied (Figures 5−7,
solution with addition of TBP (full curves) and without addition of Figure S10). This evidences that the chemical or structural
TBP (dashed curves). Red curves are for oxide layers calcined at 450 changes caused by doping are erased by thermal treatment.
°C. Pt and Pt1 are symbols for a platinum-disc electrode in the Yet, SnO2(D) could be applicable for solar cells that do not
presence and absence of TBP, respectively. require calcination, such as “low-temperature” DSSCs, which
do not require heat treatment of the TiO2 photoanode during
because all our three redox mediators exhibit fast electro- their fabrication. Eventually, SnO2(D) appears to be an
chemistry at platinum. (For the same reason, Pt is the attractive material for PSCs because high temperatures are
ubiquitous electrocatalyst, which is used for the cathodes of normally avoidable in their preparation, and doping of SnO2 is
DSSCs with these mediators). All three redox mediators show known to improve the photovoltaic parameters of PSC.31,50,67
sluggish electrochemical kinetics on FTO where the corre- The occurrence of anodic current (if any) detects pinholes
sponding overpotentials for oxidation/reduction define the in the ESL, similarly to the standard “ferrocyanide” test (cf.
electrochemical window in which the dark-current is negligible. Figure S2 and discussion thereof). These pinholes are clearly
This window is quite broad for I3−/I− (which is one of the detectable in the I3−/I− medium and in the calcined TiO2(A),
reasons why these DSSCs work well even without ESL, see TiO2(R), and SnO2(D) (Figure 5). The corresponding anodic
above) but gets narrower for Cu(II/I) and Co(III/II). The Cu currents through pinholes are much smaller in the Co(III/II)-
mediator is unique by the strong effect of TBP, which and Cu(II/I)-containing media. We ascribe this difference to
influences (in contrast to Co and I mediators) also the redox the small ionic radius of I−. Iodide penetrates even across small
processes on Pt (Figure 7). This is ascribed to the coordination pinholes, which are not permeable for bulky molecules of
of TBP to Cu, as it is detailed in earlier works.44−46 Co(bpy)32+ and Cu(tmby)2+. (An analogous situation occurs if
Nevertheless, the overpotential for the reduction of I3− and we compare oxidation of ferrocyanide and spiro-OMeTAD on
Co(III) on FTO is also enhanced by TBP. This provides clear partly cracked films21,41).
6518 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

The undoped SnO2 is perfectly intact against thermal I3−/I− < Cu(II/I). For each particular redox mediator, the
cracking, which confirms earlier electrochemical tests of cathodic overvoltages line up in the series FTO < SnO2 <
blocking quality27,32 and mapping of leakage current by TiO2(A) ≈ TiO2(R) < SnO2(D). Independent of the electrode
AFM.32 Interestingly, however, the tin dioxide film loses its material and mediator type, the overvoltages are always
unique thermal stability upon electrochemical doping when enhanced in the presence of TBP. The flat band potentials
small cracks (transparent selectively for I−) appear in the and cathodic overvoltages tend to scale linearly, but the
SnO2(D) sample upon calcination (Figure 5). These small material-specific (electrocatalytic) effects can influence this
cracks manifest themselves solely as the anodic leakage current correlation, too.
of iodide oxidation at the denuded FTO surface (Figure 5).
Absence of these leakage currents for Co(bpy)32+ oxidation
(Figure 6) and Cu(tmby)2+ oxidation (Figure 7) is attributed

*
ASSOCIATED CONTENT
sı Supporting Information
to larger molecular volume of these organometallic complexes, The Supporting Information is available free of charge at
which cannot penetrate the small pinholes in the calcined https://pubs.acs.org/doi/10.1021/acs.jpcc.9b11883.
SnO2(D). SEM images (Figures S12 and S13; Supporting
Info) show that the morphology of SnO2(D) and blank SnO2 Equivalent circuit for fitting of electrochemical impe-
is dominated by the FTO substrate and is virtually intact by dance data, additional electrochemical data (cyclic
calcination. We observe solely the crystallization-induced voltammograms), Mott−Schottky plots, XPS spectra,
coarsening of the surface,27,32 but the compact nature of and SEM images (PDF)


these films is nearly unperturbed (at the resolution of SEM).
At the same time, the high sensitivity of electrochemical AUTHOR INFORMATION
analysis of ultrasmall pinholes in the compact oxide film is thus
documented. Corresponding Author
Ladislav Kavan − J. Heyrovsky Institute of Physical Chemistry,
4. CONCLUSIONS Czech Acad Sci, Prague 8 182 23, Czech Republic; Laboratory
of Photonics and Interfaces, Institute of Chemical Sciences and
Titanium dioxide (anatase, rutile) and quasi-amorphous tin Engineering, É cole Polytechnique Fédérale de Lausanne,
dioxide were prepared on F-doped SnO2 (FTO) in the form of Lausanne 1015, Switzerland; orcid.org/0000-0003-3342-
compact, nm-thin films, particularly suitable for the develop- 4603; Email: kavan@jh-inst.cas.cz
ment of electron-selective layers in photovoltaics. Hydrolysis of
TiCl4 can provide pure rutile TiO2 at temperatures below 100 Authors
°C, but interestingly, TiO2 (anatase) grows in these layers Zuzana Vlckova Zivcova − J. Heyrovsky Institute of Physical
upon calcination at 450 °C. Chemistry, Czech Acad Sci, Prague 8 182 23, Czech Republic
In acetonitrile electrolyte solution, the flat band potential of Magda Zlamalova − J. Heyrovsky Institute of Physical
TiO2 (rutile) is more negative than that of TiO2 (anatase), Chemistry, Czech Acad Sci, Prague 8 182 23, Czech Republic;
which is opposite to the flat band potential ordering of Department of Inorganic Chemistry, Faculty of Science, Charles
anatase/rutile in aqueous media. This observation helps to University in Prague, Prague 2 128 43, Czech Republic
resolve a long-term controversy in the field. The conduction Shaik M. Zakeeruddin − Laboratory of Photonics and
band minimum of TiO2 (anatase) equals −4.15 ± 0.07 eV at Interfaces, Institute of Chemical Sciences and Engineering, É cole
the conditions mimicking the DSSC’s environment. This value Polytechnique Fédérale de Lausanne, Lausanne 1015,
is recommended to be the energy reference for the Switzerland
corresponding DSSC photoanodes. Michael Grätzel − Laboratory of Photonics and Interfaces,
Simple electrochemical reduction (doping) of tin dioxide in Institute of Chemical Sciences and Engineering, É cole
Li+/acetonitrile medium produces a Li-containing material, Polytechnique Fédérale de Lausanne, Lausanne 1015,
SnO2(D), exhibiting the most negative flat band potential of all Switzerland; orcid.org/0000-0002-0068-0195
the tested oxide electrodes. At the same time, the SnO2(D) has
also the largest overpotential for the reduction of I3−, Complete contact information is available at:
Co(bpy)33+, and Cu(tmby)22+. These beneficial properties https://pubs.acs.org/10.1021/acs.jpcc.9b11883
are promising for both the dye-sensitized and perovskite solar
Notes
cells. However, SnO2(D) is also quite sensitive to calcination
The authors declare no competing financial interest.


when these reduction overpotentials strongly drop and
ultrasmall cracks (permeable selectively to I−) appear in the
film. ACKNOWLEDGMENTS
Systematic screening of all the layers, together with the This work was supported by the European Union’s Horizon
reference materials, blank FTO and Pt, was carried out by 2020 Research and Innovation Program under grant agreement
cyclic voltammetry in six different electrolyte solutions, no. 826013. LK thanks the Grant Agency of the Czech
relevant to DSSC applications. It provided detailed informa- Republic (contract no. 18-08959S) for financial support. We
tion about the mediator type, namely, I3−/I−, Co(III/II), thank Dr. Yameng Ren and Dr. Dan Ren for their help in
Cu(II/I), and effects of calcination and the presence of 4-tert- preparing some TiO2 and SnO2 films.
butylpyridine. The latter effect is boosted in Cu(II/I)-
containing solution in which the collective influence of
coordination of TBP to copper and adsorption to TiO2 are
■ REFERENCES
(1) Burke, A.; Ito, S.; Snaith, H.; Bach, U.; Kwiatkowski, J.; Grätzel,
at play. M. The Function of a TiO2 Compact Layer in Dye-Sensitized Solar
The overvoltages for cathodic reduction of redox mediators Cells Incorporating “Planar” Organic Dyes. Nano Lett. 2008, 8, 977−
on TiO2 or SnO2 generally increase in the series Co(III/II) < 981.

6519 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(2) Cameron, P. J.; Peter, L. M. Characterization of Titanium (21) Kavan, L.; Tétreault, N.; Moehl, T.; Grätzel, M. Electro-
Dioxide Blocking Layers in Dye-Sensitized Nanocrystalline Solar chemical Characterization of TiO2 Blocking Layers for Dye-Sensitized
Cells. J. Phys. Chem. B 2003, 107, 14394−14400. Solar Cells. J. Phys. Chem. C 2014, 118, 16408−16418.
(3) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Dye- (22) Yum, J. H.; Moehl, T.; Yoon, J.; Chandiran, A. K.; Kessler, F.;
Sensitized Solar Cells. Chem. Rev. 2010, 110, 6595−6663. Gratia, P.; Grätzel, M. Toward higher photovoltage: Effect of Blocking
(4) Cameron, P. J.; Peter, L. M.; Hore, S. How Important is the Back Layer on Cobalt Bipyridine Pyrazole Complexes as Redox Shuttle for
Reaction of Electrons via the Substrate in Dye-Sensitized Nano- Dye-Sensitized Solar Cells. J. Phys. Chem. C 2014, 118, 16799−16805.
crystalline Solar Cells? J. Phys. Chem. B 2005, 109, 930−936. (23) Amini, M.; Keshavarzi, R.; Mirkhani, V.; Moghadam, M.;
(5) Kavan, L. Electrochemistry and Dye-sensitized Solar Cells. Curr. Tangestaninejad, S.; Mohammadpoor-Baltork, I.; Sadegh, F. From
Opin. Electrochem. 2017, 2, 88−96. Dense Blocking Layers to Different Templated Films in Dye
(6) Correa-Baena, J. P.; Abate, A.; Saliba, M.; Tress, W.; Jesper Sensitized and Perovskite Solar Cells: Toward Light Transmittance
Jacobsson, T.; Grätzel, M.; Hagfeldt, A. The Rapid Evolution of Management and Efficiency Enhancement. J. Mater. Chem. A 2018, 6,
Highly Efficient Perovskite Solar Cells. Energy Environ. Sci. 2017, 10, 2632−2642.
710−727. (24) Yum, J.-H.; Baranoff, E.; Kessler, F.; Moehl, T.; Ahmad, S.;
(7) Stockhausen, V.; Mesquita, I.; Andrade, L.; Mendes, A. Insights Bessho, T.; Marchioro, A.; Ghadiri, E.; Moser, J.-E.; Yi, C.;
in Perovskite Solar Cell Fabrication: Unraveling the Hidden Nazeeruddin, M. K.; Grätzel, M. Rationally Designed Cobalt
Challenges of Each Layer. IEEE J. Photovoltaics 2018, 8, 1029−1038. Complexes as Redox Shuttle for Dye-Sensitized Solar Cells to Exceed
(8) Kavan, L. Electrochemistry and Perovskite Photovoltaics. Curr. 1000 mV. Nat. Commun. 2012, 3, 631.
Opin. Electrochem. 2018, 11, 122−129. (25) Saygili, Y.; Söderberg, M.; Pellet, N.; Giordano, F.; Cao, Y.;
(9) Nonomura, K.; Vlachopoulos, N.; Unger, E.; Häggman, L.; Muñoz-García, A. B.; Zakeeruddin, S. M.; Vlachopoulos, N.; Pavone,
Hagfeldt, A.; Boschloo, G. Blocking the Charge Recombination with M.; Boschloo, G.; Kavan, L.; Moser, J.-E.; Grätzel, M.; Hagfeldt, A.;
Diiodide Radicals by TiO2 Compact Layer in Dye-sensitized Solar Freitag, M. Copper Bipyridyl Redox Mediators for Dye-Sensitized
Cells. J. Electrochem. Soc. 2019, 166, B3203−B3208. Solar Cells with High Photovoltage. J. Am. Chem. Soc. 2016, 138,
(10) Xia, J.; Masaki, N.; Jiang, K.; Yanagida, S. Deposition of a Thin 15087−15096.
Film of TiOx from a Titanium Metal Target as Novel blocking Layers (26) Cao, Y.; Saygili, Y.; Ummadisingu, A.; Teuscher, J.; Luo, J.;
at Conducting glass/TiO2 Interfaces in Ionic Liquid Mesoscopic TiO2 Pellet, N.; Giordano, F.; Zakeeruddin, S. M.; Moser, J.-E.; Freitag, M.;
Dye-sensitized Solar Cells. J. Phys. Chem. B 2006, 110, 25222−25228. Hagfeldt, A.; Grätzel, M. 11% Efficiency Solid-state Dye-sensitized
(11) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Solar Cells with Copper(II/I) Hole Transport Materials. Nat.
Mueller, E.; Liska, P.; Vlachopoulos, N.; Graetzel, M. Conversion of Commun. 2017, 8, 15390.
Light to Electricity by cis-X2bis(2,2’-bipyridyl-4,4’-dicarboxylate)- (27) Kavan, L.; Steier, L.; Grätzel, M. Ultrathin Buffer Layers of
ruthenium(II) Charge-Transfer Sensitizers (X = Cl-, Br-, I-, CN-, SnO2 by Atomic Layer Deposition: Perfect Blocking Function and
and SCN-) on Nanocrystalline Titanium Dioxide Electrodes. J. Am. Thermal Stability. J. Phys. Chem. C 2017, 121, 342−350.
(28) Kavan, L.; Vlčková-Ž ivcová, Z.; Hubík, P.; Arora, N.; Dar, M. I.;
Chem. Soc. 1993, 115, 6382−6390.
(12) O’Regan, B.; Grätzel, M. A Low-Cost, High-Efficiency Solar Zakeeruddin, S. M.; Grätzel, M. Electrochemical Characterization of
CuSCN Hole-Extracting Thin Films for Perovskite Photovoltaics.
Cell Based on Dye-Sensitized colloidal TiO2 films. Nature 1991, 353,
ACS Appl. Energy Mater. 2019, 2, 4264−4273.
737−740.
(29) Cao, Y.; Liu, Y.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M.
(13) Kavan, L.; O’Regan, B.; Kay, A.; Grätzel, M. Preparation of
Direct Contact of Selective Charge Extraction Layers Enables High-
TiO2 (Anatase) Films on Electrodes by Anodic Oxidative Hydrolysis
Efficiency Molecular Photovoltaics. Joule 2018, 2, 1108−1117.
of TiCl3. J. Electroanal. Chem. 1993, 346, 291−307.
(30) Li, W.; Liu, C.; Zhou, Y.; Bai, Y.; Feng, X.; Yang, Z.; Lu, L.; Lu,
(14) Kavan, L.; Grätzel, M. Highly Efficient Semiconducting TiO2
X.; Chan, K.-Y. Enhanced Photocatalytic Activity in Anatase/TiO2(B)
Photoelectrodes Prepared by Aerosol Pyrolysis. Electrochim. Acta
Core−Shell Nanofiber. J. Phys. Chem. C 2008, 112, 20539−20545.
1995, 40, 643−652. (31) Chen, Y.; Meng, Q.; Zhang, L.; Han, C.; Gao, H.; Zhang, Y.;
(15) Krýsová, H.; Krýsa, J.; Kavan, L. Semi-automatic Spray Yan, H. SnO2-based Electron Transporting Layer Materials for
Pyrolysis Deposition of Thin, Transparent, Titania Films as Blocking Perovskite Solar Cells: A Review of Recent Progress. J. Energy Chem.
Layers for Dye-sensitized and Perovskite Solar Cells. Beilstein J. 2019, 35, 144−167.
Nanotechnol. 2018, 9, 1135−1145. (32) Jeong, S.; Seo, S.; Park, H.; Shin, H. Atomic Layer Deposition
(16) Hurum, D. C.; Agrios, A. G.; Gray, K. A.; Rajh, T.; Thurnauer, of a SnO2 Electron-transporting Layer for Planar Perovskite Solar
M. C. Explaining the Enhanced Photocatalytic Activity of Degussa Cells with a Power Conversion Efficiency of 18.3%. Chem. Commun.
P25 Mixed-phase TiO2 Using EPR. J. Phys. Chem. B 2003, 107, 2019, 55, 2433−2436.
4545−4549. (33) Zhang, G.; Yang, D.; Sacher, E. X-ray Photoelectron
(17) Wang, P.; Shao, Z.; Ulfa, M.; Pauporté, T. Insights into the Spectroscopic Analysis of Pt Nanoparticles on Highly Oriented
Hole Blocking Layer Effect on the Perovskite Solar Cell Performance Pyrolytic Graphite, Using Symmetric Component Line Shapes. J.
and Impedance Response. J. Phys. Chem. C 2017, 121, 9131−9141. Phys. Chem. C 2007, 111, 565−570.
(18) Yella, A.; Heiniger, L.-P.; Gao, P.; Nazeeruddin, M. K.; Grätzel, (34) Aygüler, M. F.; Hufnagel, A. G.; Rieder, P.; Wussler, M.;
M. Nanocrystalline Rutile Electron Extraction Layer Enables Low- Jaegermann, W.; Bein, T.; Dyakonov, V.; Petrus, M. L.; Baumann, A.;
temperature Solution Processed Perovskite Photovoltaics with 13.7% Docampo, P. Influence of Fermi Level Alignment with Tin Oxide on
Efficiency. Nano Lett. 2014, 14, 2591−2596. the Hysteresis of Perovskite Solar Cells. ACS Appl. Mater. Interfaces
(19) Liu, I. P.; Lin, W. H.; Tseng-Shan, C. M.; Lee, Y. L. Importance 2018, 10, 11414−11419.
of Compact Blocking Layers to the Performance of Dye-Sensitized (35) Anaraki, E. H.; Kermanpur, A.; Steier, L.; Domanski, K.;
Solar Cells under Ambient Light Conditions. ACS Appl. Mater. Matsui, T.; Tress, W.; Saliba, M.; Abate, A.; Grätzel, M.; Hagfeldt, A.;
Interfaces 2018, 10, 38900−38905. Correa-Baena, J. P. Highly Efficient and Stable Planar Perovskite Solar
(20) Masood, M. T.; Weinberger, C.; Sarfraz, J.; Rosqvist, E.; cells by Solution-processed Tin Oxide. Energy Environ. Sci. 2016, 9,
Sandén, S.; Sandberg, O. J.; Vivo, P.; Hashmi, G.; Lund, P. D.; 3128−3134.
Ö sterbacka, R.; Smått, J.-H. Impact of Film Thickness of Ultrathin (36) Tavakoli, M. M.; Saliba, M.; Yadav, P.; Holzhey, P.; Hagfeldt,
Dip-Coated Compact TiO2 Layers on the Performance of Mesoscopic A.; Zakeeruddin, S. M.; Grätzel, M. Synergistic Crystal and Interface
Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2017, 9, 17906− Engineering for Efficient and Stable Perovskite Photovoltaics. Adv.
17913. Energy Mater. 2019, 9, 1802646.

6520 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(37) Roose, B.; Baena, J.-P. C.; Gödel, K. C.; Graetzel, M.; Hagfeldt, (54) Kavan, L.; Grätzel, M.; Gilbert, S. E.; Klemenz, C.; Scheel, H. J.
A.; Steiner, U.; Abate, A. Mesoporous SnO2 Electron Selective Electrochemical and Photoelectrochemical Investigation of Single-
Contact Enables UV-stable Perovskite Solar Cells. Nano Energy 2016, Crystal Anatase. J. Am. Chem. Soc. 1996, 118, 6716−6723.
30, 517−522. (55) Moehl, T.; Suh, J.; Sévery, L.; Wick-Joliat, R.; Tilley, S. D.
(38) Hsu, C. H.; Mansfeld, F. Technical Note: Concerning the Investigation of (Leaky) ALD TiO2 Protection Layers for Water-
Conversion of the Constant Phase Element Parameter Y0 into a Splitting Photoelectrodes. ACS Appl. Mater. Interfaces 2017, 9,
Capacitance. Corrosion 2001, 57, 747−748. 43614−43622.
(39) Berger, T.; Monllor-Satoca, D.; Jankulovska, M.; Lana- (56) Schulz, P.; Edri, E.; Kirmayer, S.; Hodes, G.; Cahen, D.; Kahn,
Villarreal, T.; Gómez, R. The Electrochemistry of Nanostructured A. Interface Energetics in Organo-metal Halide Perovskite-based
Titanium Dioxide Electrodes. ChemPhysChem 2012, 13, 2824−2875. Photovoltaic Cells. Energy Environ. Sci. 2014, 7, 1377−1381.
(40) Frank, O.; Zukalova, M.; Laskova, B.; Kürti, J.; Koltai, J.; Kavan, (57) Kashiwaya, S.; Morasch, J.; Streibel, V.; Toupance, T.;
L. Raman Spectra of Titanium Dioxide (Anatase, Rutile) with Jaegermann, W.; Klein, A. The Work Function of TiO2. Surfaces
Identified Oxygen Isotopes (16, 17, 18). Phys. Chem. Chem. Phys. 2018, 1, 73−89.
2012, 14, 14567−14572. (58) Balajka, J.; Hines, M. A.; DeBenedetti, W. J. I.; Komora, M.;
(41) Krysova, H.; Zlamalova, M.; Tarabkova, H.; Jirkovsky, J.; Frank, Pavelec, J.; Schmid, M.; Diebold, U. High-affinity Adsorption Leads to
O.; Kohout, M.; Kavan, L. Rutile TiO2 Thin Film Electrodes with Molecularly Ordered Interfaces on TiO2 in Air and Solution. Science
Excellent Blocking Function and Optical Transparency. Electrochim. 2018, 361, 786−789.
Acta 2019, 321, 134685. (59) Boschloo, G.; Hagfeldt, A. Characteristics of the Iodide/
(42) Bonu, V.; Gupta, B.; Chandra, S.; Das, A.; Dhara, S.; Tyagi, A. triiodide Redox Mediator in Dye-sensitized Solar Cells. Acc. Chem.
K. Electrochemical Supercapacitor Performance of SnO2 Quantum Res. 2009, 42, 1819−1826.
(60) Redmond, G.; Fitzmaurice, D. Spectroscopic Determination of
Dots. Electrochim. Acta 2016, 203, 230−237.
Flatband Potentials for Polycrystalline Titania Electrodes in Non-
(43) Wu, J.; Lan, Z.; Lin, J.; Huang, M.; Huang, Y.; Fan, L.; Luo, G.
aqueous Solvents. J. Phys. Chem. 1993, 97, 1426−1430.
Electrolytes in Dye-Sensitized Solar Cells. Chem. Rev. 2015, 115,
(61) Jung, H. S.; Kim, H. Origin of Low Photocatalytic Activity of
2136−2173.
Rutile TiO2. Electron. Mater. Lett. 2009, 5, 73−76.
(44) Kavan, L.; Saygili, Y.; Freitag, M.; Zakeeruddin, S. M.; Hagfeldt,
(62) Kavan, L. Conduction Band Engineering in Semiconducting
A.; Grätzel, M. Electrochemical Properties of Cu(II/I)-Based Redox Oxides (TiO2, SnO2): Applications in Perovskite Photovoltaics and
Mediators for Dye-Sensitized Solar Cells. Electrochim. Acta 2017, 227, Beyond. Catal. Today 2019, 328, 50−56.
194−202. (63) Kullgren, J.; Aradi, B.; Frauenheim, T.; Kavan, L.; Deák, P.
(45) Saygili, Y.; Stojanovic, M.; Michaels, H.; Tiepelt, J.; Teuscher, Resolving the Controversy About the Band Alignment Between Rutile
J.; Massaro, A.; Pavone, M.; Giordano, F.; Zakeeruddin, S. M.; and Anatase: the Role of OH‑/H+ Adsorption. J. Phys. Chem. C 2015,
Boschloo, G.; Moser, J.-E.; Grätzel, M.; Muños-García, A. B.; 119, 21952−21958.
Hagfeldt, A.; Freitag, M. Effect of Coordination Sphere Geometry (64) Deák, P.; Kullgren, J.; Aradi, B.; Frauenheim, T.; Kavan, L.
of Copper Redox Mediators on Regeneration and Recombination Water Splitting and the Band Edge Positions of TiO2. Electrochim.
Behavior in Dye-Sensitized Solar Cell Applications. ACS Appl. Energy Acta 2016, 199, 27−34.
Mater. 2018, 1, 4950−4962. (65) Scanlon, D. O.; Dunnill, C. W.; Buckeridge, J.; Shevlin, S. A.;
(46) Ferdowsi, P.; Saygili, Y.; Zakeeruddin, S. M.; Mokhtari, J.; Logsdail, A. J.; Woodley, S. M.; Catlow, C. R. A.; Powell, M. J.;
Grätzel, M.; Hagfeldt, A.; Kavan, L. Alternative Bases to 4-tert- Palgrave, R. G.; Parkin, I. P.; Watson, G. W.; Keal, T. W.; Sherwood,
butylpyridine for Dye-sensitized Solar cells Employing Copper Redox P.; Walsh, A.; Sokol, A. A. Band Alignment of Rutile and Anatase
Mediator. Electrochim. Acta 2018, 265, 194−201. TiO2. Nat. Mater. 2013, 12, 798−801.
(47) Boschloo, G.; Häggman, L.; Hagfeldt, A. Quantification of the (66) Quesada-Cabrera, R.; Sotelo-Vazquez, C.; Bear, J. C.; Darr, J.
Effect of 4-tert-butylpyridine addition to I‑/I3‑ Redox Electrolytes in A.; Parkin, I. P. Photocatalytic Evidence of the Rutile-to-Anatase
Dye-sensitized Nanostructured TiO2 Solar Cells. J. Phys. Chem. C Electron Transfer in Titania. Adv. Mater. Interfaces 2014, 1, 1400069.
2006, 110, 13144−13150. (67) Akin, S. Hysteresis-Free Planar Perovskite Solar Cells with a
(48) Kim, J.-Y.; Kim, J. Y.; Lee, D.-K.; Kim, B.; Kim, H.; Ko, M. J. Breakthrough Efficiency of 22% and Superior Operational Stability
Importance of 4-tert-butylpyridine in Electrolyte for Dye-sensitized over 2000 h. ACS Appl. Mater. Interfaces 2019, 11, 39998−40005.
Solar Cells Employing SnO2 Electrode. J. Phys. Chem. C 2012, 116, (68) Kashiwaya, S.; Toupance, T.; Klein, A.; Jaegermann, W. Fermi
22759−22766. Level Positions and Induced Band Bending at Single Crystalline
(49) Kwoka, M.; Ottaviano, L.; Passacantando, M.; Santucci, S.; Anatase (101) and (001) Surfaces: Origin of the Enhanced
Czempik, G.; Szuber, J. XPS Study of the Surface Chemistry of L- Photocatalytic Activity of Facet Engineered Crystals. Adv. Energy
CVD SnO2 Thin Films after Oxidation. Thin Solid Films 2005, 490, Mater. 2018, 8, 1802195.
36−42. (69) Pfeifer, V.; Erhart, P.; Li, S.; Rachut, K.; Morasch, J.; Brötz, J.;
(50) Park, M.; Kim, J. Y.; Son, H. J.; Lee, C. H.; Jang, S. S.; Ko, M. J. Reckers, P.; Mayer, T.; Rühle, S.; Zaban, A.; Mora Seró, I.; Bisquert,
Low-temperature Solution-processed Li-doped SnO2 as an Effective J.; Jaegermann, W.; Klein, A. Energy Band Alignment Between
Electron Transporting Layer for High-performance Flexible and Anatase and Rutile TiO2. J. Phys. Chem. Lett. 2013, 4, 4182−4187.
Wearable Perovskite Solar Cells. Nano Energy 2016, 26, 208−215. (70) Nebel, R.; Macounová, K. M.; Tarábková, H.; Kavan, L.; Krtil,
(51) Park, J. W.; Park, C. M. A Fundamental Understanding of Li P. Selectivity of Photoelectrochemical Water Splitting on TiO2
Insertion/Extraction Behaviors in SnO and SnO2. J. Electrochem. Soc. Anatase Single Crystals. J. Phys. Chem. C 2019, 123, 10857−10867.
2015, 162, A2811−A2816.
(52) Dixon, D.; Á vila, M.; Ehrenberg, H.; Bhaskar, A. Difference in
Electrochemical Mechanism of SnO2 Conversion in Lithium-Ion and
Sodium-Ion Batteries: Combined in Operando and Ex Situ XAS
Investigations. ACS Omega 2019, 4, 9731−9738.
(53) Baena, J. P. C.; Steier, L.; Tress, W.; Saliba, M.; Neutzner, S.;
Matsui, T.; Giordano, F.; Jacobsson, T. J.; Kandada, A. R. S.;
Zakeeruddin, S. M.; Petrozza, A.; Abate, A.; Nazeeruddin, M. K.;
Grätzel, M.; Hagfeldt, A. Highly Efficient Planar Perovskite Solar
Cells Through Band Alignment Engineering. Energy Environ. Sci.
2015, 8, 2928−2934.

6521 https://dx.doi.org/10.1021/acs.jpcc.9b11883
J. Phys. Chem. C 2020, 124, 6512−6521

You might also like