Ti

You might also like

You are on page 1of 7

LETTER

pubs.acs.org/NanoLett

Facile Synthesis of Highly Photoactive r-Fe2O3-Based Films for Water


Oxidation
Gongming Wang,† Yichuan Ling,† Damon A. Wheeler,† Kyle E. N. George,‡ Kimberly Horsley,‡
Clemens Heske,‡ Jin Z. Zhang,† and Yat Li*,†

Department of Chemistry and Biochemistry, University of California Santa Cruz, Santa Cruz, California 95064, United States

Department of Chemistry, University of Nevada, Las Vegas, Nevada 89154, United States

bS Supporting Information
ABSTRACT: This work reports a facile method for preparing
highly photoactive R-Fe2O3 films as well as their implementation as
photoanodes for water oxidation. Transparent R-Fe2O3 films were
prepared by a new depositionannealing (DA) process using
nontoxic iron(III) chloride as the Fe precursor, followed by
annealing at 550 °C in air. Ti-doped R-Fe2O3 films were prepared
by the same method, with titanium butoxide added as the Ti
precursor. Impedance measurements show that the Ti-dopant
serves as an electron donor and increases the donor density by
2 orders of magnitude. The photoelectrochemical performance of
undoped and Ti-doped R-Fe2O3 photoanodes was characterized and optimized through controlled variation of the Fe and Ti
precursor concentration, annealing conditions, and the number of DA cycles. Compared to the undoped sample, the photocurrent
onset potential of Ti-doped R-Fe2O3 is shifted about 0.10.2 V to lower potential, thus improving the photocurrent and incident
photon to current conversion efficiency (IPCE) at lower bias voltages. Significantly, the optimized Ti-doped R-Fe2O3 film achieved
the highest photocurrent density (1.83 mA/cm2) and IPCE values at 1.02 V vs RHE for R-Fe2O3 photoanode. The enhanced
photocurrent is attributed to the improved donor density and reduced electronhole recombination at the time scale beyond a few
picoseconds, as a result of Ti doping.
KEYWORDS: Photoelectrochemical water splitting, deposition-annealing process, R-Fe2O3, Ti doping

A number of semiconductor metal oxides such as TiO2, ZnO,


WO3, and R-Fe2O3 have been investigated for photoelec-
trochemical (PEC) water splitting.113 Among these metal
These dopants have been shown to play important roles in
improving the optical absorption coefficient, the donor density,
electrical conductivity, and the flat band potential of R-Fe2O3.
oxides, R-Fe2O3 (hematite) has received much attention due Second, nanomaterials with very large interfacial area between
to its favorable optical band gap (∼2.2 eV),14 excellent chemical electrolyte and semiconductor would facilitate the charge collec-
stability, natural abundance, and low cost. It has been theoreti- tion by reducing the diffusion length for photoexcited holes.
cally predicted that a semiconductor with this band gap can Nanostructured R-Fe2O3 materials have thus been synthesized
achieve a solar-to-hydrogen efficiency of 16.8%.15 However, the by numerous approaches, including hydrothermal methods,13,24
reported efficiencies of R-Fe2O3 are notoriously lower than the spray pyrolysis,14,21,23 DC magnetron sputtering,22 electro-
predicted value, mainly due to its very short lifetime of photo- chemical deposition,25 atmospheric pressure chemical vapor de-
generated charge carriers (<10 ps)16,17 and short hole diffusion position (APCVD),1,2,10,20,26 colloidal solution assembly,19,27
length (∼24 nm).18 In this regard, very thin R-Fe2O3 films anodization,28 flame oxidation of iron foils,29 and atomic layer
should be used for facilitating transport/collection of electrons deposition (ALD).4 Among these methods, the nanostructured
and holes. However, R-Fe2O3 is an indirect band gap semicon- R-Fe2O3 films prepared by spray pyrolysis, APCVD, and ALD
ductor, and thus it requires a relatively thick layer of active have shown excellent PEC performance. However, these tech-
material for the complete absorption of solar light. Besides, the niques require special instruments such as atomizers14 or chemi-
band positions of R-Fe2O3 do not necessarily straddle the H2 and cal vapor deposition and ALD4 systems that use toxic and
O2 evolution potentials. Therefore, a relatively large external flammable metalorganics as precursors.1,2,20,26
bias (0.81.0 V) is typically required for water splitting using Developing a facile, safe, and cost-effective method to fabricate
R-Fe2O3 materials as photoanodes.2,4,19 element-doped nanostructured R-Fe2O3 films is highly desirable.
To address these limitations, a number of unique element- In this work, we report a simple depositionannealing (DA)
doped, nanostructured R-Fe2O3 materials have been designed and
fabricated. First, both nonmetal1,2,20 and metal elements13,19,2123 Received: July 7, 2011
have been extensively investigated as dopants for R-Fe2O3 materials. Published: July 18, 2011

r 2011 American Chemical Society 3503 dx.doi.org/10.1021/nl202316j | Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

Figure 1. (a) A scheme showing the DA procedure for the growth of R-


Fe2O3 films. (bf) SEM images collected for R-Fe2O3 films prepared
with 80, 50, 30, 10, and 5 mM FeCl3 precursor solutions. All scale bars
are 500 nm.

method for the fabrication of transparent, mesoporous R-Fe2O3


films on a conducting glass substrate, using nontoxic iron(III)
chloride (FeCl3) as the Fe precursor. This is also a versatile
method that allows dopants such as Ti to be easily incorpo-
rated into R-Fe2O3 structures. The size of R-Fe2O3 nanopar-
ticles, film thickness, as well as the doping level can be tailored by
tuning the growth conditions. In comparison to the spray
pyrolysis and APCVD methods, this growth approach avoids Figure 2. (a) Plots of the photocurrent density of R-Fe2O3 films
the use of specialized equipment and metalorganic precursors. prepared with different precursor concentrations as a function of
Significantly, the undoped and Ti-doped R-Fe2O3 films prepared the number of DA cycles. All photocurrent densities were measured at
by this method showed excellent PEC performance. In particular, 1.47 V vs RHE. (b) Linear sweeps collected for R-Fe2O3 films prepared
the photocurrent onset potential of Ti-doped R-Fe2O3 is shifted with 10 mM FeCl3 precursor solution at different annealing tempera-
about 0.10.2 V to lower potential and achieves the best tures (the annealing time was 4 h in each case).
photocurrent density of 1.83 mA/cm2 at relatively low bias,
1.02 V vs RHE. To shed light on the origins of the improvement densely packed when precursor concentrations of 10 and 5 mM
induced by Ti doping, R-Fe2O3 films were investigated by X-ray were used, possibly due to the reduced size of the nanoparticles
photoelectron spectroscopy, electrochemical measurements, and (Figure S2, Supporting Information). Raman spectra were
ultrafast laser spectroscopy. collected for these samples to shed further light on the correla-
R-Fe2O3 films were prepared on FTO-coated glass substrates tion between concentration and particle size (Figure S3, Sup-
by a DA method (Experimental Methods, Supporting In- porting Information). The Raman spectrum of the R-Fe2O3 film
formation), as shown in Figure 1a. When the sample was heated prepared at the highest precursor concentration of 80 mM
at 350 °C in air, the color of the FeCl3-covered FTO substrate exhibits all six characteristic Raman peaks of R-Fe2O3.30 These
turned pale yellow, which is believed to be due to the conversion peaks become broader and featureless as the precursor concen-
of FeCl3 into a thin β-FeOOH film. This film was subsequently tration is decreased, indicating the decrease of grain size in the
converted into a red film by further sintering at 550 °C for 4 h. material.31 The results support the hypothesis that the film
X-ray diffraction (XRD) data collected from the yellow and red morphology is correlated with a size reduction of the R-Fe2O3
films were indexed to the characteristic peaks of β-FeOOH nanoparticles.
(JCPDS 34-1266) and R-Fe2O3 (JCPDS 33-0664), respectively PEC performance of R-Fe2O3 films was measured in a 1.0 M
(Figure S1a, Supporting Information). These data confirm the NaOH electrolyte solution using a three-electrode electroche-
successful fabrication of the R-Fe2O3 phase and indicate that the mical cell coupled with a solar simulator. The measured poten-
β-FeOOH phase was the process intermediate. The as-prepared tials vs Ag/AgCl were converted to the reversible hydrogen ele-
films appear to be uniform. The film transparency, related to the ctrode (RHE) scale according to the Nernst equation
film thickness and/or the size of individual R-Fe2O3 nanoparti-
cles (i.e., the density of the film), decreases with increasing ERHE ¼ EAg=AgCl þ 0:059pH þ E°Ag=AgCl
number of DA cycles and concentration of the FeCl3 solution
(Figure S1b, Supporting Information). where ERHE is the converted potential vs RHE, E°Ag/AgCl = 0.1976 V
Scanning electron microscopy (SEM) images were collected at 25 °C, and EAg/AgCl is the experimentally measured potential
for the R-Fe2O3 films prepared with increasing precursor con- against a Ag/AgCl reference. Figure 2a shows plots of the photo-
centration in order to understand the correlation of these key current densities of R-Fe2O3 films, prepared with various FeCl3
growth parameters to the material morphology. Panels bf of precursor concentrations, as a function of the number of DA cycles.
Figure 1 show the SEM images of the R-Fe2O3 films prepared For each precursor concentration and with increasing number of
with FeCl3 precursor concentrations in a range from 5 to 80 mM. DA cycles, the photocurrent densities of R-Fe2O3 films initially
These images reveal that the size of R-Fe2O3 nanoparticles in the increase to a maximal value and then decrease again. This suggests
films decreases with precursor concentration. The films appear that the photocurrent is strongly correlated with the thickness and
3504 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

density of the R-Fe2O3 film. The initial increase of photocurrent annealing could also increase interparticle contacts and/or improve
can be ascribed to the increased amount of photoactive material, the contact at the interface of R-Fe2O3 film and the substrate. In
which enhances the light absorption as well as the total number of contrast, the prolonged annealing at temperatures of 550 °C
photoexcited carriers. Once the R-Fe2O3 film thickness reaches a or above caused glass distortion. Furthermore, the decrease of
certain threshold, charge carrier recombination becomes increas- photocurrent after annealing at temperatures higher than 600 °C
ingly important, since the very short carrier diffusion length effec- is ascribed to an increased resistance of the FTO substrate (from
tively limits the ability to remove charges from the semiconductor. 22 to 167 Ω after 4 h of annealing at 700 °C). Therefore, 550 °C
On the basis of the maximal photocurrent densities obtained at was chosen as the optimal annealing temperature.
each precursor concentration, the optimal R-Fe2O3 film thickness The photocurrent densities of the R-Fe2O3 films were also
is estimated to be in a range of 80160 nm (Figure S4, Supporting measured as a function of annealing time (Figure S7b, Supporting
Information). Information). The photocurrent densities achieved a maximal
Moreover, the maximal photocurrent densities of R-Fe2O3 level after 34 h of annealing. The resistance of the substrate was
films increase with decreasing precursor concentration (Figure slightly increased from 22 to 32 Ω after 4 h of annealing at
S5, Supporting Information). The film prepared with the lowest 550 °C. For an undoped R-Fe2O3 film with optimized precursor
precursor concentration of 5 mM achieved the highest peak concentration (10 mM), number of DA cycles (10), film thickness
photocurrent density of 1.78 mA/cm2 at 1.47 V vs RHE. The (ca. 140 nm), annealing temperature (550 °C), and annealing
improved PEC activity of this R-Fe2O3 film is due to the reduced time (4 h), a maximized photocurrent density of 0.91 mA/cm2 at
size of R-Fe2O3 nanoparticles, although an increased number of 1.23 V vs RHE was achieved. To our knowledge, this photocurrent
DA cycles (24) was required to achieve the optimal film thickness density is considerably higher than that for the best undoped
when using such a diluted precursor solution. The effect of particle R-Fe2O3 films prepared by hydrothermal methods using the
size on the PEC performance of R-Fe2O3 can be understood in same FeCl3 precursor.13 The enhanced photocurrent density
terms of the charge separation and transport at the semiconductor/ can be attributed to the small size of the R-Fe2O3 particles,
electrolyte interface. In small particles, the average distance which reduces the impact of unfavorably short diffusion lengths
between the photoexcited holes and the semiconductor/electro- for photoexcited holes and facilitates the charge separation
lyte interface is reduced, which improves the hole collection due to a large contribution of the depletion region in these
efficiency. Furthermore, the decrease of particle size can enhance R-Fe2O3 particles.
the charge separation by depleting the R-Fe2O3 nanoparticle. To further enhance the PEC performance, we have studied the
When a semiconductor nanoparticle is brought into contact with impact of deliberate doping of the R-Fe2O3 films. Despite the fact
the electrolyte, a space-charge layer32 and an interface dipole that R-Fe2O3 has a favorable optical band gap for light absorp-
forms at the semiconductor/electrolyte interface, replacing the tion, its potential use as a photoanode for water splitting is
surface band bending and surface dipole at the former semicon- limited by the relatively poor electronic properties, as discussed
ductor surface. In contrast, the particle core (of sufficiently large above, as well as the large potential difference between flatband
particles) can be regarded as a diffusion region (Figure S6, position and H+/H2 reduction potential. Elemental doping is
Supporting Information). Photoexcited electrons and holes are known to be a promising approach to alter the electronic pro-
more effectively separated by the electric field in the space-charge perties of semiconductors. In this work, we report the synthesis
layer than in the diffusion region. For R-Fe2O3, most of the of Ti-doped R-Fe2O3 films and investigate the Ti-doping effect
electronhole pairs that are created beyond the depletion layer on donor density, electronic structure, and excited state lifetime
are very likely lost via recombination, due to the very short of R-Fe2O3 films. As mentioned in the experimental section
electron and hole diffusion lengths. Therefore, the increased (Supporting Information), Ti was introduced into R-Fe2O3 films
contribution of the depletion layer in small R-Fe2O3 particles can by mixing titanium butoxide into the FeCl3 precursor solution. In
reduce the electronhole recombination loss. comparison to undoped samples, R-Fe2O3 films prepared in the
The PEC performance of R-Fe2O3 films was also analyzed by presence of titanium butoxide show a less densely packed
varying the annealing temperature and time. Here, we use R- structure with relatively large R-Fe2O3 particles (Figure S9,
Fe2O3 films prepared with 10 mM FeCl3 solution as an example Supporting Information).
to elucidate the effects of the annealing process. Figure 2b shows X-ray photoelectron spectroscopy (XPS) survey spectra of a
the linear sweep curves of R-Fe2O3 films annealed at different Ti-doped R-Fe2O3 film with 10 mM FeCl3 precursor concentra-
temperatures. The photocurrent densities of these films are tion confirm the presence of Ti (Figure 3a), while no Ti signal
enhanced as the annealing temperature is increased from 400 is found for the undoped sample (Figure S10, Supporting
to 600 °C and then decrease with further increase of annealing Information), as expected. We note that the Ti signal at the
temperature (the photocurrent densities measured at 1.47 V vs surface of the Ti-doped R-Fe2O3 film is substantial, indicating
RHE are shown in Figure 7a, Supporting Information). Raman that Ti concentration levels are higher than traditional “doping”
spectra collected for R-Fe2O3 films prepared at different anneal- levels, as discussed in the experimental section in the Supporting
ing temperatures (Figure S8, Supporting Information) again Information. The Ti-doped sample shows significant Na and C
confirm the as-prepared β-FeOOH film was converted into signals at the surface, while the Na signal and, to some degree, the
R-Fe2O3 upon annealing. All six characteristic Raman peaks30 of C signal are reduced for the undoped sample. Both samples show
R-Fe2O3 were observed after annealing up to 600 °C. Importantly, small Cl peaks, while only the undoped sample exhibits a Sn
the full width at half-maximum of the Raman peaks decreased as signal. Sn and Na most likely originate from the FTO layer and
the annealing temperature was increased, suggesting an enhance- the glass substrate, respectively. The presence of Cl is presumably
ment of R-Fe2O3 crystallinity.33 Therefore, the increased photo- due to the use of FeCl3 solutions as the Fe precursor. Carbon
current with increased annealing temperature (from 400 to is believed to be included during sample preparation (in air)
600 °C) is believed to be the result of a reduced defect density and subsequent handling. Note that the ratio between the Fe
in annealed R-Fe2O3 samples. Furthermore, the high-temperature signals (e.g., the Fe 2p) and the O 1s signal is much lower for the
3505 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

Figure 3. (a) XPS survey spectrum and high-resolution (b) Ti 2p, (c) Fe 2p, and (d) O 1s XPS spectra collected for the Ti-doped R-Fe2O3 film (10 mM
FeCl3 precursor). The Ti 2p spectrum of a metallic Ti reference is also included in (b), and the O 1s spectrum of the undoped R-Fe2O3 film was added
in (d) for comparison.

Ti-doped sample than for the undoped sample. This indicates This satellite is most likely indicative of the presence of Fe3+
that the Ti atoms present at the surface of the Ti-doped sample species.38 Again, there is no evidence to support the presence of
very likely are bound in an oxide environment, as will be dis- Fe2+, which should give rise to a satellite peak located at 717.2 eV,
cussed next. 6.0 eV from the main peak (as indicated by the vertical dashed
The Ti 2p detail spectrum of the Ti-doped sample is shown in line in Figure 3c). In comparison to the undoped R-Fe2O3 film,
Figure 3b, together with a metallic Ti reference spectrum. From the Ti-doped sample shows evidence for an additional Fe species
the peak position and line shape, it is evident that the Ti atoms in at a binding energy of ∼708 eV (Figure S11a, Supporting
the Ti-doped sample are not in a metallic environment. The Ti Information). This shoulder does not match the binding energy
2p3/2 binding energy of 458.6 eV is fully consistent with typical of metallic Fe (706.7707.1 eV34,35), FeOOH, or any of the
values reported for TiO2 (458.7 eV34 and ca. 458.6 to 459.3 eV35). other oxides discussed above. We speculate that it might be due
Furthermore, the peaks of the Ti-doped sample exhibit a sym- to the presence of FeC or FeTi bonds in the Ti-doped sample;
metric line shape, which is generally found when a band gap Peng et al. reported 706.9 eV as the Fe 2p3/2 binding energy in
inhibits low-energy excitations in the photoemission process; the FeTi and FeTiN thin films,39 and the Fe 2p3/2 binding
metallic Ti peaks, in contrast, show the well-known asymmetry energy for Fe3C is reported to be 708.1 eV.34 Figure 3d shows the
toward higher binding energy (Figure 3b). Note that the Ti 2p1/2 O 1s XPS spectra collected from the doped and undoped R-
peak is significantly broader than the Ti 2p3/2 peak due to pre- Fe2O3 films. The binding energy of the main line (530.1 eV for
sence of a CosterKronig decay channel of the Ti 2p1/2 core hole, the undoped and 530.3 eV for the Ti-doped sample) is consistent
leading to lifetime broadening of the photoemission line. The Fe with the reported value for Fe2O3 (530.035,38 and 529.9 eV37)
2p3/2 line of the Ti-doped R-Fe2O3 film in Figure 3c is found at and Fe3O4 (530.034 and 530.3 eV37). Both samples show a shoulder
a binding energy of ∼711.2 eV. Within the common variations at higher binding energy (more pronounced for the undoped
of binding energy determination of band gap materials, this sample), which could be attributed to OH and/or OC bonds.
value is consistent with typical values observed for Fe2O3 (Fe3+, Finally, the valence band region of the two R-Fe2O3 films was
710.9 eV34 or 711.2 eV;36 FeOOH (Fe3+, ∼711.2 eV36 and also measured by Mg KR XPS (Figure S11b, Supporting In-
711.3711.9 eV35) and Fe3O4 (a mix of Fe2+/Fe3+, 709.5/ formation). The two valence band spectra are similar, and an
711.2 eV36 and 711.6 eV).37 The binding energy is not compatible estimate of the valence band maximum by linear extrapolation to
with that of FeO (Fe2+, 709.6 eV34 and 709.1709.5 eV35). the baseline derives a band edge position of ∼1.70 ((0.20) eV
A satellite peak of the Fe 2p3/2 main line is observed at approx- below the Fermi energy in each case. Assuming that the optical
imately 719.3 eV, i.e., approximately 8.1 eV below the main line. band gap of (bulk) R-Fe2O3 of ∼2.2 eV14 is preserved at the
3506 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

suggesting that the carrier transport in the Ti-doped sample is


more efficient. Finally, the photocurrent density saturates at higher
applied bias as a result of insufficient numbers of photoexcited
carriers (depending on the light intensity). The Ti-doped R-Fe2O3
film shows a 2-fold enhanced photocurrent density compared to
the undoped sample over the entire potential window studied.
The Ti-doped R-Fe2O3 film achieved a remarkable photocurrent
density of 2.80 mA/cm2 at 1.23 V vs RHE, which is comparable
to the benchmark values previously reported for R-Fe2O3 struc-
tures at the same potential.2,4,10 Compared to the undoped
sample, the photocurrent onset potential of Ti-doped R-Fe2O3 is
shifted about 0.10.2 V to lower potential, thus improving the
photocurrent at lower bias voltage. A similar shift of onset poten-
tial was demonstrated in previous studies for Si-20 and Ti-doped
samples.42,43 Significantly, our Ti-doped R-Fe2O3 film achieved
a photocurrent density of 1.83 mA/cm2 at 1.02 V vs RHE, which
is substantially higher than that observed in other R-Fe2 O3
structures.2,4,10 To our knowledge, this is the highest photo-
current density for R-Fe2O3 obtained at this bias.
To quantitatively investigate the photoactivity of the R-Fe2O3
films as a function of wavelength, IPCE measurements were
performed on the optimized undoped and Ti-doped R-Fe2O3
films at 1.02 V vs RHE. IPCE can be expressed as
IPCE ¼ ð1240IÞ=ðλJlight Þ
Figure 4. (a) Comparison of photocurrent densities collected for
where I is the measured photocurrent density, λ is the wavelength
optimized undoped and Ti-doped R-Fe2O3 films. (b) IPCE spectra of
these two films collected at 1.02 V vs RHE. of the incident light, and Jlight is the measured irradiance at the
specific measurement wavelength. In comparison to the undoped
surface, this would place the surface conduction band minimum sample, the optimized Ti-doped R-Fe2O3 film shows substan-
of the R-Fe2O3 films approximately 0.5 eV above the Fermi tially enhanced IPCE values over the entire wavelength range of
energy (note that the assumption of identical bulk and surface 350610 nm (Figure 4b). Above 620 nm, the photoresponse of
band gaps has been shown to be incorrect for a variety of com- both R-Fe2O3 films drops to zero, in accordance with the band
pound semiconductors).40,41 The fact that both films exhibit the gap of R-Fe2O3. The Ti-doped sample exhibits the highest IPCE
same position of the valence band maximum indicates a negligible values for R-Fe2O3 obtained at 1.02 V vs RHE to date. Given that
effect of Ti-doping on the valence band position at the surface the irradiance spectrum may vary for different light sources (e.g.,
(most likely due to Fermi level pinning at the surface). lamp lifetime and power) and filters used in the measurements, it
The PEC performance of Ti-doped R-Fe2O3 films was inves- is challenging to compare the measured photocurrent densities
tigated and optimized as a function of Ti precursor concentration reported by different research laboratories.15 In this regard,
and the number of DA cycles (Figure S12, Supporting Information). IPCE, which is independent from the specific light source, is a
Photocurrent densities of R-Fe2O3 films (prepared with 10 mM better parameter for comparison. We also calculated the photo-
FeCl3 solution and 10 DA cycles) increase drastically with the current density of R-Fe2O3 films by integrating their IPCE
addition of a small amount of Ti precursor. The photocurrent spectra with a standard AM 1.5G solar spectrum (ASTM
reaches a maximal value at 10% (atomic ratio of Ti:Fe) Ti pre- G-173-03), using the following equation
cursor concentration, and decreases gradually for higher precursor Z 650
concentration. By fixing the Ti precursor concentration at 10%, 1
I ¼ λIPCEðλÞEðλÞ dðλÞ
the PEC performance of the R-Fe2O3 films was further optimized 300 1240
by varying the number of DA cycles. The highest photocurrent
density was achieved at 14 DA cycles (Figure S12b, Supporting where E(λ) is the solar irradiance at a specific wavelength (λ),
Information). Vbias is the applied bias vs RHE, and IPCE(λ) is the obtained
Figure 4a compares the linear sweeps of optimized undoped photoresponse profile of the R-Fe2O3 samples at a specific
and Ti-doped R-Fe2O3 films. Both current-bias curves can be wavelength (λ) at 1.02 V vs RHE. The photocurrent density is
discerned into three regions with the increase of applied bias. At calculated to be 0.78 mA/cm2 for Ti-doped R-Fe2O3 and 0.075
small bias, the photocurrent density increases slowly and non- mA/cm2 for undoped R-Fe2O3 samples, which are smaller than
linearly, being indicative of a limited separation rate for the that obtained experimentally using a xenon lamp at the same
photogenerated electronhole pairs in R-Fe2O3 under a low applied bias. The discrepancy is believed to be due to the
electric field. As the applied bias is increased, the electronhole mismatch between the xenon lamp spectrum and the standard
pairs are separated more efficiently and the charge transport in solar spectrum.
the R-Fe2O3 film becomes the rate-limiting step. Therefore, the To elucidate the strong correlation between Ti-doping and
photocurrent density shows a linear variation with the applied the enhanced photocurrent and IPCE values, electrochemical
bias. The Ti-doped R-Fe2O3 film shows an earlier photocurrent impedance measurements and ultrafast spectroscopy were carried
onset (i.e., at a lower voltage) than the undoped sample, out. Capacitances were derived from the electrochemical impedance
3507 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

that the Faradaic charge transfer is the limiting step for the oxi-
dation process in the electrode surface (Figure S13, Supporting
Information). Under illumination (100 mW/cm2), two capacitive
arcs were observed in the Nyquist plot. The arcs observed in
the high (left) and the low (right) frequencies are correlated to
the charge transfer resistance and the mass transfer limitation,
respectively. The capacitive arcs obtained under illumination
have much smaller radii (i.e., smaller charge transfer resistance)
than those in the dark, indicating that the photoexcited carriers
increase the conductivity of the R-Fe2O3 film.
Ultrafast spectroscopy studies44 were carried out to probe the
fundamental charge carrier dynamics following photoexcitation
and to understand the possible effect of Ti-doping on the photo-
excited electron dynamics in R-Fe2O3 films. The normalized
ultrafast transient absorption profiles of undoped and Ti-doped
R-Fe2O3 films are shown on two different time scales in Figure 5b.
A pulse-width-limited rise of the signal is followed by a decay that
can only be fit to a multiple exponential. To achieve a satisfactory
fit result, at least three exponential functions need to be employed.
The fit-derived time constants for the undoped and Ti-doped
R-Fe2O3 films are 0.329, 2.225, and 70.84 ps and 0.336, 2.310,
and 70.55 ps, respectively. These time constants were convolved
with a Gaussian (FWHM 130 fs), representing the cross correla-
tion of the 440 nm pump (150 nJ/(pulse cm2), attenuated with
neutral density filters) and 620 nm probe pulses. The overall
charge carrier decay for both R-Fe2O3 films is very fast, without
measurable transient absorption beyond 300 ps. The very fast
Figure 5. (a) MottSchottky plots of the undoped and Ti-doped decay is likely due to a high density of electronic states in the
R-Fe2O3 films. Capacitances were derived from the electrochemical band gap caused by internal defects and/or surface defects. This
impedance obtained at each potential with 10 kHz frequency in the dark. indicates that the early time dynamics of the photogenerated
Inset: magnified MottSchottky plot of the Ti-doped R-Fe2O3 film.
charge carriers are dominated by the intrinsic properties of the
(b) Normalized ultrafast transient absorption decay profiles of undoped
and Ti-doped R-Fe2O3 films in 010 ps (inset) and the 0300 ps R-Fe2O3 and consistent with the decay profiles reported pre-
windows. They are fit simultaneously using a nonlinear least-squares viously for R-Fe2O3 nanoparticles.16,17 In comparison to the
fitting algorithm to a triple-exponential decay convolved with a Gaussian, undoped sample, the Ti-doped R-Fe2O3 film decays slightly
representing the cross correlation of the 440 nm pump and 620 nm faster on the shortest time scale (Figure 5b, inset). For time scales
probe pulses. The solid lines are fitted curves. above 4 ps, the absorption decay profile of the Ti-doped sample is
slightly higher than that of the undoped sample, indicating a
obtained at each potential with 10 kHz frequency in the dark. reduced electronhole recombination on that time scale. This
MottSchottky plots were generated from the capacitance might contribute to the enhanced photocurrent of the Ti-doped
values. Both samples show a positive slope in the MottSchottky sample observed in the PEC measurements.
plots, indicating that they are n-type semiconductors with In summary, we have demonstrated a facile DA method to
electrons as majority carriers (Figure 5a). The slopes determined fabricate highly photoactive undoped and Ti-doped R-Fe2O3
from the MottSchottky plots were used to estimate the carrier films on FTO substrates. The size of R-Fe2O3 particles and the
densities using the equation film thickness can be controlled by varying the Fe precursor
concentration and the number of DA cycles. In comparison to
Nd ¼ ð2=e0 εε0 Þ½dð1=C2 Þ=dV 1 undoped R-Fe2O3, the Ti-doped film shows enhanced photo-
current density and IPCE as a result of improved charge carrier
where e0 is the electron charge, ε the dielectric constant of density. The optimized Ti-doped R-Fe2O3 films showed the
R-Fe2O3, ε0 the permittivity of vacuum, Nd the dopant density, highest IPCE values and photocurrent density for R-Fe2O3
and V the potential applied at the electrode. With an ε value of photoanode at a low bias of 1.02 V vs RHE. Further improvement
80 for R-Fe2O3,20 the electron densities of the undoped and of PEC performance of these R-Fe2O3 films could potentially be
Ti-doped R-Fe2O3 films were calculated to be 2.7  1017 and accomplished by coupling them with efficient oxygen-evolving
1.36  1019 cm3, respectively. The Ti-doping thus led to an catalysts, such as Co2+-based compounds45 and iridium oxide,10
increase of 2 orders of magnitude in the carrier density of as well as improving the intrinsic electronic structure of R-Fe2O3
R-Fe2O3 films. The electron density of Ti-doped R-Fe2O3 is in order to achieve longer photoexcited charge carrier lifetimes.
consistent with reported values for Si-doped R-Fe2O3 structures
prepared by APCVD20 and is believed to be a major contributing ’ ASSOCIATED CONTENT
factor for the pronounced photocurrent density enhancement.
Moreover, electrochemical impedance spectroscopy was used to bS Supporting Information. Experimental details, optical
investigate the kinetics of the oxidation process at the electrode and SEM images, XPS, XRD, PEC, and impedance data. This
surface. The Nyquist plot collected for Ti-doped R-Fe2O3, in material is available free of charge via the Internet at http://
the dark at 1.02 V vs RHE, exhibits one capacitive arc, suggesting pubs.acs.org.
3508 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509
Nano Letters LETTER

’ AUTHOR INFORMATION (23) Sartoretti, C. J.; Alexander, B. D.; Solarska, R.; Rutkowska, W. A.;
Augustynski, J.; Cerny, R. J. Phys. Chem. B 2005, 109 (28), 13685–13692.
Corresponding Author (24) Lindgren, T.; Wang, H. L.; Beermann, N.; Vayssieres, L.;
*E-mail: yli@chemistry.ucsc.edu. Hagfeldt, A.; Lindquist, S. E. Sol. Energy Mater. Sol. Cells 2002, 71 (2),
231–243.
(25) Kleiman-Shwarsctein, A.; Hu, Y. S.; Forman, A. J.; Stucky,
’ ACKNOWLEDGMENT G. D.; McFarland, E. W. J. Phys. Chem. C 2008, 112 (40), 15900–15907.
(26) Le Formal, F.; Gratzel, M.; Sivula, K. Adv.Funct. Mater. 2010, 20 (7),
We thank Fang Qian for the helpful discussion. Y.L. acknowl- 1099–1107.
edges the financial support of this work in part by NSF (DMR- (27) Brillet, J.; Gratzel, M.; Sivula, K. Nano Lett. 2010, 10 (10),
0847786) and faculty startup funds granted by the University of 4155–4160.
California, Santa Cruz. J.Z.Z. thanks the BES Division of the U.S. (28) Mohapatra, S. K.; John, S. E.; Banerjee, S.; Misra, M. Chem.
DOE (DE-FG02-ER46232) for financial support. D.W. was Mater. 2009, 21 (14), 3048–3055.
partially supported by the W.M. Keck Center for Nano- and (29) Rao, P. M.; Zheng, X. L. Nano Lett. 2009, 9 (8), 3001–3006.
Optofluidics through a QB3 Fellowship. C.H. acknowledges (30) Massey, M. J.; Baier, U.; Merlin, R.; Weber, W. H. Phys. Rev. B
1990, 41, 7822–7827.
financial support by the U.S. Department of Energy, Energy
(31) Campbell, I. H.; Fauchet, P. M. Solid State Commun. 1986, 58
Efficiency and Renewable Energy (EERE), under subcontract (10), 739–741.
#RF-05-SHGR-004 of prime contract #DE-FG36-03GO13062 (32) Gratzel, M. Nature 2001, 414 (6861), 338–344.
and subcontract #NFH-8-88502-01 of prime contract #DE- (33) Wang, W.; Howe, J. Y.; Gu, B. H. J. Phys. Chem. C 2008, 112
AC36-08GO28308. (25), 9203–9208.
(34) Briggs, D.; Seah, M. P., Auger and X-Ray Photoelectron Spectroscopy:
Practical Surface Analysis; Wiley: New York, 1990; Vol. 1, Appendix 1.
’ REFERENCES (35) Moulder, J. F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D.
(1) Cesar, I.; Kay, A.; Martinez, J. A. G.; Gratzel, M. J. Am. Chem. Soc. Handbook of X-ray Photoelectron Spectroscopy. In Physical Electronics
2006, 128 (14), 4582–4583. Division, Perkin-Elmer Corporation: Eden Prairie, MN, 1992.
(2) Kay, A.; Cesar, I.; Gratzel, M. J. Am. Chem. Soc. 2006, 128 (49), (36) Wandelt, K. Surf. Sci. Rep. 1982, 2, 1–121.
15714–15721. (37) Mills, P.; Sullivan, J. L. J. Phys. D: Appl. Phys. 1983, 16, 723–732.
(3) Wang, G. M.; Yang, X. Y.; Qian, F.; Zhang, J. Z.; Li, Y. Nano Lett. (38) Fujii, T.; de Groot, F. M. F.; Sawatzky, G. A.; Voogt, F. C.;
2010, 10 (3), 1088–1092. Hibma, T.; Okada, K. Phys. Rev. B 1999, 59 (4), 3195–3202.
(4) Lin, Y.; Zhou, S.; Sheehan, S. W.; Wang, D. J. Am. Chem. Soc. (39) Peng, D. L.; Sumiyama, K.; Oku, M.; Li, D. X.; Suzuki, K. Phys.
2011, 133 (8), 2398–2401. Status Solidi A 1996, 157 (1), 139–152.
(5) Liu, R.; Lin, Y.; Chou, L.; Sheehan, S. W.; He, W.; Zhang, F.; (40) B€ar, M.; Weinhardt, L.; Pookpanratana, S.; Heske, C.; Nishiwaki,
Hou, J. M.; Wang, D. Angew. Chem., Int. Ed. 2011, 50 (2), 499–502. S.; Shafarman, W.; Fuchs, O.; Blum, M.; Yang, W.; Denlinger, J. D. Appl.
(6) Yang, X. Y.; Wolcott, A.; Wang, G. M.; Sobo, A.; Fitzmorris, Phys. Lett. 2008, 93, 244103.
R. C.; Qian, F.; Zhang, J. Z.; Li, Y. Nano Lett. 2009, 9 (6), 2331–2336. (41) Weinhardt, L.; Blum, M.; B€ar, M.; Heske, C.; Cole, B.; Marsen,
(7) Hensel, J.; Wang, G. M.; Li, Y.; Zhang, J. Z. Nano Lett. 2010, 10 (2), B.; Miller, E. L. J. Phys. Chem. C 2008, 112, 3078–3082.
478–483. (42) Zhang, M. L.; Luo, W. J.; Li, Z. S.; Yu, T.; Zou, Z. G. Appl. Phys.
(8) Solarska, R.; Krolikowska, A.; Augustynski, J. Angew. Chem., Int. Lett. 2010, 97 (4), 3.
Ed. 2010, 49 (43), 7980–7983. (43) Morin, F. J. Phys. Rev. 1951, 83 (5), 1005–1010.
(9) Su, J.; Feng, X. J.; Sloppy, J. D.; Guo, L.; Grimes, C. A. Nano Lett. (44) Newhouse, R. J.; Wang, H. N.; Hensel, J. K.; Wheeler, D. A.;
2011, 11 (1), 203–208. Zou, S. L.; Zhang, J. Z. J. Phys. Chem. Lett. 2011, 2 (3), 228–235.
(10) Tilley, S. D.; Cornuz, M.; Sivula, K.; Gratzel, M. Angew. Chem., (45) Kanan, M. W.; Nocera, D. G. Science 2008, 321, 1072–1075.
Int. Ed. 2010, 49 (36), 6405–6408.
(11) Qian, F.; Wang, G.; Li, Y. Nano Lett. 2010, 10, 4686–4691.
(12) Klahr, B. M.; Martinson, A. B. F.; Hamann, T. W. Langmuir
2011, 27, 461–468.
(13) Ling, Y.; Wang, G.; Wheeler, D. A.; Zhang, J. Z.; Li, Y. Nano
Lett. 2011, 11 (5), 2119–2125.
(14) Duret, A.; Gratzel, M. J. Phys. Chem. B 2005, 109 (36),
17184–17191.
(15) Murphy, A. B.; Barnes, P. R. F.; Randeniya, L. K.; Plumb, I. C.;
Grey, I. E.; Horne, M. D.; Glasscock, J. A. Int. J. Hydrogen Energy 2006, 31
(14), 1999–2017.
(16) Joly, A. G.; Williams, J. R.; Chambers, S. A.; Xiong, G.; Hess,
W. P.; Laman, D. M. J. Appl. Phys. 2006, 99 (5), 6.
(17) Cherepy, N. J.; Liston, D. B.; Lovejoy, J. A.; Deng, H. M.;
Zhang, J. Z. J. Phys. Chem. B 1998, 102 (5), 770–776.
(18) Khaselev, O.; Turner, J. A. Science 1998, 280 (5362), 425–427.
(19) Sivula, K.; Zboril, R.; Le Formal, F.; Robert, R.; Weidenkaff, A.;
Tucek, J.; Frydrych, J.; Gratzel, M. J. Am. Chem. Soc. 2010, 132 (21),
7436–7444.
(20) Cesar, I.; Sivula, K.; Kay, A.; Zboril, R.; Graetzel, M. J. Phys.
Chem. C 2009, 113 (2), 772–782.
(21) Khan, S. U. M.; Akikusa, J. J. Phys. Chem. B 1999, 103 (34),
7184–7189.
(22) Glasscock, J. A.; Barnes, P. R. F.; Plumb, I. C.; Savvides, N.
J. Phys. Chem. C 2007, 111 (44), 16477–16488.

3509 dx.doi.org/10.1021/nl202316j |Nano Lett. 2011, 11, 3503–3509

You might also like