Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/263808984

Hematite (α-Fe2O3) Photoanodes for the Photooxidation of Water

Article  in  ECS Transactions · May 2012


DOI: 10.1149/1.3702427

CITATIONS READS
5 370

4 authors, including:

Iris Herrmann-Geppert Peter Bogdanoff


Hochschule Mittweida Helmholtz-Zentrum Berlin
55 PUBLICATIONS   3,177 CITATIONS    126 PUBLICATIONS   9,033 CITATIONS   

SEE PROFILE SEE PROFILE

Sebastian Fiechter
Helmholtz-Zentrum Berlin
322 PUBLICATIONS   10,534 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Microgravity Materials Science View project

CleanSi View project

All content following this page was uploaded by Iris Herrmann-Geppert on 10 July 2014.

The user has requested enhancement of the downloaded file.


ECS Transactions, 41 (33) 201-212 (2012)
10.1149/1.3702427 © The Electrochemical Society

Hematite (α-Fe2O3) Photoanodes for the Photooxidation of Water

I. Herrmann-Geppert, P. Bogdanoff, L. Hepperle, and S. Fiechter

Institute for Solar Fuels, Helmholtz-Zentrum


Berlin für Materialien und Energie Hahn-Meitner-Platz 1, D-14109 Berlin, Germany

Nanostructured hematite films prepared by a sol-gel procedure


were evaluated for the photo-induced oxygen evolution reaction
(OER). Surface treatments such as plasma and post-annealing at
500 °C in oxygen gas flow were used in order to condition the
electrode surface. Electrochemical measurements of the plasma
treated samples revealed that the surface was reorganized in terms
of enhanced catalytic properties and reduced recombination centers
at the surface. Photocurrents in the milli-Ampere range were
yielded at 0.92 V(NHE) in 1 M KOH. In contrast to the plasma
samples, the post-annealing samples displayed higher bulk
crystallinity resulting in a more enhanced photovoltage. XRD and
TEM analysis confirmed the texturing of a beneficial highly
oriented hematite film (110) parallel to the back contact.
Nevertheless, lower photocurrents for this post-annealed sample
compared to the plasma one indicated that the catalytic centers
were not fully recovered at the surface.

Introduction

Based on Fujishima and Honda’s pioneering work (1), international research has focused
on the conversion of sunlight into chemical energy producing hydrogen as a clean and
renewable energy resource. Among other systems, photoelectrochemical cells were
developed consisting of monolithic semiconductor devices as photoabsorbers which are
combined with electrocatalysts for the electrochemical water splitting reaction. A more
direct way, however, is the development of so-called photocatalysts in which a
semiconducting material simultaneously functions as absorber as well as catalyst in
electrochemical reactions. Nevertheless, in all these systems the total efficiency is often
limited to the water oxidation kinetic at the anode. Therefore, appropriate
(photo)electrocatalysts are required to lower the activation barrier and to utilize the
yielded photovoltage for the reaction.

For decades, hematite has been discussed as promising catalyst for the photo-induced
oxygen evolution reaction (OER) due to its appropriate band gap of 2.2 eV, suitable
valence band position, high corrosion stability, non-toxicity and earth abundance (2 - 5).
Therefore, various preparation routes were tested to develop highly active hematite
photoanodes for water oxidation. However, the semiconductor is characterized by some
unfavorable physical parameters such as a low diffusion length (LD = 2 – 4 nm), low
electrical conductivity and a low mobility; thus, the transport of photogenerated holes to
the electrode surface is generally limited. Consequently, recombination of the electron-
hole pairs has to be overcome by increasing the diffusion length, by enhancing the

201
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

conductivity by carrier doping and by the passivation of recombination centers at the


surface and in the bulk. Hence, the implementation of nanotechnology for an optimized
architecture of the electrodes could potentially help to surmount this problem. Recent
progress in the development of the hematite photoanode has been achieved by the group
of Grätzel (6). They produced nanostructured hematite layers by chemical vapor
deposition. Finally, the combination of the appropriate hematite nanostructure with a co-
catalyst (IrO2) leads to a highly efficient electrode material. Nanostructured hematite can
also be synthesized by wet chemical methods, which offer the additional opportunity to
use dopants and deposit co-catalysts simultaneously. Although nanostructured hematite
grown by a wet-chemical process revealed equal structural parameters compared to layer
deposition from a gas phase reaction, a high photoactivity towards the OER was only
observed for the hematite produced from the latter technique until now (7).

This work is dedicated to further develop and improve a sol-gel method to grow
Fe2O3-based photoanodes. For the preparation of efficient Fe2O3-based photoanodes,
recombination of the generated charge carriers in the electrode bulk and on the surface
had to be hindered; thus, plasma treatment of the surface as well as post-annealing of the
hematite films were performed. Cyclic voltammogram and potentiostatic measurements
were performed to evaluate the (photo)electrochemical activity. These results are
discussed in correlation with the structural changes detected by X-ray diffraction and
transmission electron microscopy analysis.

Experimental Methods

Sol-gel Preparation

0.01 mol Fe(NO3)3 was added to a solution containing 0.05 mol citric acid and
0.1 mol ethylene glycol at 80 °C under reflux. The resulting yellow solution was then
heated at 130 °C for two hours. After cooling down, the brown sol-gel (resin consistency)
was diluted in ethylene glycol in a 4:3 ratio and constantly stirred for thirty minutes.
Previously cleaned fluorine-doped tin oxide glasses (FTO) were coated with the sol-gel
solution by dip-coating at 2.27 cm/min. The coated glasses were calcined in a split-hinge
furnace equipped with a quartz tube under air atmosphere with a slow heating rate
(1 K/min) up to 500 °C and held at this temperature for two hours. The dip-coating
procedure and subsequent calcination were repeated twice in order to reach a layer
thickness of 300 to 500 nm.

Surface Treatments

Plasma treatment was performed in a plasma etcher by Diener electronics at the


Optotransmitter Umweltschutz Technologie e.V. (OUT e.V. Berlin). The films were
placed in the reaction chamber with their conducting side face upwards. The chamber
was evacuated and an oxygen partial pressure of 0.2 mbar was reached. For the
optimization of the operation parameters, the oxygen gas flow was varied between 5 and
20 sccm. Furthermore the treatment time duration was varied between five and ten
minutes. For the post-annealing in oxygen gas flow, the already prepared hematite films
were placed in a quartz tube of a split-hinge furnace again. The tube was connected to the
oxygen gas supply so that a flowing oxygen gas atmosphere was established within the

202
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

tube. The temperature was increased up to 500 °C with a heating rate of 1 K/min and held
at this temperature for two hours. Samples were undergone once (18 h), twice (36 h) and
three times by this treatment (54 h).

Electrochemical Characterization

Electrochemical measurements were performed in a three-electrode compartment


controlled by a EG&G potentiostat (Type 273A). Thereby, the films were contacted as a
working electrode by a gold wire via the FTO glass outside of the electrolyte. A platinum
wire was used as counter electrode and a Hg/HgO reference electrode was utilised in 1 M
KOH (+ 0.32 V vs. NHE). Iron oxide samples were cycled with 10 mV/s from 0.3 to
0.7 V (NHE) in 1 M KOH until the cyclic voltammogram (CV) curve showed steady
state characteristics. CVs under dark conditions and illumination were performed.

For the evaluation of the photoelectrochemical properties of the films, potentiostatic


measurements were performed at a potential of 0.92 V(NHE) in the alkaline medium.
After the potentiostatic curve showed steady state behaviour, light (400 mWcm-2 of a
tungsten halogen lamp from Xenophot) with intermittent pulses for 30 s was focussed on
a sample section with an area of 0.3117 cm2. For the evaluation of the electrochemical
activity the potential was read, which is needed to obtain a current density of
0.01 mA/cm2.

Structural Characterization

X-ray diffraction (XRD) analysis. XRD measurements of the catalysts were carried
out employing a Bruker diffractometer D8 Advance in Bragg-Brentano Θ - 2Θ geometry
with Göbel mirror using Cu-Kα-radiation. A silicon disc was used as a sample holder.
Phase determination was done according to the Powder Diffraction File (PDF-2)
maintained by the International Centre for Diffraction Data (ICDD) in its 2002-version.
Crystal particle size G (in Å) was determined by the Scherrer equation (8):

G = K*λ / B*cosθ [1]


with K = 0.5, λ = 1.5406 Å and

B = (K12 - K22)0.5 [2]


with K1 – FWHM of the sample and K2 = 0.0033 (FWHM of the Bruker diffractometer)

Scanning and transmission electron microscopy (SEM and TEM) and UV/Vis
spectroscopy. SEM images of the hematite films were taken using a LEO 1530 scanning
electron microscope at 5 kV. The prepared layers were investigated by employing a
Philips CM12 transmission microscope. Optical analysis was performed using a Lab950
UV-Vis spectrometer (by Perkin Ellmer) equipped with a 150 mm InGaAs integrating
sphere. A Tauc plot was applied to determine direct band edge in accordance to Tauc et
al. (9). The relation between the absorption coefficient (α) and the energy of the incident
light (h ν) was given by

(α *h ν)n = A (h ν – Eg) [3]


where h is the Planck constant, Eg the band gap of the material and n equals 2 for the
direct transition.

203
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

Results and Discussion

Figure 1 presents the obtained films after the calcination of the dip-coated sol-gel on
fluorine-doped tin oxide glasses (FTO). SEM analysis of the red-orange films revealed
the typical morphology of the sol-gel product from the Pecchini method (10). The
particles were rod-type shaped, ca. 100 nm long and 50 nm thick.

iron oxide

FTO

200 nm

Figure 1. (left) Photograph and (right) SEM image of the hematite photoanode.
9
80 1.5x10

60
573nm 1.0x10
9
Absorbance [%]

2
(α h ν)
40

8
5.0x10

20

0.0
300 400 500 600 700 800 900 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0
Wave length [nm] Energy [eV]

Figure 2. (left) UV-Vis measurement of the hematite film and (right) the Tauc plot for
the direct band gap.

UV-Vis measurement revealed a pronounced absorption edge in the red region from
500 to 600 nm (see Fig. 2). The Tauc plot of the direct transition was calculated. From
the extrapolation area a direct band gap at 2.17 eV was determined for the investigated

204
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

hematite sample. These values agreed with the literature data of hematite synthesized via
diverse preparation routes (11 - 14). XRD analysis of the films confirmed that hematite
was formed after the calcination steps (see Fig. 3). The X-ray diffractogram of the films
revealed the presence of pure hematite (PDF 85-0987) between the dominating
reflections of the FTO substrate. The crystal size of the hematite films calculated from
the 110 reflection was ca. 35 nm. Moreover, the hematite was textured with 001
preferential orientation. Iordanova et al. demonstrated that the conductivity of the basal
plane (001) was up to three orders of magnitude higher than the orthogonal direction.
Therefore, such orientation of the grains in our films should favor the transport of charges
to the catalytic reactive surfaces (110) (15).

(110)

(121)
(220)

FTO

Hematite

10 15 20 25 30 35 40 45 50 55 60 65
Diffraction angle 2Θ [°]

Figure 3. X-Ray diffractogram of the substrate (fluorine-doped tin oxide – red) and of
the sol-gel prepared hematite layer with the signal pattern of hematite (PDF 085-0987 –
blue).
-4
2.0x10

-4
1.5x10
Current density [A/cm ]
2

-4
1.0x10

-5
5.0x10

0.0
0.5 0.6 0.7 0.8 0.9 1.0 1.1
Potential E(NHE) [V]

Figure 4. CV measurements of the hematite layer prepared from the sol-gel procedure in
1 M KOH with 10 mV/s (0.3117 cm2) under dark conditions (black) and under
illumination (blue).

205
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

Nevertheless, only low electrochemical activity of the hematite films towards the
oxygen evolution reaction (OER) was observed in 1 M KOH. The measured cyclic
voltammograms (CV) are presented in figure 4. For the evaluation of the electrochemical
activity, the potential at 0.01 mA/cm2 was consulted. Thereby, the sol-gel hematite films
yielded 0.01 mA/cm2 at a potential of ca. 955 mV(NHE), which implies a high
overvoltage compared to the thermodynamical redox potential of the OER at pH = 14
(390 mV(NHE)). Under illumination, the potential was shifted to ca. 900 mV(NHE).
Apparently, only a small photovoltage is generated by the illumination which can
contribute to the water oxidation reaction.

In a preliminary investigation (16), XPS analysis showed that hematite prepared from
our sol-gel process exhibited carbon-based residuals on the surface. Evidently, residues
of the organic-based polymer remained in the interstices of the nano-sized hematite
particles; thus protected against the oxidation to CO2 during the heat treatment step in air.
These adsorbents could potentially block the catalytic centers and limit the OER kinetic,
which results in a low electrochemical activity under dark conditions and under
illumination. In order to solve this problem, surface treatments were required to remove
these carbon-based species.

Plasma Treatment of the Sol-gel Prepared Hematite Layers in Oxygen-rich Plasma

Plasma etching is a well-established method for cleaning surfaces. For the removal of
the carbon-based species on the surface the hematite films were treated in oxygen-rich
plasma as described in the experimental section. CVs of the plasma treated sample are
presented in figure 5.
-3
2.5x10
Hematite layer
after plasma treatment under
2.0x10
-3 dark conditions
illumination Current density [A/cm ]
2

-3
1.5x10

-3
1.0x10

-4
5.0x10

0.0
0.6 0.8 1.0 1.2
Potential E(NHE) [V]

Figure 5. CV measurements of plasma treated hematite layer (20 sccm O2 and 5 min) in
1 M KOH with 10 mV/s (0.3117 cm2) under dark conditions (grey) and under
illumination (blue).

Under dark conditions the potential at 0.01 mA/cm2 of the anodic current was shifted
by about 50 mV to negative potentials compared to the non-treated sample (ca.
900 mV(NHE)). Evidently, the plasma treated sample revealed similar catalytic
properties for the dark reaction like the untreated one. But under illumination the

206
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

potential at 0.01 mA/cm2 was shifted to 550 mV(NHE) (obtained by extrapolation) which
revealed an enhancement of ca. 350 mV compared to the untreated film. This enhanced
catalytic photoactivity indicated the formation of a photovoltage and an increased number
of photogenerated charge carriers, which participated in the water oxidation reaction. The
improvement of the catalytic properties for the photo-induced oxidation became more
noticeable in the following potentiostatic measurements (see Fig. 6).
Light on Light off before O2 plasma treatment
-4
4.0x10 after O2 plasma treatment

Current density at 0.92V(NHE) [A/cm ]


2
-4
3.0x10

-4
2.0x10

-4
1.0x10
Light on Light off

0.0
0 50 100 150 200
Time [s]

Figure 6. Potentiostatic measurements of hematite layer before (blue) and after the
plasma treatment (20 sccm O2 and 5 min) (red) in 1 M KOH at 0.92 V(NHE)
(0.3117 cm2).

While the untreated sample revealed photocurrents in the micro-Ampere range at


0.92 V(NHE), an exceeding increase of the photocurrent was yielded. The photocurrent
was increased in the milli-Ampere range due to the plasma treatment. The increased
photocurrent points to a successful removal of the addressed carbon-based species. For
the investigation of the structural change of the electrode surface XPS analysis is required
which will be presented in a forthcoming work. The observation of an improved
photovoltage also points to a reorganization of the semiconductor surface so that defects
and crystal imperfections were eliminated. This seemed to be likely by a plasma
treatment. Such a surface recovery led to a reduced amount of recombination and resulted
in a higher photovoltage and thereby also in a higher photocurrent. Optimization of the
operation parameters during the plasma process was performed by varying the gas flow
and the time of treatment. For the evaluation of the effect the yielded photocurrents at
0.92 V(NHE) from potentiostatic measurements were evaluated and presented in figure 7.

Thereby, it was noted that with increasing plasma time and oxygen gas flow the
photocurrent was increased. The presence of an increased number of reactive oxygen
species in contact with the hematite surface and their longer dwell time led to a
successful removal of the carbon-based species and reorganisation of the surface. The
sample treated for ten minutes with 20 sccm oxygen exhibited again lower photocurrent.
An explanation for this behavior is not well understood yet. Longer attack of the oxygen

207
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

species in plasma might lead to a destruction and abrasion of the hematite layer. In
further experiments the reversal of this surface modification will be investigated.
Oxygen Gas Flow
20sccm

Photocurrent at 0.92V(NHE) [A/cm ]


10sccm

2
-4
3.0x10 5sccm

-4
2.0x10

-4
1.0x10

0.0
0 2 4 6 8 10
Time of Plasma Treatment [min]

Figure 7. Photocurrent at 0.92 V(NHE) (1 M KOH) of the plasma treated samples with
varied oxygen gas flow (blue circle – 20 sccm, black square – 10 sccm and red triangle –
5 sccm) versus their treatment time.
-5
4.0x10
before O2 heat treatment
dark
-5 under light
3.0x10 after O2 heat treatment

Current density [A/cm ]


2
at 500°C
under light
-5
2.0x10

-5
1.0x10

0.0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Potential E(NHE) [V]

Figure 8. CV measurements of hematite layer in 1 M KOH with 10 mV/s (0.117 cm2)


under illumination after post-treatment in oxygen atmosphere at 500 °C (blue solid line)
and before this treatment (dark – dotted black line, under illumination – dotted blue line).

Post Heat Treatment of the Sol-gel Prepared Hematite Layers in Oxygen at 500 °C

In the above described plasma treatment of the hematite layers, a reorganization of


the surface was obtained. In addition to the recombination centers on the surface,
crystalline imperfections in the bulk were also considered as eventual traps and
recombination centers for the photogenerated charge carriers. In to order to improve the
crystallinity of the hematite films, the sol-gel prepared hematite layers were post-treated
in oxygen atmosphere at 500 °C. In Figure 8 the obtained CVs of the photo-

208
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

electrochemical measurements are presented. Compared to the sample treated in oxygen-


rich plasma, the onset potential of the anodic current under illumination was significantly
shifted to more negative potentials to 370 mV (NHE).

A higher photovoltage was yielded by the heat-treatment of the hematite layers


compared to ones acquired from the plasma treatment. These results indicated a reduced
recombination rate. While in the plasma treatment only the surface was attacked by the
reactive gas species, the complete bulk was affected in the heat treatment.
Consequentially, the reduced recombination results from a healing of crystalline
imperfections in the inner part of the crystal. However, compared to the plasma samples
the current densities and the slope of the anodic photocurrent was lowered which
indicated a poor catalytic activity of the surface. Hence, only a potential of
890 mV(NHE) under illumination is yielded for the treated hematite layer at
0.01 mA/cm2.

This observation was mirrored in the potentiostatic measurement of the sample (see
Fig. 9). While the plasma-treated sample yielded a photocurrent in the milli-Ampere
range, the photocurrent was indeed increased due to the post-treatment in oxygen
atmosphere but lower than for the plasma treated sample.

Light on Light off after O2 heat treatment


-5
8.0x10 before O2 heat treatment

Current density at 0.92V(NHE) [A/cm ]


2
-5
6.0x10

-5
4.0x10

-5
2.0x10

0.0

0 50 100 150 200 250


Time [s]

Figure 9. Potentiostatic measurement of hematite layer before (blue) and after the second
post-treatment (500 °C in O2 atmosphere) (red) in 1 M KOH at 0.92 V(NHE)
(0.3117cm2).

These results led us to conclude that the semiconductor properties in the bulk were
improved due to the post-heat treatment step at 500 °C in oxygen gas flow. The reduction
of defects in the lattice which can act as recombination centers led to a higher
photovoltage. Despite of this enhancement, however, the post-heat treated sample
displayed lower photocurrents compared with the plasma treated samples. This effect is
caused by the different catalytic activity of the electrode/electrolyte interface. Apparently,
the plasma treatment forms a chemical structure at the hematite crystal surface which is
more favorable for the photo-induced OER than at the post-heat treated modification.
Thereby, it was assumed that the reaction mechanism differs from the photo-induced and

209
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

dark reaction. It was noted that the catalytic properties of the hematite layer has been
significantly enhanced under illumination after the plasma treatment whereas no
noteworthy change can be observed for the dark reaction. While the dark current is
caused from a strong bending of the valence band under high positive bias potentials,
surface states within the band gap could be responsible for the photoreaction.

before O2
140
(110) heat treatment
after 18h
120 O2 heat treatment
after 54h O2
100 heat treatment

80

60

40

34.0 34.5 35.0 35.5 36.0 36.5 37.0


Diffraction angle 2Θ [°]

Figure 10. Detailed XRD analysis of the (110) hematite reflex of the untreated hematite
sample (grey) and O2 heat treated ones (blue and red).

60 nm

200 nm

Figure 11. Cross-sectional TEM analysis (left) of the untreated hematite film and (right)
of the O2 heat treated hematite film.

Under illumination less photovoltage but an enhanced catalytic efficiency lead to the
observed high photocurrent the plasma treated samples. In contrast to that, a high
photovoltage was generated for the heat treated sample but the catalytic properties on the
surface were less improved. This hypothesis was confirmed by the structural
characterization using XRD and TEM analysis. In the X-ray diffractogram (see Figure

210
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

10) of the O2 heat treated samples compared to the untreated one an increase of the (110)
hematite reflection was observed while the intensity of the other investigated reflections
remained constant.

This results indicated a texture of the (001) crystal direction which was favorable for
an increased mobility of the charge carriers. In accordance to the XRD, TEM analysis of
the O2 heat treated sample showed that the film consisted of more enlarged crystals
compared to the untreated reference (see Fig. 11). Apparently the original nanostructured
hematite crystals aggregated so that the nanostructure seemed to be lost due to the heat
treatment. Nevertheless this was the basis for improved semiconductor properties. In
figure 11, it is also noted that the particles are rod-type arranged in the layer partially. As
mentioned above in the XRD analysis, a texture of the (001) crystal direction is identified
after the heat treatment. This was confirmed in the TEM analysis in which turning-up of
the crystals was visualized. Finally, it was confirmed that the crystal properties were
improved so that the generation and the transport of charge carriers were benefited.

Conclusion

The here presented sol-gel technique leads to nanostructured hematite layers for the
photo-induced water oxidation. In the surface analysis, carbon-based species were
detected which origin from the polymer network of the sol-gel after the calcination step.
They might inhibit the catalytic centers on the electrode surface for the oxygen evolution
reaction. Thus, surface treatments were proven to enhance the catalytic efficiency of the
photoanode.

By applying oxygen-rich plasma on the hematite film the photocurrent was increased
from the µA- to the milli-Ampere range (at 0.92 V(NHE) in 1 M KOH). Nevertheless, a
more improved onset potential of the photocurrent was obtained if the hematite layers
were heat treated at 500 °C in an oxygen gas flow. But lower photocurrents were gained
by this post-treatment. Apparently, the surface was recovered by the applied plasma
treatment so that more catalytic centers were able to participate in the water oxidation
reaction. Furthermore the reorganization led to reduced recombination centers on the
surface. The enhanced photovoltage obtained by the heat treatment in an oxygen gas flow
at 500 °C was related to the improved crystallinity of the samples. Nevertheless, this
post-annealed electrode only displayed poor catalytic activity in the OER. It is supposed
that such beneficial surface modification like in the plasma treatment is not obtained in
this case. In order to improve the OER of the heat treated samples, a subsequent plasma
treatment could be a promising preparation step to enable both improved bulk
crystallinity for high utilization of the photovoltage and a surface recovery for enhanced
catalytic centers.

Acknowledgments

Financial support by a grant of BMBF under contract # 03SF0353A "H2-NanoSolar"


is gratefully acknowledged. The authors would like to thank U. Bloeck for the cross-
section preparation of the hematite layers for the TEM analysis.

211
Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp
ECS Transactions, 41 (33) 201-212 (2012)

References

1. A. Fujishima and K. Honda, Nature, 238, 37 (1972).


2. K. Sivula and F. Le Formal, ChemSusChem, 4, 432 (2011).
3. J. A. Glasscock and P. R. F. Barnes, J. Phys. Chem. C, 111, 16477 (2007).
4. N. Beermann and L. Vayssieres, J. Electrochem. Soc., 147, 2456 (2000).
5. R. Shinar and J.H. Kennedy, Solar Energy Materials, 6, 323 (1982).
6. S.D. Tiley and M. Cornuz, Angewandte Chemie Int. Edition, 49, 1 (2010).
7. K. Sivula and R. Zboril, J. Am. Chem. Soc., 132, 7436 (2010).
8. A. L. Patterson, Physical Review (American Physical Society), 56, 978(1939).
9. J. Tauc and R. Grigorovic, Phys. Stat. sol. 15, 627 (1966).
10. M. P. Pechini, US Patent No. 3.330.697, (1967).
11. E. L. Miller and D. Paluselli, Thin Solid Films, 466, 307 (2004).
12. N. Özer and F. Tepehan, Solar Energy Materials and Solar Cells, 56, 141 (1999).
13. S. U. M. Khan and J. Akikusa, J Phys Chem B, 103, 7184 (1999).
14. Z. Zhang and M. F. Hossain, Applied Catalysis B: Environmental, 95, 423 (2010).
15. N. Iordanova and M. Dupuis, The Journal of Chemical Physics, 122, 144305 (2005).
16. I. Herrmann-Geppert and P. Bogdanoff, Applied Physics Letter, (in preparation).

212
View publication stats Downloaded 21 May 2012 to 134.30.13.105. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

You might also like