Вы находитесь на странице: 1из 10

www.advenergymat.

de
www.MaterialsViews.com

FULL PAPER

Creating Highly Active Atomic Layer Deposited NiO


Electrocatalysts for the Oxygen Evolution Reaction
Katie L. Nardi, Nuoya Yang, Colin F. Dickens, Alaina L. Strickler, and Stacey F. Bent*
Atomic layer deposition (ALD) provides a promising route for depositing uniform thin coatings of electrocatalysts useful in many technologies, including
the splitting of water. For materials such as NiOx that readily form hydrous
oxides, however, the smooth, compact films deposited by ALD may result
in higher overpotentials due to low catalyst surface area compared to other
deposition methods. Here, the use of ALDNiO thin films as oxygen evolution reaction (OER) electrocatalysts is explored. Thin films of crystalline ALD
NiO are deposited and OER activity is tested using cyclic voltammetry (CV).
Fe incorporated from the electrolyte can increase the activity of NiO, and it is
shown that the turnover frequency (TOF) increases tenfold by going from an
Fe-poor to Fe-rich KOH electrolyte. Applying a potential exfoliates the NiO,
increasing the number of electrochemically accessible Ni sites. Interestingly,
by X-ray photoelectron spectroscopy (XPS) and CV, it is found that an Fe-rich
electrolyte reduces the amount of restructuring and oxidation is found. It is
shown that a high surface area, high TOF catalyst may be created by using a
two-step process in which the sample is sequentially conditioned in Fe-poor
then Fe-rich KOH. This work highlights the importance of pretreatment on
catalytic activity for compact NiO films deposited by ALD.

1. Introduction
Atomic layer deposition (ALD) offers many interesting
possibilities for the engineering of devices for energy conversion and storage.[13] Consisting of a series of self-limited, saturated reactions, ALD offers an unrivaled ability to deposit thin,
conformal films with fine control over film thickness and uniformity.[46] These advantages have been successfully leveraged
to begin to address some of the most challenging problems for
the photoelectrochemical (PEC) splitting of water, a process in
which water is converted to oxygen and hydrogen that can later
be used as fuel.[7] For example, ALD has been used to deposit

Dr. K. L. Nardi,[+] C. F. Dickens, A. L. Strickler,


Prof. S. F. Bent
Department of Chemical Engineering
Stanford University
Stanford, CA 94305-4125, USA
E-mail: sbent@stanford.edu
N. Yang
Department of Materials Science and Engineering
Stanford University
Stanford, CA 94305-4125, USA
[+]ne Katie L. Pickrahn

DOI: 10.1002/aenm.201500412

Adv. Energy Mater. 2015, 5, 1500412

high-quality, pinhole-free films that protect underlying light absorbers from corrosion under the harsh operating conditions
of water splitting.[811] ALD has also been
used to apply both conformal and nanoparticulate catalysts with low parasitic
light absorption to nanostructured light
absorbers.[1215] The importance of ALD in
the development of PEC devices is likely
to continue to grow as nanostructuring
appears to be a promising way to increase
efficiency in these devices.[16,17]
One of the main efficiency losses in the
splitting of water occurs in the driving of
the oxygen evolution reaction (OER). OER
is a four-electron process, which evolves
oxygen in alkaline conditions as shown
below
4OH O2 + 2H2 O + 4e

(1)

Unfortunately, this reaction is often associated with sluggish kinetics. Catalysts that
drive OER in alkaline conditions include
binary compounds such as RuOx, IrOx, PtOx, CoOx, MnOx;
mixtures of these binary compounds; and perovskites.[1720]
ALD processes for many of the binary OER catalysts already
exist and show good OER activity,[21] although the activity may
be sensitive to the preparation of the film.[22]
NiFe oxides are particularly promising OER electrocatalysts
in alkaline conditions. NiFe oxides have one of the lowest
reported overpotentials for OER in alkaline conditions of 0.35
V to obtain a current density of 10 mA cm2,[19 ] and devices
with >12% solar-to-hydrogen efficiency have been created using
these catalysts.[19,23] Until recently, undoped NiOx was believed
to be a good catalyst for OER, with -NiOOH as the most active
phase, and the addition of Fe was thought to mainly improve
the conductivity of the film.[2426] Trotochaud and Boettcher
recently showed that trace amounts of Fe incorporated from the
electrolyte are largely responsible for the high activity of NiOx,
and that these effects go beyond simply the increasing of the
conductivity.[27] It has been suggested that Fe has an electronic
effect on the NiO bond, favorably altering the bonding of intermediates,[28] and more recently it has been suggested that Fe
sites in a NiOx matrix serve as the active sites.[29] Calculations
have pointed to -NiOOH with trace Fe as having the lowest
overpotential among the NiFe oxides, followed by NiFe2O4.[30]
Interestingly, under the belief that NiOx alone was active,
numerous studies have used Ni and its oxides in highly active
devices, likely unintentionally doping it with Fe in the testing

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(1 of 10) 1500412

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com

process.[10,31] Overall, it appears that the turnover frequency


(TOF) of NiOx can readily be enhanced with trace amounts
of Fe.
A high surface area can also positively impact the geometric
activity of electrocatalysts.[26,32] This is a particularly important
consideration in hydrous catalysts, whose surface area evolves
with testing and is highly dependent on the experimental conditions.[20] Unfortunately, in water splitting applications, the formation of high surface area films on light absorbers is often in
conflict with the need for conductive, stabilizing catalysts with
minimal parasitic light absorption for which dense, thin films
may be advantageous. In principle, the ability of ALD to conformally coat nanostructured substrates provides an opportunity
to achieve both desired characteristics in concert: dense, stable
thin films together with high surface area. However, many
ALD processes for transition metal oxides catalysts, particularly
hydrous catalysts, do not deposit a phase that is stable under
OER-relevant conditions, resulting in catalysts that undergo
oxidation and hydration under OER. For example, despite the
existence of many precursors for this process, most NiOx ALD
processes in literature deposit NiO,[3338] whereas under OER
conditions, NiOOH is likely to be the most relevant phase. Conversion into surface hydrates may alter the activity of the NiOx
thin film, and it is important to understand this effect.
Here, we study ALD-grown NiO thin films as OER electrocatalysts, increasing the activity of the catalysts through
the incorporation of Fe from the electrolyte. Although NiFe
oxides can be deposited directly by ALD,[39] the deposition of
ternary oxides is often complicated. Because literature points
to Fe being necessary only on the surface sites for thin films,
and because Fe is readily incorporated from the electrolyte, we
explore a two-step route to synthesize highly active NiO-based
OER electrocatalysts using electrolyte conditioning. We first
deposit NiO by ALD and then add Fe in situ through the electrolyte. We characterize its activity for OER in both trace Fe and
Fe-saturated 0.1 M KOH electrolytes. A maximum concentration
of 1:9 Fe:Ni can be achieved using the Fe-saturated electrolyte
conditioning. We find that the Fe-saturated electrolyte significantly increases the geometric activity of the films, as expected,
but that the Fe-saturated electrolyte also reduces the number of
electrochemically accessible redox sites within the film. X-ray
photoelectron spectroscopy (XPS) shows that samples cycled in
an Fe-saturated solution are less hydrated than samples tested
in trace Fe electrolyte. With this knowledge, we develop a route
to enhancing the geometric activity through conditioning with
different electrolytes. Similar routes may be applicable to other
hydrous metal oxide electrocatalysts.

2. Experimental Section
Deposition of NiO thin films was performed in a custombuilt ALD reactor. The reactor has a showerhead configuration, with precursors introduced through an opening above
the 4 in. substrate holder and removed through a pump line
located beneath the substrate holder. This design assists in
the uniformity of the deposited films. N2 was used as the carrier gas and regulated by mass flow controllers. Nickelocene
(Sigma-Aldrich, Ni(Cp)2) was used as the reactant and ozone,
1500412 (2 of 10)

wileyonlinelibrary.com

O3, was used as the counter-reactant. The stage temperature was maintained at 275 C, and the Ni(Cp)2 was maintained at 87 C. O3 was generated with an IN USA OG 5000
Series ozone generator at a concentration between 160 and
180 g Nm3.
Fluorine-doped tin oxide (FTO) coated glass was used as a
conductive substrate for the electrochemistry measurements
(Hartford Glass, TEC 8). The FTO was rinsed with acetone and
then sonicated for 20 min in both ethanol and deionized water.
Immediately prior to use, samples were O3-cleaned for 15 min.
Single-side polished n-doped Si(100) wafers (WRS Materials)
with an 1.5 nm thick native oxide were placed in the reactor
during each ALD run to monitor deposition.
Film thickness was characterized using ellipsometry
(Woolam -SE). Measurements at 65, 70, and 75 were taken
in a minimum of four different locations, and the average film
thickness was calculated. For saturation curves, film thickness
was measured after 75 ALD cycles. Grazing incidence X-ray diffraction (GIXRD) was done on a PANalytical XPert Pro Materials Research Diffractometer using CuK radiation and an
omega angle of 3. XPS measurements were completed on a
PHI VersaProbe Scanning XPS Microprobe with Al(K) radiation (1486 eV). Scanning electron microscopy was performed
on a FEI Magellan 400 XHR scanning electron microscope
(SEM).
The catalytic activity of the NiO films for OER was investigated using cyclic voltammetry (CV). A Biologic SP-200 potentiostat controlled the potential and measured the current and
EC-Lab was used to monitor the experiments. Films were
placed in a custom-made Teflon compression cell that exposed
0.5 cm2 of the sample. Electrical contact with the NiO/FTO film
was made with conductive copper tape. A coiled Pt wire served
as the counter electrode and an Ag/AgCl electrode was used as
a reference electrode. The electrolyte was continuously purged
with N2. Equipment was cleaned in MilliQ water (18.2 M cm)
and dried before use.
Two electrolytes were utilized for this work. One consisted
of 0.1 M KOH (Fischer Scientific, Reagent Grade). The solution was made from MilliQ water (18.2 M cm) and stored in
polyethylene containers to reduce contamination. The second
consisted of a 0.1 M KOH solution saturated with Fe(NO3)3. To
make the Fe-saturated electrolyte, 0.03 g of Fe(NO3)3 (Sigma
Aldrich) was added to 170 mL of 0.1 M KOH. In basic conditions, most of the Fe(NO3)3 precipitates out as iron hydroxide.
The iron oxide precipitate was allowed to sink to the bottom
and only the supernatant was used. Fresh electrolyte solutions
were made every 2 d.
Catalysts were preconditioned before the measurements
using a series of CVs and chronoamperometry holds to systematically activate the catalysts while characterizing them. First,
catalysts were soaked for 5 min in the electrolyte as N2 purged
the electrolyte. The potential was then swept between 1.1 and
1.8 VRHE two times to initially characterize the sample. A CV
over this voltage sweep will be referred to as our standard CV
from this point on. The catalyst was then held at 0.85 VRHE
for 30 min, a potential below the Ni2+/Ni3+ redox feature at
which no reaction occurs, to allow equilibration in the electrolyte. We explored systematic variation of this potential, but the
pretreatment was not observed to be sensitive to the potential of

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1500412

www.advenergymat.de
www.MaterialsViews.com

FULL PAPER
Figure 1. Characterization of the Ni(Cp)2/O3 ALD system. Saturation curves of thickness versus a) Ni(Cp)2 and b) O3 pulse show that the system takes
8 s for Ni(Cp)2 and 3 s for O3 to saturate. The thickness is measured after 75 ALD cycles. c) Growth curve for the Ni(Cp)2/O3 ALD under saturated
conditions. The growth rate is 0.63 cycle1 after the initial nucleation period. d) XRD of 18 nm thin film. The indexed peaks are for NiO.

the hold. Two additional CVs were taken, and then the catalyst
was held for 10 min at 1.65 VRHE to investigate changes under
OER-relevant potentials. The subsequent CVs are reported in
this paper, and are generally similar to the second set of CVs
(see the Supporting Information). Measurements were iRcompensated at 85%. TOFs were calculated as the moles of O2
evolved per mole of electrochemically accessible Ni per second
at different overpotentials using the area under the redox curve
in the cathodic-going direction to estimate the moles of electrochemically active Ni sites. It was assumed that each charge
passed is equivalent to one atom of Ni, which is likely an overestimate of the actual number of Ni sites, but can serve to
define a lower bound for the TOF.[40] Standard CVs were taken
at a scan rate of 20 mV s1 and measurements for the Tafel
slope were taken at a scan rate of 1 mV s1.

3. Results and Discussion


NiO deposited by ALD was explored as the initial starting phase
of the OER catalyst, using Ni(Cp)2 as the reactant and O3 as the
counter-reactant. Because of discrepancies in previous reports
of this NiO ALD process, we performed a systematic characterization of the ALD deposition.[37,41,42] NiO was deposited at a
stage temperature of 275 C and a Ni(Cp)2 bubbler temperature
of 8590 C. Under these conditions, measurements of film
thickness after 75 cycles for different pulse lengths showed
that saturation was achieved after 8 s for the Ni(Cp)2 pulse
and 3 s for the O3 pulse as seen in Figure 1a,b. These are rela-

Adv. Energy Mater. 2015, 5, 1500412

tively long saturation times compared to other processes performed in this reactor,[43,44] perhaps reflecting slow kinetics of
the Ni(Cp)2 on the surface. Bachmann et al. also used a relatively long Ni(Cp)2 pulse of 2 s compared to an O3 pulse of just
0.2 s[41] and Lu et al. used a Ni(Cp)2 pulse time of 7 s for their
process.[37] The thickness measured as a function of number of
cycles under saturation conditions is shown in Figure 1c. The
steady state growth rate is 0.63 cycle1 and there is a nucleation delay of 20 cycles on O3-cleaned Si. The crystallinity of
the film was probed by GIXRD on an 18 nm thick NiO film on
a Si wafer (Figure 1d). The diffraction peaks match well with
reported values for the cubic, crystalline NiO phase. A standard
from PDF Card 00-044-1159 is plotted in Figure 1d for reference. The additional feature observed in our experimental spectrum near 52 is a shoulder of a peak from the Si substrate.
The film appears smooth by SEM (Figure S1a, Supporting
Information).
To characterize the activity of the ALD NiO thin films toward
OER, we used CV. Figure 2a shows a typical CV of a NiO sample
in 0.1 M KOH after preconditioning. The NiO film thickness
was 8 nm thick. NiO tested in these conditions will be referred
to as the NiO/Fetrace. The OER onset voltage of the NiO/Fetrace
sample is 1.51 VRHE, and the overpotential needed to achieve
a current density of 10 mA cm2 is 0.54 V. The activity is consistent with other NiO substrates whose only source of Fe is
from trace amounts in a KOH electrolyte.[45,46]
All NiO samples display a redox feature at 1.45 VRHE.
Figure 2b expands this region for greater clarity of the redox
feature. The redox feature is believed to be related to the

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(3 of 10) 1500412

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com
Table 1. Effect of electrolyte on catalytic properties of NiO thin films.
TOF ( = 0.3 V)
[s1]
NiO/Fetrace

0.03 0.03

32

NiO/Fesatd

0.5 0.1

21 6

NiO/two-step

0.3 0.1

52

Ni0.75Fe0.25OOH46
Aged NiOx (in KOH with trace Fe)28
Ni0.9Fe0.1Ox

47

NiOOH (Fe-free)46

Figure 2. a) Representative CVs of NiO films after preconditioning in


0.1 M KOH (light blue) and 0.1 M KOH with Fe(NO3)3 (purple). The samples tested in Fe saturated solution showed greater activity than samples
tested in electrolyte containing only trace Fe. b) The region before the
OER onset is magnified to show the redox features present in both samples. The redox feature of the NiO/Fesatd sample is smaller and shifted to
more anodic potentials relative to that for NiO/Fetrace.

oxidation of Ni(OH)2 to NiOOH, and is reflective of the number


of electrochemically active Ni sites.[47,48] Although NiO is initially deposited, NiO is known to spontaneously form a surface of Ni(OH)2 upon exposure to alkaline conditions, and the
Ni(OH)2 can be oxidized upon applied potential to NiOOH at
potentials around 1.45 VRHE. The small size of this redox feature, corresponding to a small amount of charge exchange, is
much less than that reported for many other NiOx films,[24,45,46]
and is likely reflective of the smooth, dense film deposited by
ALD. The redox feature grows during the preconditioning process, consistent with an increasingly hydrated and porous film
due to prolonged exposure to the electrolyte (see Figure S2a in
the Supporting Information). Different film thicknesses do not
show a large difference in the size of these redox features over
the film thicknesses tested (from 1 to 18 nm). This suggests
that only the outer surface layers are active in these films over
the time period tested, reflecting the dense nature of the ALDdeposited NiO. This behavior is in contrast to electrodeposited
NiOOH, where almost all Ni atoms in the film are estimated to
be active.[28]
Given the low surface area of these films, it is important
to calculate the TOF for better comparison between catalysts.

1500412 (4 of 10)

wileyonlinelibrary.com

TOF ( = 0.4 V)
[s1]

0.11
0.2
0.21 0.03
0.010.09

The TOF is defined as the number of O2 molecules generated


per electrochemically accessible Ni site per second, using the
amount of charge passed in the redox feature in the cathodicgoing direction. We assume a one charge per Ni atom, which
may overestimate the actual number of Ni active sites,[40] but
provides a lower bound for the TOF. Although it is possible
that the Fe may be the active site, scaling to Ni is still useful
because it captures how efficiently the NiOx surface is used
and allows for easy monitoring of changes in surface roughness. Similar NiFe oxide based catalysts have reported near
100% faradaic efficiency, thus we assume that all current at the
selected overpotentials is due to oxygen evolution.[19] The TOFs
at 0.3 and 0.4 V overpotential after the preconditioning process
(NiO/Fetrace) are calculated to be 0.03 and 3 s1, respectively
(Table 1). As expected, the Ni/Fetrace TOF is much higher than
the TOF of NiOx samples where additional purification steps
removed Fe from the KOH electrolyte (0.02 s1 at = 0.4 V),
but lower than the TOF of NiFe oxide films reported in the
literature (>0.2 s1 at = 0.3 V).[28,46,47]
In the NiO/Fetrace samples, Fe is likely incorporated into the
ALD NiOx from the 0.1 M KOH electrolyte during the preconditioning process. During the preconditioning process, the
geometric activity increases (Figure S2a, Supporting Information), in contrast to studies of Ni(OH)2 performed in KOH
with additional Fe-purification steps, where the geometric
activity decreases with cycling.[46] Although the surface area
is increasing (see discussion below), the main contribution
to the increased activity is believed to be related to incorporation of trace Fe from the KOH electrolyte into the NiO thin
films.[46,49] However, other NiFe oxide films have shown much
greater activity than the NiO/Fetrace samples studied here, likely
related to higher Fe content created during fabrication of the
catalyst.[19,47,50,51] Louie and Bell showed that the activity of
electrodeposited NiOx films increases with increasing Fe concentration to about 10% Fe:(Fe + Ni), and maintains similarly
high activity between 15% and 50%.[28] Based on the low activity
of the NiO/Fetrace samples compared to more Fe-rich NiFe
oxides, it is likely that the NiO/Fetrace samples studied here contain below the optimum Fe concentration. With these results in
mind, we attempted to further enhance the activity of the NiO
catalyst through additional Fe.
To introduce additional Fe into the films, pristine NiO samples
were pretreated and tested in an Fe(NO3)3 saturated 0.1 M KOH
electrolyte (NiO/Fesatd samples). As seen in Figure 2a, NiO/Fesatd
samples show significantly enhanced catalytic activity compared

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1500412

www.advenergymat.de
www.MaterialsViews.com

Adv. Energy Mater. 2015, 5, 1500412

FULL PAPER

to NiO/Fetrace. The OER onset is at 1.48 VRHE, and a current


density of 10 mA cm2 is achieved at an overpotential of 0.48 V.
As shown in Table 1, the TOFs are 0.5 and 21 s1 at 0.3 and
0.4 V overpotential, respectively, which compare well with other
NiFe oxide OER catalysts.[27,28,47] The redox feature for the NiO/
Fesatd (Figure 2b) is shifted to more anodic potentials compared
to the redox feature for NiO/Fetrace, consistent with an increase
in the difficulty to oxidize Ni2+ with increasing Fe incorporation,
as observed previously.[46] Unlike the NiO/Fetrace, the activity of
the NiO/Fesatd stays relatively constant during preconditioning
(see Figure S2b in the Supporting Information), likely indicating
that Fe is quickly incorporated into the thin films.
Literature varies on the pretreatment process used to activate NiO thin films, ranging from drawing 10 mA cm2 current for a set amount of time[15,47] to soaking in an electrolyte
without any applied potential.[25,28] The standard pretreatment
we perform consists of two CVs, followed by a 30 min hold at
0.85 VRHE and then two CVs followed by a 10 min hold at
1.65 VRHE (see the Experimental Section). Although this process has not been fully optimized, our process allows for a systematic measurement of changes in the films while activating
the catalysts, and we find it provides an improvement over
soaking alone (see below). The CVs interspersed throughout
allowed us to monitor changes during the pretreatment. The
largest change in activity was generally found after the 30 min
hold. We probed the sensitivity of the activity to the pretreatment by varying the potential of the first 30 min hold. Although
there is a slight variance in the activity between samples before
the pretreatment, after this treatment, the activity tends to
converge within both the NiO/Fetrace and NiO/Fesatd sample
sets to CVs similar to Figure 2a. As seen in Figure 3a, which
plots TOF number versus pretreatment voltage, there is little
variation in the final TOF upon changing the applied voltage
for the samples. The low correlation between TOF and applied
potential suggests that Fe incorporation is not assisted by the
applied voltage, but is a diffusion-limited process. This is in
agreement with previous experiments in which the activity of
NiO was enhanced by simply soaking in a KOH electrolyte.[25,28]
The slight enhancement in activity for the NiO/Fetrace samples
at higher potentials may be related to increased turbulence due
to bubble formation. Regardless of the potential applied during
the pretreatment, the NiO/Fesatd consistently showed improved
activity compared to the NiO/Fetrace.
In addition, the differences between the NiO/Fetrace and NiO/
satd
Fe remain pronounced over time, according to chronoamperometry measurements. After pretreatment, the catalysts were
held at 1.6 VRHE for 2.5 h as shown in Figure 3b. The NiO/Fesatd
displays excellent stability over the first 2.5 h, even increasing in
activity slightly with time. The noisier current of the NiO/Fesatd
samples compared to the NiO/Fetrace samples is related to a
higher rate of bubble formation at the higher current density. The
NiO/Fetrace shows moderate stability over the first 2.5 h, but does
decrease in activity slightly. The decrease may be related to a transition to a less active Ni-Fe oxide,[30] or to a restructuring/dissolution that leads to less Fe on the surface.[45] The long-term stability
enhancement of the NiO/Fesatd sample is limited (Figure S6,
Supporting Information). After 5 h, the activity begins to decay
similar to previous NiOx thin film electrocatalysts,[31,45] possibly
also due to a transition to a less active NiFe oxide.[30]

Figure 3. a) TOF of NiO thin films versus the applied potential during a
30 min chronoamperometry preconditioning step for NiO/Fesatd (purple
squares) and NiO/Fetrace (blue triangles). The TOF did not depend heavily
on the applied potential, whereas addition of Fe to the electrolyte had a
strong effect. b) The NiO/Fesatd (purple) samples showed good stability and
activity compared to the NiO/Fetrace sample (blue) when held at 1.61 VRHE for
2.5 h. c) Both the NiO/Fesatd (purple) and NiO/Fetrace (blue) showed similar
low Tafel slopes of 30 mV decade1 at low current densities. The Tafel slope
increases at higher current densities, characteristic of NiOx OER catalysts.

To better understand the kinetics, Tafel plots were created as


seen in Figure 3c. Interestingly, both the NiO/Fetrace and NiO/
Fesatd show similar behaviors. At low potentials, both films have
low Tafel slopes of 30 mV decade1, characteristic of Fe-doped
NiOx thin films.[51] At higher overpotentials, the slope begins
to increase. Although this may be partly related to mass-transport limitations due to our experimental setup, it is worthwhile

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(5 of 10) 1500412

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com

noting that this is characteristic of Fe-doped NiOx films.[25] The


change in the slope has been attributed to a change in the film
that affects the rate-limiting step of the reaction, or even to
problems with charge transport.[52] The similar behavior of the
Tafel slopes suggests that there are similar active sites within
the two films, and that the main difference can be attributed to
the density of the active sites. Interestingly, a comparison of the
redox features of the NiO/Fetrace and NiO/Fesatd (Figure 2b) consistently shows that the number of electrochemically accessible
sites is greater in the NiO/Fetrace, the opposite of what would
be expected from the Tafel slopes alone. This suggests that the
number of active sites is closely tied to Fe concentration.
The different sizes of the NiO/Fetrace and NiO/Fesatd samples Ni(OH)2/NiOOH redox features in the CVs (e.g., see
Figure 2b) deserve greater exploration. To study this in greater
detail, we scanned 3 nm thick samples between 1.30 and
1.52 VRHE for 375 sweeps. Select sweeps for the NiO/Fetrace are
shown in Figure 4a. With each additional sweep, the redox features become larger. The charge passed in the cathodic-going
feature is calculated and plotted versus each cycle number in
Figure 4b for both the NiO/Fetrace and NiO/Fesatd samples. It is
readily apparent that the charge passed is larger for the NiO/
Fetrace samples than the NiO/Fesatd samples, and that the charge
passed for the NiO/Fetrace sample continues to grow while the
NiO/Fesatd feature appears to plateau.
The growth of the redox feature with cycling is not surprising
and is characteristic of the formation of hydrous oxides. In Ni
metal, for example, a porous hydrous layer on a compact anhydrous oxide layer is formed when in contact with an electrolyte
as described by the BurkeOSullivan duplex layer model.[20]
Potential cycling increases the hydrous layer at the expense of
the metal, increasing the number of electrochemically accessible metal sites, thereby increasing the size of the redox feature. While the extent of the oxidation is dependent on the pH,
range of applied potential, and scanning rate, a trend is generally observed in which the amount of charge rises then plateaus
with continued cycling.[53] This is reflective of the difficulty in
disturbing the growth of the dense oxide underlying the porous
hydroxide layer. It appears that NiO follows a similar mechanism, except that there is no Ni metal underlying the dense
oxide layer. The NiO/Fetrace redox feature grows quickly in the
first cycles, then the growth rate begins to slow. In NiO/Fesatd
samples, the presence of Fe(NO3)3 in the electrolyte appears to
inhibit the growth of hydroxide, perhaps through the stabilization of the underlying NiO phase or the formation of a denser
hydroxide that better decreases ion transfer.[20] The anodic shift
in the Ni2+/Ni3+ redox feature in the presence of an Fe rich electrolyte supports the conclusion that the Ni(OH)2/NiOOH phase
change becomes more difficult in the presence of Fe.
We used ex situ XPS to investigate the changes in the catalyst under OER conditions. We explored three different samples: a pristine ALD NiO, a NiO/Fetrace sample, and a NiO/Fesatd
sample. The testing procedure included preconditioning and ten
additional CVs. All samples were 18 nm thick. Figure 5a shows
survey XPS scans of a pristine 18 nm NiO film. The pristine
film contains both Ni and O at a ratio of 15:14 O:Ni, close to
the expected 1:1 ratio for NiO thin films. There is also some
adventitious C on the surface. Sputtering was not attempted
because oxygen is preferentially sputtered from NiOx thin films,
1500412 (6 of 10)

wileyonlinelibrary.com

Figure 4. a) Select scans from CVs of a NiO/Fetrace sample tested over


a span of 375 cycles. b) Amount of charge passed in the cathodic-going
direction of both the NiO/Fesatd and NiO/Fetrace over 375 cycles. The
amount of charge is correlated with the number of electrochemically
accessible Ni sites. The NiO/Fesatd consistently has less charge transferred and reaches a steady state value unlike the NiO/Fetrace over the
time-span tested, suggesting that Fe inhibits the restructuring of the NiO
thin films.

resulting in a change in the oxidation state. Survey scans taken


after the preconditioning treatment contain predominately O
and Ni at a ratio of 1.3 and 1.5 O:Ni for the NiO/Fesatd and
NiO/Fetrace, respectively, suggesting some oxidation of the
thin films.
High-resolution scans of the Ni 2p3/2 XPS peak were taken
to better understand changes in the oxidation state of the thin
films (Figure 5b). The spectrum for Ni is complicated because
of the significant overlap of the Ni 3d and O 2p orbitals,
resulting in broad peaks with multiple satellite peaks.[54]
For NiOx samples, peaks at 854 and 856 eV are typically
deconvoluted to indicate the contribution from the Ni2+ and
Ni3+ to the Ni 2p3/2 feature, respectively.[45] In Figure 5b, the
lower and higher binding energy peaks are shown in red and
yellow, respectively. Although the exact contribution is likely
more complicated than this simple deconvolution,[54] particularly for NiO, we choose to follow this convention as a means
of tracking changes in oxidation. NiO and Ni(OH)2 contain
Ni2+, while NiOOH has Ni3+ with some Ni4+ contributions, so
a growth in the yellow peak is indicative of greater concentration of NiOOH. The pristine ALD film shows two strong peaks
at 853.7 eV (red) and 855.7 eV (yellow) along with broad satellite peaks at higher energies (green and blue). The ratio of the
lower and higher energy peak areas of the NiO pristine sample
are 1.5:1 Ni2+:Ni3+, a ratio which is characteristic of other

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1500412

www.advenergymat.de
www.MaterialsViews.com

FULL PAPER
Figure 5. a) XPS survey scan of pristine ALD NiO confirming that NiO was deposited on the surface. b) High-resolution scans of the Ni 2p3/2 peak
before and after testing in either 0.1 M KOH or 0.1 M KOH with added Fe(NO3)3. The feature highlighted in red is often associated with Ni2+ and
feature highlighted in yellow is often associated with Ni3+. The green and blue peaks are satellite features. c) High-resolution scans of the O 1s peak
for pristine and treated samples. The red, green, and blue peaks are assigned to O bonded as NiONi, NiOH, and adsorbed H2O, respectively.
The NiO/Fetrace sample became more hydrated than the NiO/Fesatd based on the greater proportion of OH and H2O in the films. d) High-resolution
scans of the Fe 3p peak show that the NiO/Fesatd sample has more Fe at its surface than the other samples tested. The increase in counts at higher
binding energies is due to the Ni 3p feature.

NiO XPS spectra in the literature.[45,47,54,55] After testing in the


Fe saturated solution, there is little change in the spectra, with
only a slight increase in the Ni3+ peak relative to the Ni2+ peak
(1.6:1). However, after identical testing in 0.1 M KOH, the size
of the Ni3+ peak increases more relative to the Ni2+ peak (2.1:1).
Upon continued testing in 0.1 M KOH electrolyte for 2.5 h, the
size of Ni2+ feature continues to decrease and the Ni3+ feature
continues to increase (Figure S3, Supporting Information),
becoming more like the spectrum for NiOOH in literature.
Since the higher energy feature is indicative of the NiOOH
contribution, the XPS results support a time-dependent formation of a NiOOH from Ni2+ upon exposure to 0.1 M KOH electrolytes. Interestingly, the presence of additional Fe appears to
reduce the extent of additional oxidation of the thin films, consistent with the CV measurements.

Adv. Energy Mater. 2015, 5, 1500412

High-resolution scans of the O 1s peak can also provide evidence for restructuring, specifically the level of hydration within
a thin film. The binding energy of the O varies depending on
whether it is bonded as NiONi, NiOH, or adsorbed H2O.[55]
Peaks at 529 (red), 531 (green), and 532 eV (blue) are assigned
to O bonded as NiONi, NiOH, and adsorbed H2O, respectively.[56,57] The spectrum for the pristine ALD film contains
oxygen predominately bonded as NiONi (66%), followed by
NiOH (31%), with a small amount of adsorbed H2O (3%), as
seen in Figure 5c. Testing in Fe saturated KOH results in a slightly
more hydrated film (62%, 37%, and 1%, respectively). Similar to
the Ni 2p3/2 XP spectra, testing NiO in 0.1 M KOH with only trace
Fe once again results in even larger changes in the ALD film, with
the film containing 48% NONi, 48% NiOH, and 5% H2O. These
changes further support the hypothesis that a hydrated film is

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(7 of 10) 1500412

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com

Fe may actually be the active site. When Fe occupies Ni sites in


formed under testing conditions, and that the presence of Fe
a NiOOH matrix at concentrations <25%, changes in both the
during testing reduces the amount of restructuring in the film.
oxidation and metal-to-oxygen bond distance make the binding
Despite the change in oxidation of the Ni, after testing there is no
of Fe to intermediates near ideal for the OER.[29] Both explanadetectable change in morphology by SEM (Figure S1, Supporting
Information). While this result contrasts with our previous
tions of the high activity of NiFe oxides are consistent with our
work on oxidation of ALD MnO under OER conditions,[58] the
results and explain how despite the lower surface area, the Fedoped catalyst shows superior activity. Preliminary screening
preservation of the continuity of the NiO film upon oxidation/
on the effects of Mn, Co, Cu, Ce, and Rh through the addition
increasing amorphization is similar to other highly active OER
of their nitrate salts did not show an enhancement in TOF, nor
catalysts where the crystalline surface becomes amorphous with
did the addition of KNO3, further pointing toward the imporOER testing.[59] The oxidation is believed to follow a process similar to the BurkeOSullivan duplex layer model.[20]
tance of Fe to the high TOF, and not another constituent in the
KOH nor a by-product of the Fe(NO3)3 salt. For the most active
It is difficult to directly measure the small amounts of Fe in
the film by XPS using an Al(K) source because the strongest
catalyst on a geometric basis, it would be beneficial to have both
high TOF and high surface area.
Fe signal (Fe 2p3/2) overlaps with a Ni Auger signal. We instead
To this end, we explored a two-step process to synthesize
look at the Fe 3p feature at 56 eV, shown in Figure 5d. The
highly active NiO OER electrocatalysts that combined both high
Fe 3p signal is below the detection limit for the pristine ALD
TOF and higher surface area. The procedure is summarized in
sample and the NiO/Fetrace sample. However, for the NiO/Fesatd
Figure 6a. A 3 nm ALD NiO catalyst was first preconditioned
sample, a clear Fe feature is observed. This result supports
following the standard procedure in 0.1 M KOH with only trace
our conclusions from the electrochemistry measurements that
more Fe is incorporated in the film when using the Fe satuFe. This produced a catalyst with high surface area. A CV after
rated electrolyte compared to the trace Fe electrolyte. By exampreconditioning of this sample is shown in Figure 6b (blue).
ining the ratio of Ni 3p and Fe 3p peaks, we can estimate that
The electrolyte was then exchanged for an Fe(NO3)3 saturated
between 10% and 15% Fe is present in the NiO/Fesatd films.
electrolyte, the sample was soaked for 10 min, and then another
CV was taken (orange). As seen in Figure 6b, the activity of the
Additional exposures to Fe saturated electrolyte do not signifiNiO/two-step sample was greatly enhanced over the activity of
cantly enhance this amount, suggesting an upper limit on Fe
the NiO/Fetrace sample. The TOF is 0.3 s1 at = 0.3 V, comcontent in NiOx catalysts achievable under these reaction conditions. Interestingly, this range is close to the transition at which
pared to 0.05 s1 before the Fe(NO3)3 addition, likely reflecting
regions of FeOOH begin to precipitate in a NiFe oxide solid
the beneficial role of Fe to solution. In addition, it has a higher
solution.[29]
geometric activity than a sample that was simply preconditioned
The XPS analysis in combination with the
behavior of the redox feature for both the
NiO/Fetrace and NiO/Fesatd samples point to
a greater restructuring in the NiO film when
pretreated with an electrolyte containing trace
Fe compared to one that is saturated with
Fe(NO3)3. Similar reductions in the degree
of restructuring of Ni have been observed in
CoNi mixtures, where it was hypothesized
that the NiCo oxide spinel was unable to
transform into a layered hydroxide/oxyhydroxide material.[47] The higher surface area
is beneficial for obtaining highly active catalysts on a geometric basis.[32] This would be
expected to improve the activity of the NiO/
Fetrace sample since it has the largest increase
in surface area.
However, the effect of surface area is
dwarfed by the much higher TOF of the NiO/
Fesatd sample compared to the NiO/Fetrace
sample due to the higher intrinsic activity of
NiFe oxides. It has long been believed that
the presence of Fe may affect the binding of
OER intermediates to Ni active sites. It has
been shown that the presence of Fe affects
the redox properties of Ni making it harder
to oxidize, similar to our results shown in
Figure 6. a) Schematic for a two-step procedure to fabricate high surface area to make NiO/
Figure 2b.[28] This in turn reduces the average two-step samples. b) CVs to compare OER performance of NiO/Fetrace, NiO/Fesatd, and NiO/
oxidation state and produces a Ni site that is two-step. The NiO/two-step samples have the best OER activity of the samples tested on a
more active for OER.[28,30,60] Alternatively, the geometric scale. c) CVs of samples soaked in the electrolyte for 24 h.

1500412 (8 of 10)

wileyonlinelibrary.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1500412

www.advenergymat.de
www.MaterialsViews.com

Adv. Energy Mater. 2015, 5, 1500412

will enable the creation of nanoengineered devices with conformal catalytic coatings. Its fine control over film thickness
and uniformity will allow for the minimization of parasitic light
absorption and charge transport resistances, and improved
stability of unstable light absorbers.[1,52,62] As illustrated in
this work, care must be taken to condition the ALD catalyst
to the desired surface area and active state. Many variables
can be used to optimize this process, helping make ALD of
electrocatalysts a versatile method suitable for many material
systems.

4. Conclusions
Highly active, crystalline NiO catalysts were synthesized by
ALD. The activity of these catalysts can be tailored by the concentration of Fe in the electrolyte, with higher Fe concentrations allowing for higher TOF. Interestingly, it appears that
the addition of Fe also leads to a reduction in the electrochemically active surface area compared to cycling in electrolytes
with lower concentrations of Fe. Preconditioning that involves
potential cycling and chronoamperometry increases catalyst
surface area compared to a procedure involving only soaking
in an electrolyte, resulting in a higher geometric activity. Preconditioning in two different electrolytes can be used to create
catalysts with both high TOF and higher surface area. These
results are especially important when the starting catalyst is a
compact oxide, such as those often deposited by ALD, and may
provide important guidance to other electrocatalytic material
systems.

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

Acknowledgements
This work was supported as part of the Center on Nanostructuring
for Efficient Energy Conversion (CNEEC), an Energy Frontier Research
Center funded by the U.S. Department of Energy, Office of Science, Basic
Energy Sciences, under Award # DE-SC0001060. The authors would like
to thank Vijay Parameshwaran for helpful discussions.
Received: February 26, 2015
Revised: May 10, 2015
Published online: June 18, 2015

[1] Q. Peng, J. S. Lewis, P. G. Hoertz, J. T. Glass, G. N. Parsons, J. Vac.


Sci. Technol. A 2012, 30, 010803.
[2] J. R. Bakke, K. L. Pickrahn, T. P. Brennan, S. F. Bent, Nanoscale
2011, 3, 3482.
[3] C. Marichy, M. Bechelany, N. Pinna, Adv. Mater. 2012, 24, 1017.
[4] S. M. George, Chem. Rev. 2010, 110, 111.
[5] H. Kim, H.-B.-R. Lee, W. J. Maeng, Thin Solid Films 2009, 517, 2563.
[6] R. W. Johnson, A. Hultqvist, S. F. Bent, Mater. Today 2014, 17, 236.
[7] M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. X. Mi,
E. A. Santori, N. S. Lewis, Chem. Rev. 2010, 110, 6446.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

(9 of 10) 1500412

FULL PAPER

in Fe(NO3)3 alone (purple line), despite having half the TOF


(see Table 1). This greater geometric activity is related to the
increased surface area from the 0.1 M KOH conditioning; however the TOF is lower likely due to a lower ratio of Fe to Ni
electrochemically accessible sites in the film. Performing the
reverse experiment, in which a NiO/Fesatd sample was prepared
and then tested in 0.1 M KOH, produced catalysts with similar
activity to NiO/Fesatd (see Figure S4 in the Supporting Information). This demonstrates that the concentration of Fe in
the electrolyte is itself not important for catalytic activity once
the Fe has become incorporated into the catalyst and that the
increase in activity is indeed a change in the catalyst.
Interestingly, we find that an applied potential is beneficial
to obtaining the high activity from the Fe(NO3)3 electrolyte
treatment when using ALD NiO. For comparison, NiO films
were soaked for 24 h in either 0.1 M KOH or 0.1 M KOH with
Fe(NO3)3 electrolyte before performing a CV in 0.1 M KOH. The
second CV of the NiO catalysts soaked in the different electrolytes is shown in Figure 6c, with the NiO/Fetrace CV plotted
for comparison. The presoaked samples activities are similar
to the NiO/Fetrace samples, even when presoaked in an Fe-rich
electrolyte. Subsequent cycling increases the activity of both
films due to increased hydration, but the geometric activity
remains below that of the NiO/Fesatd. We suggest that this is
related to the lack of hydration in the soaking process that is
needed to create a high surface area catalyst. An applied potential with cycling assists in disturbing the dense NiO thermal
oxide deposited by ALD and transforming it into a permeable, hydrous oxide as seen in Figure 4. As discussed above, it
appears that the Fe saturates in the NiOx matrix, and without
the additional restructuring, less Fe is incorporated into the
film. Thus, a pretreatment with applied potentials is beneficial
in achieving high activity of compact ALD NiO films. The range
of applied potentials can also determine the degree of restructuring, and a smaller range may result in less restructuring and
lower activity (see the Supporting Information).
Further tailoring of preconditioning parameters such as
pH and applied potentials may be able to further enhance the
activity of these thin films. Such pretreatments offer a facile
way to tailor both the chemistry and geometry of a catalyst.
Parameters such as concentration of Fe in the electrolyte and
duration of exposure to an Fe-containing electrolyte can be
used to alter the concentration of Fe on a surface (Figure S2,
Supporting Information). However, since the Fe saturates on
the NiO surface, and Fe effects how the NiO surface restructures, it is important to control how these materials are
tested. These pretreatments are particularly important when
working with ALD-grown films, which often have low surface
area due to the deposition of dense thin films, as seen here
by the comparatively small redox features for NiO on a per
area basis.
The use of ALD in the creation of electrocatalysts and photoelectrocatalysts for water splitting is a rich field with many
opportunities. Nanostructuring water splitting devices allows
for higher surface area, the decoupling of length scales for light
absorption and charge transport, and the modification of band
gaps, among other possibilities, and has already shown much
promise.[16,61] As the importance of nanostructuring water
splitting devices continues to grow, ALD of electrocatalysts

www.advenergymat.de

FULL PAPER

www.MaterialsViews.com
[8] Y. J. Hwang, A. Boukai, P. Yang, Nano Lett. 2008, 9, 410.
[9] A. Paracchino, N. Mathews, T. Hisatomi, M. Stefik, S. D. Tilley,
M. Gratzel, Energy Environ. Sci. 2012, 5, 8673.
[10] S. Hu, M. R. Shaner, J. A. Beardslee, M. Lichterman, B. S. Brunschwig,
N. S. Lewis, Science 2014, 344, 1005.
[11] Y. W. Chen, J. D. Prange, S. Duhnen, Y. Park, M. Gunji,
C. E. D. Chidsey, P. C. McIntyre, Nat. Mater. 2011, 10, 539.
[12] J. Yang, K. Walczak, E. Anzenberg, F. M. Toma, G. Yuan, J. Beeman,
A. Schwartzberg, Y. Lin, M. Hettick, A. Javey, J. W. Ager, J. Yano,
H. Frei, I. D. Sharp, J. Am. Chem. Soc. 2014, 136, 6191.
[13] N. P. Dasgupta, C. Liu, S. Andrews, F. B. Prinz, P. Yang, J. Am.
Chem. Soc. 2013, 135, 12932.
[14] J. Resasco, N. P. Dasgupta, J. R. Rosell, J. Guo, P. Yang, J. Am.
Chem. Soc. 2014, 136, 10521.
[15] V. O. Williams, E. J. DeMarco, M. J. Katz, J. A. Libera, S. C. Riha,
D. W. Kim, J. R. Avila, A. B. F. Martinson, J. W. Elam, M. J. Pellin,
O. K. Farha, J. T. Hupp, ACS Appl. Mater. Interfaces 2014, 6, 12290.
[16] T. M. Gr, S. F. Bent, F. B. Prinz, J. Phys. Chem. C 2014, 118, 21301.
[17] Y. Jiao, Y. Zheng, M. Jaroniec, S. Z. Qiao, Chem. Soc. Rev. 2015, 44,
2060.
[18] E. Fabbri, A. Habereder, K. Waltar, R. Kotz, T. J. Schmidt, Catal. Sci.
Technol. 2014, 4, 3800.
[19] C. C. L. McCrory, S. Jung, J. C. Peters, T. F. Jaramillo, J. Am. Chem.
Soc. 2013, 135, 16977.
[20] R. L. Doyle, I. J. Godwin, M. P. Brandon, M. E. G. Lyons, Phys.
Chem. Chem. Phys. 2013, 15, 13737.
[21] J. S. Ponraj, G. Attolini, M. Bosi, Crit. Rev. Solid State 2013, 38, 203.
[22] K. L. Pickrahn, S. W. Park, Y. Gorlin, H. B. R. Lee, T. F. Jaramillo,
S. F. Bent, Adv. Energy Mater. 2012, 2, 1269.
[23] J. Luo, J.-H. Im, M. T. Mayer, M. Schreier, M. K. Nazeeruddin,
N.-G. Park, S. D. Tilley, H. J. Fan, M. Grtzel, Science 2014, 345,
1593.
[24] B. S. Yeo, A. T. Bell, J. Phys. Chem. C 2012, 116, 8394.
[25] M. R. G. de Chialvo, A. C. Chialvo, Electrochim. Acta 1988, 33,
825.
[26] A. Singh, S. L. Y. Chang, R. K. Hocking, U. Bach, L. Spiccia, Energy
Environ. Sci. 2013, 6, 579.
[27] L. Trotochaud, S. W. Boettcher, Scr. Mater. 2014, 74, 25.
[28] M. W. Louie, A. T. Bell, J. Am. Chem. Soc. 2013, 135, 12329.
[29] D. Friebel, M. W. Louie, M. Bajdich, K. E. Sanwald, Y. Cai,
A. M. Wise, M.-J. Cheng, D. Sokaras, T.-C. Weng, R. Alonso-Mori,
R. C. Davis, J. R. Bargar, J. K. Nrskov, A. Nilsson, A. T. Bell, J. Am.
Chem. Soc. 2015, 137, 1305.
[30] Y.-F. Li, A. Selloni, ACS Catal. 2014, 4, 1148.
[31] M. J. Kenney, M. Gong, Y. Li, J. Z. Wu, J. Feng, M. Lanza, H. Dai,
Science 2013, 342, 836.
[32] F. Song, X. Hu, Nat. Commun. 2014, 5, 6191.
[33] M. Utriainen, M. Krger-Laukkanen, L. Niinist, Mater. Sci. Eng. B
1998, 54, 98.
[34] E. Lindahl, M. Ottosson, J.-O. Carlsson, Chem. Vap. Deposition
2009, 15, 186.

1500412 (10 of 10)

wileyonlinelibrary.com

[35] P. A. Premkumar, A. Delabie, L. N. J. Rodriguez, A. Moussa,


C. Adelmann, J. Vac. Sci. Technol. A 2013, 31, 061501.
[36] E. Thimsen, A. B. F. Martinson, J. W. Elam, M. J. Pellin, J. Phys.
Chem. C 2012, 116, 16830.
[37] H. L. Lu, G. Scarel, C. Wiemer, M. Perego, S. Spiga, M. Fanciulli,
G. Pavia, J. Electrochem. Soc. 2008, 155, H807.
[38] S. J. Song, S. W. Lee, G. H. Kim, J. Y. Seok, K. J. Yoon, J. H. Yoon,
C. S. Hwang, J. Gatineau, C. Ko, Chem. Mater. 2012, 24, 4675.
[39] Y. T. Chong, E. M. Y. Yau, K. Nielsch, J. Bachmann, Chem. Mater.
2010, 22, 6506.
[40] D. A. Corrigan, S. L. Knight, J. Electrochem. Soc. 1989, 136, 613.
[41] J. Bachmann, A. Zolotaryov, O. Albrecht, S. Goetze, A. Berger,
D. Hesse, D. Novikov, K. Nielsch, Chem. Vap. Deposition 2011, 17, 177.
[42] X. Tong, Y. Qin, X. Guo, O. Moutanabbir, X. Ao, E. Pippel, L. Zhang,
M. Knez, Small 2012, 8, 3390.
[43] H.-B.-R. Lee, K. L. Pickrahn, S. F. Bent, J. Phys. Chem. C 2014, 118,
12325.
[44] R. Methaapanon, S. M. Geyer, H.-B.-R. Lee, S. F. Bent, J. Mater.
Chem. 2012, 22, 25154.
[45] B. Mei, A. A. Permyakova, R. Frydendal, D. Bae, T. Pedersen,
P. Malacrida, O. Hansen, I. E. L. Stephens, P. C. K. Vesborg,
B. Seger, I. Chorkendorff, J. Phys. Chem. Lett. 2014, 5, 3456.
[46] L. Trotochaud, S. L. Young, J. K. Ranney, S. W. Boettcher, J. Am.
Chem. Soc. 2014, 136, 6744.
[47] L. Trotochaud, J. K. Ranney, K. N. Williams, S. W. Boettcher, J. Am.
Chem. Soc. 2012, 134, 17253.
[48] P. Oliva, J. Leonardi, J. F. Laurent, C. Delmas, J. J. Braconnier,
M. Figlarz, F. Fievet, A. d. Guibert, J. Power Sources 1982, 8, 229.
[49] D. A. Corrigan, J. Electrochem. Soc. 1987, 134, 377.
[50] E. L. Miller, R. E. Rocheleau, J. Electrochem. Soc. 1997, 144, 3072.
[51] Y. Qiu, L. Xin, W. Li, Langmuir 2014, 30, 7893.
[52] V. Viswanathan, K. Pickrahn, A. C. Luntz, S. F. Bent, J. K. Norskov,
Nano Lett. 2014, 14, 5853.
[53] S. Rebouillat, M. E. G. Lyons, M. P. Brandon, R. L. Doyle, Int. J. Electrochem. Sci. 2011, 6, 5830.
[54] A. G. Marrani, V. Novelli, S. Sheehan, D. P. Dowling, D. Dini, ACS
Appl. Mater. Interfaces 2013, 6, 143.
[55] M. A. Peck, M. A. Langell, Chem. Mater. 2012, 24, 4483.
[56] J. Li, F. Luo, Q. Zhao, Z. Li, H. Yuan, D. Xiao, J. Mater. Chem. A
2014, 2, 4690.
[57] D. Tian, N. Li, F. F. Wang, S. L. Mu, D. Y. Li, G. F. Xia, ECS Electrochem. Lett. 2012, 1, H5.
[58] K. L. Pickrahn, Y. Gorlin, L. C. Seitz, A. Garg, D. Nordlund,
T. F. Jaramillo, S. F. Bent, Phys. Chem. Chem. Phys. 2015, 17, 14003.
[59] K. J. May, C. E. Carlton, K. A. Stoerzinger, M. Risch, J. Suntivich,
Y.-L. Lee, A. Grimaud, Y. Shao-Horn, J. Phys. Chem. Lett. 2012, 3,
3264.
[60] M. Gong, H. Dai, Nano Res. 2015, 8, 23.
[61] S. Chen, J. Duan, M. Jaroniec, S. Z. Qiao, Angew. Chem. Int. Ed.
2013, 52, 13567.
[62] T. Wang, Z. Luo, C. Li, J. Gong, Chem. Soc. Rev. 2014, 43, 7469.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Energy Mater. 2015, 5, 1500412

Вам также может понравиться