Вы находитесь на странице: 1из 8

Journal of The Electrochemical Society, 154 3 D126-D133 2007

D126

0013-4651/2007/1543/D126/8/$20.00 The Electrochemical Society

Electrochemical Growth of Diverse Iron Oxide


(Fe3O4, -FeOOH, and -FeOOH) Thin Films by
Electrodeposition Potential Tuning
L. Martinez,a D. Leinen,a F. Martn,a M. Gabas,a J. R. Ramos-Barrado,a,*
E. Quagliata,b and E. A. Dalchieleb,*,z
a

Laboratorio de Materiales y Superficie (Unidad Asociada al CSIC), Departamentos de Fsica Aplicada &
Ingeniera Qumica, Universidad de Mlaga, E29071 Mlaga, Spain
b
Facultad de Ingeniera, Instituto de Fsica, 11000 Montevideo, Uruguay
Magnetite, goethite, and lepidocrocite thin films have been electrochemically grown on titanium substrates by the anodic oxidation
of ferrous ions in a 0.01 M FeSO4NH42SO46H2O + 0.04 M CH3COOK, pH 6.0, aqueous solution. It is demonstrated that the
deposition potential can be used as a tool to tune the obtainment of the different pure phases of the iron oxide-oxyhydroxides thin
films. Results of an exhaustive structural characterization, a morphological study, and X-ray photoelectron spectroscopy characterization are presented.
2007 The Electrochemical Society. DOI: 10.1149/1.2424416 All rights reserved.
Manuscript submitted March 31, 2006; revised manuscript received October 23, 2006. Available electronically January 17, 2007.

Iron oxides-oxyhydroxides are abundant materials in nature, and


are relevant in many scientific and technical applications, participating in a number of environmentally and industrially important processes. For the sake of simplicity, all of these iron oxidesoxyhydroxides are referred to here to as iron oxides. Iron oxides
are the subject of interest within the area of many applied sciences
such as geochemistry, mineralogy, corrosion science, soil science,
medicine, biology, metallurgy, and environmental protection.1 Iron
oxides can have many different crystalline forms. There are 15 iron
oxides, oxyhydroxides, and hydroxides known to date.1 They show a
variety of physical and chemical properties depending on their
chemical composition and crystal structure. Among the more important ones, we have magnetite Fe3O4, goethite -FeOOH, and
lepidocrocite -FeOOH. Magnetite was the first known magnetic
material and it is still now largely studied as it exhibits very interesting properties. Magnetite is a ferromagnetic mixed-valence 3d
transition metal oxide that has an inverse spinel structure.2 Magnetite undergoes a metal to insulator Verwey transition at ca. 120 K,
and the Curie temperature of magnetite is 860 K.3,4 In recent years,
magnetite has received much attention due to the presence of a
higher spin-polarization than in normal metals, suggesting applications in giant magnetoresistance GMR sensors and memories.3
Goethite is the most abundant and thermodynamically stable iron
oxyhydroxide mineral present in soils and sediments in cool
climates.5 Well-crystallized goethite is antiferromagnetic below its
Nel temperature of about 400 K, and its magnetic structure is similar to the crystalline one.1,6 It has been extensively studied from
many different perspectives with regard to structure and surface
reactivity.1,6 For instance, the formation of thin films of goethite is
of special importance to protect the steel from further corrosion.5
Lepidocrocite is generally less widespread than its polymorph goethite, but it can be encountered as orange accumulations in certain
environments.1 Lepidocrocite is sometimes used as starting material
to produce maghemite, which is used as magnetic pigment. Moreover, there have been encouraging results reported on the utilization
of goethite and lepidocrocite as electrodes for rechargeable lithiumion batteries, showing very high lithium intercalation capacity, almost complete reversibility, excellent rate capability, and nearly perfect capacity retention upon cycling.7,8
In addition to the properties of iron oxides as a bulk material that
were mentioned above, the thin-film forms of iron oxides show very
remarkable properties from a scientific and technological point of
view.9,10
Iron oxide thin films have been obtained by different techniques:

* Electrochemical Society Active Member.


z

E-mail: dalchiel@fing.edu.uy

chemical vapor deposition CVD,10 sputtering,3 plasma-enhanced


chemical vapor deposition,11 reactive sputter deposition,12 dc reactive magnetron sputtering,2 metallorganic chemical vapor
deposition,13 etc. There are very few reports dealing with the preparation of goethite and lepidocrocite thin films; the majority is devoted to the magnetite thin-film synthesis. However, until now it has
been very difficult to grow iron oxide thin films possessing welldefined surface composition.12 In general, the formation of pure iron
oxide phases using physical methods is not a simple, easy, or direct
issue.14 Moreover, in the particular case of magnetite thin films, very
strong research is aimed at a low-temperature preparation technique
573 K, which is one of the most important factors for
application.3 In fact, in the production of hybrid ferromagnetic oxide ferromagnets-semiconductor structures, processing temperatures must be kept far below 1000C to prevent reaction with the
semiconductor.15
Low-temperature, in situ fabrication of crystalline thin films is
essential in order to improve film quality, lower production costs,
and make the whole process environmentally friendly. In this sense,
soft-solution processing SSP techniques of thin-film preparation
appears to be an interesting alternative route to achieve these
goals.16 They can be defined as environmentally friendly processing,
using aqueous solutions that seem to provide results similar to other
processes using fluids such as vapor, gas, and plasma or beamvacuum processing, while consuming less total energy than other
processing routes.16-18 Electrodeposition is a typical soft-solution
processing application which enables in the present case the formation of iron oxide thin films from an aqueous solution between
room temperature and 100C under atmospheric pressure. In addition to its capability of controlling composition and morphology by
electrochemical parameters, the advantage of working in the lowtemperature domain becomes very attractive as soon as interdiffusion phenomena found at high temperatures becomes critical, as for
example as was mentioned above for hybrid ferromagneticsemiconductor structure fabrication.
However, in electrochemical synthesis of iron oxides, an accurate
control of growth potential and pH is necessary because several of
them present similar stability regions in the potential-pH diagrams.19
Growing these types of iron oxides onto foreign and inert substrates is of interest. It allows electrochemical studies of iron oxides
without the influence of iron and its possible electrochemical
response.9 From a technical point of view, there is for example much
interest in microelectronics in the production of hybrid
ferromagnetic-semiconductor structures, combining oxide ferromagnet thin-film devices with traditional semiconductors.15,20
The anodic film deposition of iron oxides from ferrous ion in
aqueous solutions was studied in the past in relation to the corrosion
of iron. Synthesis generally performed at room temperature onto

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Journal of The Electrochemical Society, 154 3 D126-D133 2007

D127

Table I. Literature survey of iron oxide-oxyhydroxide thin films obtained by electrodeposition onto inert substrates. Here are reviewed only
those works in which the iron oxide thin films were obtained by one-step electrodeposition process.
Iron oxide/
oxyhydroxide
thin film
Fe3O4
Fe3O4

Fe3O4

-FeOOH

-FeOOH

Fe3O4

Fe3O4

a
b

Substrate

Electrodeposition bath and growth conditions

ITO

0.0075 M FeCl2, pH 79, T = 2480C


Pot.a
Single
0.004 M KCH3COO + 0.01 M
crystals:
FeNH42SO426H2O
Au110,
pH 6.5, T = 90C
Au100 and Pot. and Galv.b
Au111
Au
0.01 M FeCl24H2O + 0.4 M NaCl + 2 102 M 1methyl-imidazole
pH 9.6, T = 70C
Pot.
Au
0.01 M FeCl24H2O + 0.4 M NaCl + 0.1 M 1-methylimidazole
pH 8.0, T = 25C
Pot.
Au
0.01 M FeCl24H2O + 0.4 M NaCl + 0.02 M 1-methylimidazole
pH 7.5, T = 1055C
Pot.
Au
2 gr KCH3COO + 2 gr FeNH42SO426H2O in 500
ml H2O
pH 6.3, T = 80C
Pot.
Au
50 ml of 1.0 M FeCl2 and 200 ml of 1.6 M
CH3COONH4 + NaOH
pH 9.5, T = ?
Pot.

Observations

Ref.

Polycrystalline films with a cubic spinel structure having no 28


preferred orientation.
20
On Au110 a 110 Fe3O4
orientation has been observed. On Au100, the films
presented 100, 111, and 511 orientations. On Au111,
both 111 and 511 orientations are observed.
Structural results not shown and not discussed.

Structural results not shown and not discussed.

Broad XRD peaks, at 55C growth temperature


a 002 preferential orientation can be observed.

29

Polycrystalline Fe3O4 samples without a strong preferential


orientation have been obtained.

30

Structural results not shown.


Authors comment: XRD showed the presence of a single
phase magnetite.

31

Pot.-potentiostatic electrodeposition.
Galv.-galvanostatic electrodeposition.

inert Au or Pt substrates, from different salt solutions of FeII, and


at pH within the 5.88.4 range, can be encountered in the
literature.21-26 These studies showed that the experimental conditions of preparation i.e., pH, anionic composition, oxygen concentration of the solution influenced the film composition. Furthermore, in these works, amorphous ferric oxides and a low crystalline
lepidocrocite phases have been reported. In some of the films, the
lepidocrocite phase has been identified by electron diffraction,21,22
Auger electron spectroscopy AES24,25 or Raman spectroscopy.26
The anodic film deposition of iron oxides from ferrous ion in
aqueous solution was also studied in the past by Abe et al.,27 as a
way to prepare magnetite films within their studies on ferrite preparation. Polycrystalline magnetite layers with a cubic spinel structure,
having no preferred orientation, were grown from FeCl2 solutions
onto copper substrates at pH 8 and at a temperature of 65C.
In addition to those previously reviewed works, in recent years a
new and renewed interest in the synthesis by electrodeposition, of
reproducible and well-characterized pure iron oxide thin films became evident Table I lists the more recent studies where iron oxide
thin films were grown by electrodeposition in one stage in one step
without annealing, and onto inert substrates. Only a few groups
reported the preparation of iron oxides by electrodeposition onto
inert substrates, though not all of them reported their structural properties. So, the growth by electrodeposition of high-quality iron oxides is a challenging task.
In the present work several iron oxide thin films were grown by
electrodeposition onto titanium substrates, and from the same electrolytic bath. It is demonstrated that the deposition potential can be
used as a tool to tune the different phases magnetite, goethite, and
lepidocrocite of the iron oxide films. Results of an exhaustive structural characterization, a morphological study, and X-ray photoelectron spectroscopy XPS characterization are presented.

Experimental
The diverse iron oxides were anodically deposited in a conventional three-electrode, single-compartment electrochemical cell
under potentiostatic conditions, from mildly stirred solutions.
Electrolytic bath was an aqueous solution of 0.01 M
FeSO4NH42SO46H2O and 0.04 M CH3COOK, pH 6.0 adjusted
with NaOH, was kept at a constant temperature of 90C. As the
bath temperature was near the boiling point of the solution, a condenser was used in order to avoid very high evaporation and then
concentration of the electrolytic solution. All solutions were prepared from analytical grade reagents and 18 M cm Millipore water. The electrolyte solutions were purged with nitrogen for 30 min
prior to each experimental series and then kept under flowing nitrogen during the experiments. The substrates were titanium plates
about 1 cm2 of geometric area, which before each experiment
were polished with emery paper, rinsed with deionized water, then
immersed for 10 s in 5% HF solution, and finally rinsed with water.
In both electrochemical growth and electrochemical measurements,
a platinum sheet was used as a counter electrode and a silver-silver
chloride saturated electrode Ag/AgCl, KCl satd; E = 0.197 V vs
NHE served as reference electrode. All potential values in this paper are reported vs this reference electrode.
The electrochemical measurements and electrodepositions were
carried out using a model 1286 Solartron electrochemical potentiostat. The total electrical charge passed was monitored to estimate
the film thickness assuming a faradaic efficiency of 100% using
the equation d = MQ/nAF; where d is the film thickness; M molecular weight; Q the total charge passed during deposition; A film
area; the density of the film; F the Faraday constant; and n the

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

D128

Journal of The Electrochemical Society, 154 3 D126-D133 2007


posited iron oxide thin films in the potential E range 400 E
1000 mV see the next section evidences formation of the following phases as the potential is made more positive and in this
sequential order: magnetite, goethite, and lepidocrocite. Then, in the
following, Fig. 1 is discussed on the basis of knowing that those are
the possible phases which can be obtained, and by using the
potential-pH diagrams for the iron-water system.
According to Pourbaix potential-pH diagrams of iron-water
system,32 the growth of Fe3O4 films would proceed via the oxidation
of Fe2+ by the reaction
3Fe2+ + 4H2O Fe3O4 + 8H+ + 2e

Based on the Nernst equation, the thermodynamic equilibrium electrode potential at 25C can be expressed as
E = 0.98 0.236pH 0.0886 logFe2+

Figure 1. Anodic potentiodynamic curve of the titanium substrate in the iron


oxide electrodeposition bath: 0.01 M FeSO4NH42SO46H2O and 0.04 M
CH3COOK, pH = 6.06.5, at 90C. Scan rate: 3 mV/s. The oxidation features are indicated by lower case letters; see the text.

number of electrons exchanged. In this equation n = 2 when


deposition of Fe3O4 takes place, and n = 1 when -FeOOH or
-FeOOH is being electrodeposited.
Powder X-ray diffraction spectra of films were recorded with a
Siemens D5000 diffractometer using Cu K radiation. The accelerating voltage was set at 40 KV with a 25 mA flux. Scatter and
diffraction slits of 1 and 0.1 mm collection slit were used.
Scanning electron microscopy SEM pictures were obtained
with a JEOL JSM-5410 apparatus. Transmission electron microscopy TEM and selected area electron diffraction SAED were
carried out with a Philips CM-200 microscope operated at 200 kV.
Specimens for TEM were prepared by removing the electrodeposited material by grating with a scalpel, collected, and ultrasonically
dispersed in 1 ml of ethanol. A small drop of the suspended solution
was placed onto a porous carbon film supported on a TEM nickel
grid, and was dried in air prior to observation.
XPS was carried out with a PHI 5700 equipment using a standard X-ray source 15 KV, 300 W, Mg K 1253, 6 eV. The pressure in the chamber was about 107 Pa. Binding energies BE were
referenced to the C 1s peak at 284.8 eV. Spectra were handled by
PHI-ACCESS V.6 and MULTIPAK software, both from Physical Electronics PHI, Eden Prairie, MN.
Results and Discussion
Electrochemical study. Different iron compounds can be obtained by chemical oxidation of Fe2+ salts in water solutions depending on the solution concentration, temperature, pH, as well as
starting materials.1 It is well known that inorganic anions like chloride, phosphate, or sulfate have a strong influence on final product
characteristics. In the same way different iron compounds could be
produced by electrochemical oxidation of Fe2+ ions given that several iron oxides-oxyhydroxides show similar stability regions in
potential-pH diagrams. In order to study the electrochemical processes taking place during the electrodeposition of the iron oxides,
and to check the voltage domain of each of them in our experimental
conditions, potentiodynamic experiments have been done. Figure 1
shows a typical potentiodynamic curve of the titanium substrate in
the electrodeposition electrolytic bath at 90C and pH 6. The sweep
started at the open-circuit potential and was scanned anodically at
3 mV/s. Several oxidation features indicated in Fig. 1 as a, b, c, and
d, can be observed, suggesting the formation of different phases
depending on the applied potential. Structural analysis of electrode-

with its corresponding equilibrium boundary line that shifts to the


lower-pH side in the Pourbaix diagram as the bath temperature is
increased, resulting in a thermodynamic potential at 90C of
0.63 V vs Ag/AgCl.20 As can be seen in Fig. 1, the anodic current
onset was at approximately 450 mV, followed by the process a
that can be assigned to oxidation of Fe2+ to magnetite. The difference between thermodynamic data and anodic reaction could be due
to the overpotential which is necessary to active this reaction on
titanium. This electrochemical process becomes diffusion controlled
at a potential around 300 mV, and with a limiting current density
of about 0.3 mA/cm2.
From a thermodynamic point of view and according to the Pourbaix diagram, only a narrow window leads to the deposition of
Fe3O4, and at more anodic potentials other oxidation processes
would occur resulting in different FeIII oxides. X-ray diffraction
XRD characterization see the next section allows us to associate
the sharp increase in anodic current at potentials more positive than
250 mV with the beginning of the deposition of goethite, presenting an anodic peak at 130 mV see feature b in Fig. 1. In fact,
according to potential-pH diagrams,33 the growth of -FeOOH films
would proceed via the oxidation of Fe2+ by the reaction
Fe2+ + 2H2O FeOOH + 3H+ + e

presenting the following equilibrium potential at 25C


E = 0.89 0.177pH 0.059 logFe2+

which is 240 mV more positive than the potential of the oxidation


reaction of magnetite formation. This is in good accordance with
features observed in the voltammogram and XRD characterization.
Finally, the growth of -FeOOH films would procced via the
oxidation of of Fe2+ by the reaction
Fe2+ + 2H2O FeOOH + 3H+ + e

which presents an equilibrium potential 100 mV more positive than


Reaction 3.32 The narrow window observed for the electrodeposition
of goethite pure films see below agrees with this result, and the
oxidation peak at 620 mV observed in the voltammogram process c
in Fig. 1 is associated with lepidocrocite deposition.
At further, more positive potentials in the linear sweep voltammogram, the current onset at 1.28 V is attributed to the reaction of
O2 evolution process d in Fig. 1
2H2O O2 + 4H+ + 4e

During the experiment the corresponding reaction at the cathode


would be the hydrogen evolution reaction
2H+ + 2e H2gas

with gas evolving on the cathode surface, as was observed.


After obtaining the potentiodynamic curve of Fig. 1, a set of
films was deposited potentiostatically on Ti substrates at various
deposition potentials in the range 500 to 1000 mV with respect to

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Journal of The Electrochemical Society, 154 3 D126-D133 2007

Figure 2. Representative XRD patterns of iron oxide films deposited onto a


titanium substrate at different applied potentials as indicated. JCPDS patterns
of magnetite and goethite are included for comparison. Only those diffraction
planes of interest are labeled in the JCPDS goethite pattern. M magnetite;
G goethite, and T titanium substrate.

Ag/AgCl electrode, in order to study the film properties. The pH of


the bath and the charge passed during deposition of each film were
monitored.
A trend to a decrease in pH was observed in all the electrodepositions carried out in accordance with the proposed Reactions 1, 3,
and 5. In fact, according to those reactions the whole H+ balance
would always be the increase of H+ concentration. So, to maintain
the bath pH equal to 6, KOH solution was added during the potentiostatic film electrodeposition experiments.
Structural characterization: XRD analysis. The influence of
the electrodeposition potential on the phase composition of the iron
oxide films was studied by XRD. Figures 2-4 show representative
XRD patterns of films obtained at different electrodeposition potentials. XRD results show that different iron oxide phases can be
obtained depending on the applied potential. Comparison of these
patterns with JCPDS file cards suggests that the obtained phases can
be assigned to magnetite Fe3O4,34 goethite -FeOOH,35 and
lepidocrocite -FeOOH,36 which is discussed in more detail, vide
infra.
Figure 2 shows typical XRD patterns for iron oxide films deposited at E = 340, 270, and 180 mV, which are representative for
films deposited in the potential range from 450 to 180 mV. At
E = 340 mV, besides the titanium peaks originated from the substrate, the XRD pattern does not reveal the presence of any other
phase than single-phase Fe3O4. The results also show that the magnetite film is polycrystalline, where the crystallites are randomly
oriented. Moreover, the very sharp and intense XRD diffraction
peaks are an indication of very well crystallized layers, better than
similar films obtained also by electrodeposition, and other lowtemperature techniques low-temperature reactive sputtering, sol-gel
method, etc..30 When the growth deposition potential was

D129

Figure 3. Representative XRD patterns of iron oxide films deposited onto a


titanium substrate at different applied potentials as indicated. JCPDS patterns
of goethite and lepidocrocite are included for comparison. Only the 002
diffraction plane of interest is labeled in the JCPDS lepidocrocite pattern. G
goethite; L lepidocrocite, and T titanium substrate.

270 mV, new XRD peaks corresponding to goethite appear, indicating then the formation of a mixture of phases: magnetite goethite. Despite the presence of this new goethite phase, the diffraction
pattern corresponding to the magnetite has not changed, maintaining
the peaks relative intensity ratios. At this potential, the goethite
phase is polycrystalline, presenting a slight 111 preferred orientation. When the electrodeposition potential has been moved anodically to E = 180 mV, the magnetite phase disappears, and the film
is constituted exclusively by a single goethite phase. The corresponding diffractogram shows a decrease in the 110 peak intensity,
and a very important increase of 021 and 002 diffraction peaks.
Then, the electrodeposited goethite thin films have a polycrystalline
nature with a 002 crystallographic preferred orientation.
Further moving the growth potential toward more anodic values,
the corresponding typical XRD patterns for films deposited at
E = 130, 100, 75, and 0 mV are presented in Fig. 3. These
patterns are representatives of the structural behavior of films deposited in the 180 to 0 mV potential range. The films electrodeposited
at E = 130 and 100 mV are constituted by a single-phase goethite, having a polycrystalline structure. The film deposited at
100 mV shows a further increase in the 002 peak intensity, becoming the film with a very definite and strong 002 preferred
orientation. Moreover, the sharpness of the diffraction peaks indicates a very high crystallinity of these goethite thin films. When the
applied potential is 75 mV, the appearance of new diffraction
peaks that have been assigned to lepidocrocite has been observed,
with two phases coexisting: goethite and lepidocrocite. Continuing
this trend, it can be observed that at E = 0 mV the films present
lepidocrocite and goethite as main and secondary phases, respectively.
Figure 4 shows that from 200 to 1000 mV, -FeOOH is the only

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

D130

Journal of The Electrochemical Society, 154 3 D126-D133 2007

Figure 5. Scheme of the different iron oxide phases obtained as a function of


the electrodeposition potential.

Figure 4. Representative XRD patterns of iron oxide films deposited onto a


titanium substrate at different applied potentials as indicated. The XRD pattern for lepidocrocite powder detached from a previously electrodeposited
film is also shown. JCPDS pattern of lepidocrocite is included for comparison. Only the 002 diffraction plane of interest is labeled in the JCPDS
lepidocrocite pattern. L lepidocrocite and T titanium substrate.

single phase present in the films. Moreover, the presence of a very


high 002 diffraction peak in the pattern of the films deposited at
potentials more positive than +200 mV indicates a very strong preferred orientation, with the 002 plane parallel to the substrate. A
002 crystallographic preferred orientation on electrodeposited lepidocrocite films has also been reported by Antony et al.29 Moreover,
as in the case of magnetite and goethite thin films see above, the
very narrow diffraction peaks reveal very good crystallinity nature
of these films. As the diffraction patterns of this lepidocrocite films
were so different with respect to the JCPDS ones, it was very difficult to index them. So, in order to confirm that -FeOOH was the
only phase present in the films obtained at potentials more positive
than +200 mV, several films deposited in this potential range were
removed from the corresponding substrate. Afterward, an XRD
analysis of the resulting powder was carried out; the corresponding
spectra are shown in Fig. 4. Then, comparing the resulting diffraction pattern with the JCPDS one, it can be seen that the crystalline
phase can be indexed as -FeOOH.
Finally, and in order to resume all these XRD data, a phase
diagram scheme showing the different iron oxide phases obtained as
a function of the electrodeposition potential is shown in Fig. 5.
Morphological study. All the films prepared were adherent
and homogeneous, without showing any cracks. The films identified
by XRD as pure phases of magnetite, goethite, and lepidocrocite
were in appearance black, yellowish-brown, and reddish-brown, respectively.
The morphology of the films generated after electrodeposition at
different potentials was studied by scanning electron microscopy
SEM and TEM. SEM micrographies performed on the samples

electrodeposited at 340, 100, and +600 mV corresponding to


pure phases of magnetite, goethite, and lepidocrocite, respectively,
are shown in Fig. 6. All the examined films showed well-faceted
crystals, more typical for a kinetic rather than diffusion-limited electrochemical growth process. Figure 6a1 and a2 shows the SEM images of the top and cross-sectional views of a typical film of magnetite obtained at E = 340 mV during 1 h after a charge of
0.48 C/cm2 has been passed. These images revealed a dense and
continuous film composed of well-faceted polyhedral crystallites
randomly oriented, which is consistent with XRD results. The length
of the largest polyhedral edge is about 500 nm. Observe also the
growth of smaller polyhedra on the faces of the biggest crystals.
Using the reported density value of bulk magnetite
= 5.2 g/cm3, a film thickness of 1.1 m was estimated, in good
accordance with the cross-sectional SEM image. This type of morphology, shown in Fig. 6a, has also been observed in magnetite thin
films deposited at other potentials within its corresponding pure
phase electrodeposition potential window, but different than
340 mV. Magnetite films formed by polyhedral crystals were also
reported in a previous work.9
Figure 6b1 and b2 show the SEM images of the top and crosssectional views of a film of goethite obtained at E = 100 mV during 30 min with a deposition charge of 1.65 C/cm2. This 001 oriented film, according to XRD results, is the aggregation of
rhomboidal columns with some little gaps between them. Each individual column grows normal to the substrate surface. The SEM
micrographs show over the whole surface of the electrode a deposit
made of crystals exhibiting well-defined edges of about 1 m. To
the best of our knowledge, similar film structures have not been
previously reported for goethite. In the Peulon et al. work,9 the
goethite films exhibits acicular crystals not preferentially ordered,
which correspond to one of the typical shapes of goethite particles.1
In the present work, a needle-lath-like shape can be observed, with
the goethite needles-laths very well ordered and their length easily
controlled by the electrodeposition time. The film thickness obtained
using the bulk density of goethite = 4.27 g/cm3 was 3.6 m,
slightly smaller than the one observed in Fig. 6b2. This can be due
to a less compact morphology of the films, as can be appreciated in

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Journal of The Electrochemical Society, 154 3 D126-D133 2007

D131

Figure 7. TEM images showing different iron oxide particle morphologies


and their related SAED patterns. a magnetite; b goethite, and c lepidocrocite.

Figure 6. SEM images of electrodeposited iron oxide thin films, plane and
cross-sectional views, 1 and 2, respectively. a1, a2 magnetite; b1, b2
goethite; c1, c2 lepidocrocite.

Fig. 6b1 and b2. The goethite films which have been grown at potentials more cathodic than 100 mV present the same columnar
morphology, but with more rhomboidal columns and not as well
normal to the surface substrate as in the case of those obtained at
E = 100 mV.
Figure 6c1 and c2 shows the SEM images of the top and crosssectional views of an electrodeposited lepidocrocite thin film obtained at E = +600 mV, during 30 min with a charge transfer of
3.5 C/cm2. The crystals are lathlike, elongated in the c-direction,
with each individual lath grown normal to the substrate surface. This
is the same preferred orientation present in the lepidocrocite films
obtained by Antony et al.29 When comparing both works, an important improvement in particle crystallinity is observed here, probably
promoted by the higher bath temperature used in the present work,
90C, against 55C used by Antony et al. In the top view the lepidocrocite laths present a shape more planar than that presented by
the goethite laths above described. The film thickness obtained using
the bulk density of lepidocrocite = 3.97 g/cm3 was 8.2 m,
slightly smaller than the one observed in Fig. 6c2, which agrees with
the less compact morphology of the films, as can be appreciated in
Fig. 6c1 and c2. All the electrodeposited lepidocrocite films showed
similar morphological structures similar to the one described here
for the film deposited at E = 600 mV.
Further structural characterization by TEM was performed for
the same samples for which SEM images have been shown above.
Figure 7 shows typical TEM images and its SAED patterns. TEM
observations show crystals of different morphologies: cubes for
magnetite, and lathlike for goethite and lepidocrocite particles. The
spot diffraction patterns of the magnetite, goethite, and lepidocrocite
particles revealed that they are single-crystalline, which is in good
agreement with the XRD results.

XPS characterization. As an additional test to verify the presence of magnetite rather than maghemite -Fe2O3, which has an
almost identical XRD diffraction pattern, and to confirm the presence of the goethite and lepidocrocite phases, XPS was used to
examine and elucidate these points.
XPS Fe 2p, O 1s, and O 2p photoelectron and valence band
spectra of the Fe3O4, -FeOOH, and -FeOOH samples are shown
in Fig. 8. As can be seen from the XPS spectra, clear features permit
us to distinguish between Fe3O4 and the oxyhydroxide phases. Two
different contributions of photoelectrons are visible in the Fe 2p
spectrum of the Fe3O4 sample. They are seen in the Fe 2p3/2 peak at
binding energies of 710.8 eV as peak maximum and 708.3 eV as a
shoulder. These contributions can be assigned to Fe3+ and Fe2+
states of magnetite.37-39 Similarly, two contributions can be discerned in the Fe LMM Auger peak of the Fe3O4 sample at 700.6 and
703.3 eV of kinetic energy Auger spectra not shown. Distinct from
that, -FeOOH and -FeOOH samples show only one contribution
Fe3+ states, found at 711.3 eV in binding energy for the Fe 2p3/2
peak and around 700.6 eV in kinetic energy for Fe LMM not
shown. Furthermore, the absence of a satellite peak in the Fe 2p
spectrum of the Fe3O4 sample is characteristic for magnetite.40,41
Moreover, the lack of this satellite peak at ca. 719 eV, which is a
major characteristic of Fe3+ in -Fe2O3, clearly excluded the formation of maghemite in the electrodeposited magnetite films. This Fe
2p3/2 satellite peak is found in the - and -FeOOH samples at
719.3 eV, similar to other iron oxide samples of only Fe3+ states.42
We observe in the O 1s spectra of Fig. 8 two contributions in the
case of the oxyhydroxide samples, more clearly separated in the
-FeOOH sample. They are found at 530.0 eV corresponding to
FeO bonds and at 531.1 eV due to FeOH bonds.37,38 In the O 1s
spectrum of the Fe3O4 sample, the O 1s peak has its maximum at
530.0 eV due to FeO bonds. However, at higher binding energies a
broad shoulder can be seen. These are contributions due to CO
bonds, because this sample reveals an inherent higher carbon surface
contamination about 35 atom % compared to about 15 atom % for
-FeOOH and 10 atom % for -FeOOH samples. Due to that adventitious surface contamination of the samples, XPS atomic concentrations could not be used to confirm the stoichiometry of magnetite or iron oxyhydroxide phases. After recording the spectra
shown in Fig. 8, sample surface cleaning in the spectrometer by
4 keV Ar+ ion etching was carried out. Then, only Fe and O atoms
were detected, confirming the purity of the obtained iron oxides.
With regard to stoichiometry, Ar+ sputtering in the case of iron
oxides leads readily to ion beam induced chemical reduction, i.e.,
preferential sputtering of oxygen, thus altering the chemical states
and composition of the surface under analysis.39 The features observed in the O 1s spectra of the three samples in its original state,
before sputtering are also visible in the O 2p peak, which is shown
together with the valence band region in Fig. 8. The valence band
spectra as well clearly show the difference between iron oxyhydrox-

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

D132

Journal of The Electrochemical Society, 154 3 D126-D133 2007

Figure 8. XPS Fe 2p, O 1s, and O 2p photoelectron and valence band spectra of the Fe3O4, -FeOOH, and -FeOOH samples. Characteristic binding energies
for magnetite and Fe hydroxide are indicated. The arrow indicates the position of the Fermi level EF.

ide and magnetite phases. In the case of the Fe3O4 sample, the shape
of the XPS valence band spectrum is a typical fingerprint of
magnetite.43 Photoelectrons are detected up to energies of the Fermi
level. This shows the property of electrical conductance in magnetite, and further clearly differentiates this phase from the oxyhydroxide ones.
Conclusions
Reproducible, homogeneous, and well-adherent magnetite, goethite, and lepidocrocite thin films were electrodeposited onto titanium substrates. These different phases were obtained on the basis
of tuning the electrodeposition potential and employing the same
electrolytic bath. Three different electrodeposition potential windows for the obtainment of the iron oxide pure phases have been
identified. The very sharp and intense XRD diffraction peaks are an
indication of the very well crystallized layers. Polycrystalline magnetite thin films without a defined texture, and goethite and lepidocrocite thin films with a very strong 002 crystallographic preferred orientation have been obtained. The iron oxide thin films
present well-crystallized particles as exhibited by the SEM images:
magnetite films presenting well-faceted polyhedral crystallites randomly oriented; goethite layers showing an aggregation of rhomboidal columns with some little gaps between them, and lepidocrocite
films presenting lathlike crystallites. The magnetite phase has been
identified due to typical XPS features as two Fe 2p contributions
corresponding to Fe3+ and Fe2+ states, absence of Fe 2p satellite
peaks and a fingerprint valence band shape. The hydroxide phase
becomes evident by O 1s contributions due to FeOH bonds.
Acknowledgments
The authors are grateful to Universidad de Mlaga, Mlaga,
Spain, for the financial support received. E.A.D. also acknowledges
the financial support of PEDECIBA-Fsica, and especially to
C.S.I.C., Universidad de la Repblica, Montevideo, Uruguay. The
authors also thank A. Martnez Orellana Mlaga, Spain for the
TEM measurements.
Departamento de Fsica Aplicada assisted in meeting the publication
costs of this article.

References
1. U. Schwertmann and R. M. Cornell, in Iron Oxides in the Laboratory, Wiley-VCH,
Weinheim, 2000.

2. G. Zhang, C. Fan, L. Pan, F. Wang, P. Wu, H. Qiu, Y. Gu, and Y. Zhang, J. Magn.
Magn. Mater., 293, 737 2005.
3. S. Soeya, J. Hayakawa, H. Takahashi, K. Ito, C. Yamamoto, A. Kida, H. Asano,
and M. Matsui, Appl. Phys. Lett., 80, 823 2002.
4. F. Walz, J. Phys.: Condens. Matter, 14, R285 2002.
5. J-Y. Lee, S. J. Oh, J. G. Sohn, and S.-J. Kwon, Corros. Sci., 43, 803 2001.
6. N. Gustos, G. J. Papadopoulos, V. Likodimos, S. Patapis, D. Yarmis, A. Przepiera,
K. Prezepiera, J. Majszczyk, J. Typek, M. Wabia, K. Aidinis, and Z. Drazek, Mater.
Res. Bull., 37, 1051 2002.
7. G. Jain, C. J. Capozzi, and J. J. Xu, J. Electrochem. Soc., 150, A806 2003, and
references therein.
8. K. Kanamura, H. Sakaebe, and Z. Takehara, J. Power Sources, 40, 291 1992.
9. S. Peulon, H. Antony, L. Legrand, and A. Chausse, Electrochim. Acta, 49, 2891
2004.
10. T. Maruyama and T. Kanagawa, J. Electrochem. Soc., 143, 1675 1996 and references therein.
11. J. B. Yang, S. K. Malik, X. D. Zhou, M. S. Kim, W. B. Yelon, W. J. James, and H.
U. Anderson, J. Phys. D, 38, 1215 2005.
12. K. J. Kim, D. W. Moon, S. K. Lee, and K.-H. Jung, Thin Solid Films, 360, 118
2000.
13. K. Shalini, G. N. Subbanna, S. Chandrasekaran, and S. A. Shivashankar, Thin Solid
Films, 424, 56 2003.
14. K. W. Chung, K. B. Kim, S.-H. Han, and H. Lee, Electrochem. Solid-State Lett., 8,
A259 2005.
15. M. P. Nikiforov, A. A. Vertegel, M. G. Shumsky, and J. A. Switzer, Adv. Mater.
(Weinheim, Ger.), 12, 1351 2000.
16. H. Gmez, R. Henrquez, R. Schrebler, G. Riveros, D. Leinen, J. R. RamosBarrado, and E. A. Dalchiele, J. Electroanal. Chem., 574, 113 2004.
17. S.-H. Yu and M. Yoshimura, Adv. Funct. Mater., 12, 9 2002.
18. M. Yoshimura, W. L. Suchanek, and K. Byrappa, MRS Bull., 25, 17 2000.
19. K. W. Chung, K. B. Kim, S.-H. Han, and H. Lee, J. Electrochem. Soc., 152, C560
2005.
20. T. A. Sorenson, S. A. Morton, G. D. Waddill, and J. A. Switzer, J. Am. Chem. Soc.,
124, 7604 2002.
21. J. L. Leibenguth and M. Cohen, J. Electrochem. Soc., 119, 987 1972.
22. K. Hashimoto and M. Cohen, J. Electrochem. Soc., 121, 37 1974.
23. J. W. Schulze, S. Mohr, and M. M. Lohrengel, J. Electroanal. Chem. Interfacial
Electrochem., 154, 57 1983.
24. S. Ardizzone and L. Formaro, J. Electroanal. Chem. Interfacial Electrochem., 246,
53 1988.
25. M. Seo, K. Yoshida, H. Takahashi, and I. Sawamura, J. Electrochem. Soc., 139,
139 1992.
26. T. Ohtsuka, J. C. Ju, S. Ito, and H. Einaga, Corros. Sci., 36, 1257 1994.
27. M. Abe and Y. Tamaura, Jpn. J. Appl. Phys., Part 2, 22, L511 1983.
28. K. Nishimura, Y. Kitamoto, and M. Abe, IEEE Trans. Magn., 35, 3043 1999.
29. H. Antony, S. Peulon, L. Legrand, and A. Chauss, Electrochim. Acta, 50, 1015
2004.
30. D. Carlier, C. Terrier, C. Arm, and J.-Ph. Ansermet, Electrochem. Solid-State Lett.,
8, C43 2005, and references therein.
31. S.-Y. Wang, K.-C. Ho, S.-L. Kuo, and .N-L. Wu, J. Electrochem. Soc., 153, A75
2006.
32. M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, Pergamon
Press, New York 1996.
33. J-M.-R. Genin, G. Bourri, F. Trolard, M. Abdelmoula, A. Jaffrezic, P. Refait, V.
Maitre, B. Humbert, and A. Herbillon, Environ. Sci. Technol., 82 1058 1998.
34. JCPDS Magnetite, File 19629.

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Journal of The Electrochemical Society, 154 3 D126-D133 2007


35.
36.
37.
38.
39.
40.

JCPDS Goethite, File 810464.


JCPDS Lepidocrocite, File 080098.
N. S. McIntyre and D. G. Zetaruk, Anal. Chem., 49, 1521 1977.
D. Brion, Appl. Surf. Sci., 5, 133 1980.
P. Mills and J. L. Sullivan, J. Phys. D, 16, 723 1983.
S. Gota, J. B. Moussy, M. Henriot, M. J. Guittet, and M. Gauter-Soyer, Surf. Sci.,

D133

482/5, 809 2001.


41. C. Ruby, J. Fusy, and J. M. R. Gnin, Thin Solid Films, 352, 22 1999.
42. K. Wandelt, Surf. Sci. Rep., 2, 1 1982.
43. J. F. Moulder, N. F. Stickle, P. E. Sobol, and K. D. Bomben, Handbook of X-ray
Photoelectron Spectroscopy, J. Chastain, Editor, p. 251, PHI Perkin Elmer Corp.,
Eden Prairie, MN 1995.

Downloaded 25 May 2012 to 152.3.102.242. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Вам также может понравиться