Você está na página 1de 10

H132

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

Study of the Electrochemical Activities of Mo-Modified Pt


Catalysts, for Application as Anodes in Direct Methanol Fuel Cells:
Effect of the Aggregation Route
O. Ugalde-Reyes,a R. Hernandez-Maya,a A. L. Ocampo-Flores,a F. Alvarez- Ramrez,b
E. Sosa-Hernandez,b C. Angeles-Chavez,b and P. Roqueroa,,z
a Facultad
b Instituto

de Qumica, Universidad Nacional Autonoma de Mexico, D. F. CP 04510, Mexico


Mexicano del Petroleo, D. F. CP 07730, Mexico

Electrocatalysts with constant metallic composition, consisting of carbon-supported platinum and molybdenum phases, were synthesized following the thermolysis (thr) and borohydride (bhr) reduction methods and using different metallic precursors. The
obtained electrocatalysts were characterized by X-ray energy-dispersive spectrometry, X-ray diffraction and high resolution transmission electron microscopy. Their activities were studied by cyclic voltammetry. Different surface structures were obtained and
the electrochemical activities toward methanol oxidation were compared. Pt, MoO2 and MoO3 phases were well identified with
the characterization techniques used. However, the electrochemical responses obtained from both sample series were considerably
different, suggesting that the arrangement and relationships between active phases strongly depend on the synthesis method and
aggregation sequence of the metallic precursors, and being the cause of different catalytic activities and stabilities of molybdenum
oxide phases. The bhr method offered higher activity than the thr method. Among the sample series obtained by bhr method, the
catalyst obtained by platinum deposition on the previously synthesized molybdenum on carbon, led to the highest overall activity.
The Author(s) 2014. Published by ECS. This is an open access article distributed under the terms of the Creative Commons
Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/),
which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is not changed in any
way and is properly cited. For permission for commercial reuse, please email: oa@electrochem.org. [DOI: 10.1149/2.0521503jes]
All rights reserved.
Manuscript submitted November 5, 2014; revised manuscript received December 16, 2014. Published December 29, 2014.

Fuel Cells represent one of the most promising options for generating electricity with high efficiency and low environmental impact. The Direct-Methanol Fuel Cell (DMFC) is a low-temperature
cell where methanol is directly supplied as fuel to the anode compartment, without any previous reforming. Furthermore, methanol
can be stored and transported in liquid phase, it can be obtained
from biomass, and its complete oxidation can yield a high energy
density.1
Platinum has been recognized as the most active catalyst for the
methanol oxidation reaction and it has been employed in the formulation of such materials. However, the formation of different organic
intermediates, adsorbed on Pt surfaces during methanol electrooxidation, causes the poisoning of active sites and decreases the
catalyst efficiency. Several studies indicate that the modification
of the Pt catalyst with other transition metal can lead to good
results in oxidizing methanol at a low potential (lower than in
pure Pt).2
On the other hand, it is well known that bimetallic materials show
increased activity toward the CO electro-oxidation reaction, yielding
CO2 . Studies on catalytic effects that take place on bimetallic surfaces,
obtained from pure metals and modified by the deposition of a second
metal, are closely related to the technological development of low
temperature fuel cells. Some researchers have emphasized the use of
CO as a test molecule to show the electronic effect associated with
this reaction. The CO desorption energy is apparently related to strong
intermetallic bonds and mixed orbitals of the active phases. Hence, a
catalyst with high activity toward methanol oxidation must also have
high activity for CO oxidation.3 This activity is explained by some
authors in terms of the so-called bifunctional mechanism, in which
the formation of reactive oxygen-containing species takes place at a
lower potential than in pure Pt, promoting the complete oxidation of
CO.2,4 However, other researchers propose that an electronic effect is
responsible for the increased activity.4 In this case, the presence of a
second metal modifies and weakens the Pt-organic species bond, rendering the CO adsorption strength on these materials, different from
that in monometallic catalysts.5 Other theories suggest the existence

Electrochemical Society Active Member.


E-mail: roquero@unam.mx

of a hydrogen spillover effect. In this phenomenon, a proton (H+ )


migrates from a Pt site to form metallic bronzes in the promoting
phase and leaves Pt active sites available for adsorption of a methanol
molecule.6,7 Moreover, it has been suggested that, by modifying the
surface of a Pt catalyst with metallic molybdenum or MoO3 species,
the chemisorption of CO on Pt surfaces can be partially suppressed
due to a physical blocking effect of some Pt sites by molybdenum,
allowing the presence of free active sites for the oxidation reaction
and preventing the adsorption of CO at the crystalline platinum plane
on which it most strongly adsorbs.8
Different synthesis methods and characterization techniques for
these types of catalysts have been proposed.9 Among the commonly
used methods to synthesize these materials are: the thermolysis of
metal carbonyl compounds;10 the chemical reduction to obtain small
particles at low temperatures (below 100 C), and the support treatment to achieve optimal performance.11,12 Recent studies on bimetallic, carbon-supported PtMo electrocatalysts, prepared via conventional impregnation and NaBH4 reduction method, showed that the
behavior of these materials toward CO oxidation can also be described reasonably well with the classic bifunctional mechanism.13
Furthermore, it has been demonstrated that the chemical reduction
method produces a core-shell structure with better catalytic properties
for the oxidation of methanol and intermediates, and a higher CO
tolerance.1416 Thermolysis is a commonly used method of preparing Pt-based catalysts from carbonyl complexes (precursors), which
produces very disperse particles on the support with large active
areas that favor both, the adsorption and the oxidation of fuels
and intermediates.17 These methods have shown that the activity of
catalytic materials depends on the synthesis sequence and catalyst
parameters.
Therefore, in this work, we prepared carbon-supported Pt and Mo
catalysts, changing the aggregation sequence of the metallic precursors with two preparation methods. One series of these materials
was prepared by thermolysis of carbonyl precursors method. The
second series was obtained by a chemical reduction method, using
sodium borohydride as reductant, and chloride salts as the active
phases precursors. This study is aimed at finding the existing relationship between structure, electrochemical activity and bimetallic catalyst stability, by varying the method and sequence of metal precursor
aggregation.

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)


Experimental
Catalysts synthesis. Samples of Pt and Mo deposited on carbon
support were prepared using different aggregation sequences of metallic precursors. The reduction process was performed by two methods:
thermolysis (thr) and borohydride reduction (bhr). H2 PtCl6 (Aldrich),
Mo(CO)6 (Aldrich), and MoCl3 (Aldrich) were the chemical species
used as metal precursor sources. The Pt hexacarbonyl was obtained
by bubbling CO, at 25 cm3 min1 , during 24 h in aqueous H2 PtCl6
solution.10,18,19
In the thr method, the Mo(CO)6 and Pt carbonyl species were
added to 1,2-dichlorobenzene and subsequently, the Vulcan XC-72R
carbon was added to the solution. In order to impregnate the Mo and
Pt carbonyl species onto the carbon surface, the temperature of the
system was increased up to 110 C and maintained during 4 h with
vigorous stirring. After this stage, the solvent was filtered, the solid
washed twice with diethyl ether and placed under a N2 atmosphere at
400 C during 2 h.
In the bhr process, H2 PtCl6 and MoCl3 salts were added to a glass
flask containing water which was stirred until the metallic salts were
dissolved. Subsequently, the Vulcan XC-72R carbon was added. The
reduction process was performed by dropwise addition of an aqueous
solution of 1.0 M NaBH4 to the suspension containing dissolved
precursors and the carbon support, at 60 C. Stirring was kept during
5 hours until the addition of borohydride solution was finished. The
solids were then filtered, washed with deionized water and dried at
120 C for 1 hour.
Four aggregation sequences of the metallic species were carried
out. First, both Mo and Pt species were simultaneously added to the
container and then reduced (Pt-Mo/C). The second sequence, in both
processes, consisted of first obtaining the reduced Pt/C and then adding
the Mo species to the reduced Pt/C solution (Mo/Pt/C). In the third
sequence, the addition order was changed: first Mo/C was synthesized
and the Pt species was subsequently added to the Mo/C suspension
(Pt/Mo/C). The final sequence was physical mixing of reduced Pt/C
and Mo/C solids (Pt+Mo/C). These samples were mechanically mixed
in a diethyl ether suspension during 5 h at room temperature. The final
solids were washed with deionized water and dried at 120 C for 1 hour.
All bimetallic samples were prepared to contain 12.5 wt% of metallic loading, at a Pt:Mo atomic ratio of 4:1. Mo and Pt monometallic
samples at a concentration of 12.5 wt% were also prepared.
In previous works, we have found that there is an optimum
Mo concentration in the catalysts that leads to high electrochemical activities.18,20 Too low or too high Mo loadings lead to poor results
for methanol oxidation. In this work we only present results for the
formulation that has been shown to yield the highest activities.
Pt-Ru catalysts have long been considered the best performing
materials for the anodic methanol oxidation.21,22 The comparison of
PtRu and PtMo catalysts, carried out by current-sampled voltammetry, shows that the formulations yielding the best results had similar
activities for both metals, Mo and Ru. However, It has been shown
that Ru anodes undergo corrosion, dissolution, and the crossover of
Ru cations to the cathode.23 Ruthenium is not a low-cost metal, and
its place as preferred co-catalyst for DMFCs is in decline.

Characterization by physical techniques. The chemical composition of the samples was obtained by X-ray energy-dispersive spectroscopy (XEDS). The spectroscope was attached to a JEOL JSM5900LV scanning electron microscope, operating at 20 keV during
the tests. The metallic phases were evidenced by X-Ray Diffraction
(XRD). The measurements were carried out at room temperature using
Cu K ( = 1.5406 ) radiation on a Siemens D-500 diffractometer,
at a 1.58 min1 rate. The metal particle size and structural features
were obtained by transmission electron microscopy (TEM). The microscope used was a JEM2200FS operated at 200 kV. The samples
were dispersed in isopropanol and stirred in an ultrasound bath for
10 min. The supernatant liquid was placed on 200-mesh copper grids
covered with carbon.

H133

Electrochemical measurements. Electrochemical cell. The


experimental device used to perform the electrochemical measurements consisted of a typical 3-electrode cell with volume of 50 cm3 .
The 3-electrode array included a saturated calomel reference electrode (Hg/Hg2 Cl2 /sat. KCl), a 0.5 cm diameter graphite bar, used as
auxiliary electrode, and the catalyst placed on a 2.5 mm diameter disc
as a working electrode. The working electrode was prepared by using
the methodology described below.
Working electrode preparation. A type of ink containing
0.1 mg of each catalyst dispersed in a 10 L 2-propanol, 10 L
Nafion resin, was prepared. The dispersion was put in an ultrasound
bath during 15 min and 2.5 L of the ink were placed on the tip of
a commercial vitreous carbon disk electrode with the aid of a micropipette. The solvent was evaporated at room temperature and a
catalytic solid layer was obtained. The catalyst content in this electrode was 0.25 mg cm2 , related to the geometrical surface area of the
disk (0.25 cm diameter). Cyclic voltammetry experiments were carried out at 25 C in the three-electrode cell as described in the previous
paragraph, using the catalyst layer as working electrode.
A 0.5 M H2 SO4 aqueous solution was employed as supporting
electrolyte in all experiments. The working solution was a 1.0 M
methanol, 0.5 M H2 SO4 electrolyte. All electrolytic media were degassed by bubbling nitrogen during 15 min prior to each experiment.
Voltammetry results presented here were obtained at a potential scan
rate of 50 mV s1 , with an Autolab potentiostat. Before each experiment, electrodes were activated by applying 35 potential cycles,
between 250 and 1100 mV vs. SCE in a sulfuric acid solution.
The open circuit potential was measured and its value was used to
start all potential sweeps. Scanning was carried out first toward negative potentials, up to 0.25 V vs. SCE, and then reversed to positive
potentials. The inversion potential at positive values was changed in
some experiments.
The catalytically active area was determined by estimating the proton adsorption charge (H+ ) from the experimentally obtained cyclic
voltammograms (CVs) as follows: first, the open circuit potential
(EOCP ) was measured to obtain a reference potential of the material;
then, the electrochemical cleaning/activation of the catalyst was carried out by cycling potential in a potential window of 0.250 to 1.100
V vs. SCE. Finally, voltammetry was performed for 15 continuous
cycles at a scan rate of 50 mV s1 in a 0.5 M H2 SO4 solution.
Results and Discussion
Characterization of Mo modified Pt electrocatalysts. The elemental chemical composition of each sample is illustrated in Table I.
The concentrations obtained were near the nominal composition, but
with a slight increase in metal content, mainly that of platinum, as can
be appreciated in the average value (10.5 wt% Pt, 2.4 wt% Mo and
87.1 wt% C). However, this metal enrichment was basically observed
in the samples synthesized by thr. In average, the samples had 10.7
wt% Pt and 3.1 wt% Mo, whereas in the samples obtained by bhr

Table I. Quantitative elemental chemical analysis for Pt80:Mo20


catalysts (wt%).
Nominal
composition wt%

Real composition
by XEDS wt%

Catalyst

Pt

Mo

Pt

Mo

Pt-Mo/Cthr
Pt+Mo/Cthr
Pt/Mo/Cthr
Mo/Pt/Cthr
Pt-Mo/Cbhr
Pt+Mo/Cbhr
Pt/Mo/Cbhr
Mo/Pt/Cbhr

10
10
10
10
10
10
10
10

2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5

87.5
87.5
87.5
87.5
87.5
87.5
87.5
87.5

10.05
11.76
9.8
11.07
10.26
9.87
11.06
9.78

3.62
3
2.39
3.32
1.91
1.83
1.66
1.78

86.33
85.24
87.81
85.61
87.83
88.30
87.28
88.44

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H134

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

Table II. Particle size of bimetallic materials in comparison with


Pt. Pt active area from adsorbed H obtained from Figure 2a2b.
Catalysts

Particle size XRD (nm)

SESA a (cm2 /mg cat)

Pt/Mo/Cthr
Pt+Mo/Cthr
Pt-Mo/Cthr
Pt/Cthr
Mo/Pt/Cthr
Pt/Mo/Cbhr
Pt+Mo/C bhr
Pt-Mo/Cbhr
Pt/Cbhr
Mo/Pt/Cbhr

2.29
2.38
2.72
3.20
4.73
3.65
3.34
2.75
3.81
4.15

295.3
85.8
319.2
182.0
320.0
117.4
35
75
20.2
115

a electroactive

= Qcat / QP.

surface area (SESA ) estimated by equation, SESA

and the results obtained are displayed in Table II. The Mo/Pt/C thr
sample had a particle size of about 4.7 nm while the Pt/Mo/C thr sample had a smaller mean size (around 2.3 nm). In the samples obtained
by bhr reduction, the Mo/Pt/C bhr sample generated larger particles,
around 4.2 nm, and the Pt-Mo/C bhr sample had smaller particles,
around 2.8 nm. In both synthesis methods the aggregation sequence
labeled Mo/Pt/C produced larger particle sizes while the smaller particle sizes corresponded to Pt/Mo/C obtained by the thr method, and
to Pt-Mo/C obtained by the bhr reduction method.

Figure 1. XRD patterns of the materials synthesized by: (a) carbonyl thermolysis method. (b) Borohydride reduction method. MoO3 (JCPDS 9-0209); 
MoO2.88 (JCPDS 9-0195); Pt (JCPDS 4-0802).

reduction the metal content was 10.2 wt% Pt and 1.8 wt% Mo. Here,
the Mo concentration decreased and the metal content in the samples
was slightly lower than the nominal composition.
The crystalline phases identified in XRD patterns are displayed
in Figure 1a and 1b. All XRD patterns of the bimetallic samples
displayed diffraction peaks at 2 = 39.73, 46.21, and 67.47 degrees
which correspond to the facecentered-cubic Pt phase according to
JCPDS 40802 card. In addition, these peaks are in complete agreement with the XRD pattern of the Pt monometallic sample. The peak
located at 2 = 25 corresponds to the Vulcan carbon support. In the
sample impregnated only with Mo, two molybdenum oxide phases
were clearly identified in the XRD pattern: MoO3 (JCPDS 90209)
and MoO2.88 (JCPDS 9-0195). Therefore, formation of intermetallic
phases between Pt and Mo was not detected by XRD. The absence of
any Mo signal in the XRD pattern, but detected by XEDS, at a concentration of about 2.4 wt%, suggests that Mo concentrations below
3.6 wt% were not detectable by XRD.
On the other hand, broadening of Pt peaks in different XRD
patterns was observed. This behavior indicates that the synthesis
method and the aggregation sequence of metallic components had
a direct effect on the metal particle size. Qualitatively, the thr
method generated smaller metal particles than the bhr method and
the aggregation sequence decreased the particle size as Mo/Pt/C>PtMo/C>Pt+Mo/C>Pt/Mo/C for the thr method, whereas with bhr reduction its size was Mo/Pt/C>Pt/Mo/C>Pt+Mo/C>Pt-Mo/C. Quantitatively, the particle size was calculated using the Scherrer equation

Electrochemical behavior of Mo modified Pt electrodes. Figure


2 shows the CVs of the catalysts obtained by the thr and bhr methods,
through 4 aggregation sequences. All voltammograms were traced
in a 0.5 M H2 SO4 electrolytic medium, after having performed the
electrochemical pre-treatment of the working electrode, described in
the experimental section. The CVs have been normalized to the corresponding experimentally obtained Pt active areas of each catalyst. The
purpose of the pre-treatment is to perform structural and/or surface
modifications to the catalysts, in order to improve their electrocatalytic activity. The graphs display different potential zones where the
typical redox and charge accumulation processes are carried out at the
Mo-modified Pt nanocatalyst /electrolyte medium interface. These potential domains in Pt-based materials are well known and described
as follows: (a) cathodic region of hydrogen underpotential deposition
(H-UPD, from 0.05 to 0.250 V); (b) electrical double layer interface
region (0.05 to 0.45 V; (c) Pt oxidation potential region (0.450.950 V)
and d) Pt oxide reduction region (1 to 0.150 V). In the H-UPD region,
labeled A , anodic currents due to proton desorption from different
Pt surface planes are observed, and the corresponding cathodic currents associated with the proton adsorption and molecular hydrogen
discharge are designed A .
Typically, the voltammetric response of Pt/C electrodes in a sulfuric acid medium exhibits two cathodic and three anodic peaks in the
hydrogen region (H-UPD). The Pt/C materials obtained by thr and
bhr methods exhibit similar behaviors. The addition of Mo modifies
considerably the proton desorption-oxidation region, in the materials obtained by thr. This modification results in the disappearance
of a desorption peak (0.20 V) and broadening and displacement of
the Pt oxide reduction peak potential toward more negative values.
In materials obtained by the bhr method, the CV currents increase
with the addition of Mo. The electrical charges for hydrogen adsorption/desorption, QH , that are transferred to Pt/C catalysts or materials
modified with Mo, e.g., PtMo/C, make it possible to estimate the electroactive surface area (SESA ).24,25 The SESA value is an indicator of the
electrochemically active sites available per gram of catalyst, thereby
being an important parameter for characterizing the electrochemical
activity of Pt/C and Pt/Mo/C catalysts.
The SESA of catalysts is shown in Table II and was estimated by
the following equation, SEAS = Qcat / QPt , where Qcat is the charge
resulting from hydrogen adsorption/desorption that was quantified by

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

A'

C1'

(a)

B'

C2'

H135

C1''

-1

A''

PtMo/Cthr

B''

1
0

Current density, A/m2

-1

Pt+Mo/Cthr

1
0
-1
1

Pt-Mo/Cthr
A'

0
-1

A''

Pt/Cthr

B''

1
0
-1

MoPt/Cthr

-2
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volt vs SCE, V
10

A'

(b)

B'

Figure 2. Voltammetry characteristics of the catalysts


in 0.5 M H2SO4, at 50 mV/s. (a) carbonyl thermolysis
method. (b) Borohydride reduction method.

0
A''

Current density, A/m2

-10

B''

PtMo/Cbhr

1
0
-1
-2
10

Pt+Mo/Cbhr

0
-10
2
1
0
-1
-2
10

Pt-Mo/Cbhr
A'

A''

Pt/Cbhr

B''

0
-10

MoPt/Cbhr

-0.2

0.0

0.2

0.4
Volt vs SCE, V

0.6

0.8

including the CVs region from Figure 2; QPt is the electric charge
associated with a hydrogen monolayer adsorbed on the Pt surface,
whose value is considered to be 0.210 (mC cm2 ).26
The obtained SESA values (cm2 /mg cat.) and are downwardly
ordered as follows: Mo/Pt/C>Pt-Mo/C>Pt/Mo/C>Pt/C>Pt+Mo/C
for the materials obtained by the thr method, whereas the SESA
obtained by the bhr method were: Pt/Mo/C>Mo/Pt/C>Pt-Mo/C>
Pt+Mo/C>Pt/C. According to the parameters in Table II, there is no
relationship between particle sizes and SESA because of the different
synthesis sequences of catalysts in the two methods. The sequence
of synthesis, however, does show an effect, e.g., the SESA value for
Pt+Mo/Cthr is lower than that for Pt/C thr , due to the fact that during the
mechanical mixing there is no intimate contact between the elements,
whereas the SESA obtained for Pt/Mo/C materials is higher because Pt
is dispersed on the Mo/C surface as was evidenced by TEM-XEDS
results. In the bhr method, this value for Pt+Mo/Cbhr is only slghtly
higher than for Pt/C bhr .
In the electrical double layer region of the CV, the current density
as a function of the potential presents charge accumulation and redox
processes. The Pt/C materials only exhibit charge accumulation or

1.0

1.2

capacitive currents, whereas the other materials show, in addition to


charge accumulation, also a redox process, as observed in Figure 3
(inset). In Figure 3b, several additional peaks (C1  , C2  , and C1  ) can
be observed in detail (inset). These redox peaks may be attributed to
a Mo(VI)/Mo(IV) transformation of Mo species. The peak C1  can be
attributed to Mo (IV) to Mo (V) oxidation,20,27 while the peak C2  can
be due to Mo (V) to Mo (VI) oxidation. The corresponding cathodic
peak, C1 , is associated with the reduction of Mo (V), and the other
reduction peak C2  is not observed because it is overlapped with the
wide reduction peak of platinum oxide (fig. 2a)
In the anodic zone, the current at the B peak is associated with

the platinum oxidation that forms oxides or OH surface groups, and



the peak B corresponds to the reduction of these Pt (II) oxides to
yield metallic Pt. For Mo-modified Pt materials obtained by the thr
method, the potential of current peak B is displaced toward more cathodic values and the peak width increases (see peak B in Figure 2a).
The increase in current density in the reduction peak B , due to an
additional reduction of Mo oxides formed during the direct scan,
results in a partial passivation of the Pt surface and forms materials with greater electrical resistance. Moreover, Mo oxides seem to

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H136

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

1.0

Pt/Cthr

B'

A'

0.5
0.0

Current density, A/m2

-0.5

0.25
Pt/Cthr

-1.0

0.00

A''

-1.5

Pt/Cbhr

B''
-0.25
0.1

1.0

Pt/Cbhr

0.3

0.4

B'

A'

0.5

0.2

0.0
-0.5
-1.0
A''

-1.5

-0.2

(a)

B''

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volt vs SCE, V
PtMo/Cthr

B'

0.2

Current density, A/m2

C2'

C1'

0.1

PtMo/C thr

0.0

-1

1.5
1.0
0.5 PtMo/C bhr

10

0.0
0.1

PtMo/Cbhr

0.2

0.3

0.4

0.5

-10

(b)
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volt vs SCE, V
2

MoPt/Cthr

B'

Current density, A/m2

1
0

C1'

0.25
C2'

MoPt/C thr

-1
0.00

C2'
C1'

-2

2
1

10

MoPt/Cbrh

MoPt/C bhr

0.1

0.2

0.3

0.4

0.5

-10
(c)

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volts vs SCE, V

Figure 3. Voltammogram obtained in 0.5 M H2SO4, at 50 mV/s for the catalysts: a) Pt/C, b) PtMo/C and c) Mo/Pt/C prepared by both, thr and bhr
methods.

have a blocking effect on the adsorption sites at Pt, at the potential of


17 mV (Figure 2a) compared with Mo-containing materials obtained
by the brh method (Figure 2b).
Furthermore, Figure 3 clearly shows the influence of the preparation method and synthesis sequence on Pt/C, PtMo/C and MoPt/C,
in the three different regions of the characteristic potential described
above. Regarding these regions, the most important issue to highlight
is the current density behavior in the double layer region (0.1 V0.45
V). The Pt/Cthr material exhibits a constant current density which indicates that the accumulated charge is independent of the potential,
whereas Pt/Cbhr has lower current density and increases linearly with
the potential (see inset in Figure 3a). The values of total capacitance,

Ctot can be estimated from the equation Ctot = Jcv /v, where Jcv is the
current density as a function of potential, and v is the scan rate. The
Ctot evaluated at limit values of the potential interval, between 0.10
and 0.4 V, were 180 and 232 F cm2 for Pt/Cthr and 16 and 165 for
Pt/C bhr . These values are in the same order of magnitude as those estimated for a polycrystalline Pt electrode in a similar aqueous medium
of 0.50 M H2 SO4 ,28 which indicates that the Pt/Cthr material favors
the rapid charging of the electrical double layer, probably due to more
copious formation of metallic platinum surface in catalyst particles.
This assertion is consistent with a higher value of SESA which is 9
times greater than that of the material obtained by the bhr method
(Table II). The addition of Mo to the catalyst labeled Pt/Mo/Cbhr
(Figure 3b) results in a current density behavior with a potential similar to that of Pt/Cbhr , however the current density increases one order
of magnitude. This increase suggests that Mo causes the Pt particles to
disperse on the catalyst, as well as the Mo oxide phases on the surface
to be porous and stable (e.g., MoO3 and (MoOx )(OH)ads ). However,
concerning the catalyst obtained by the same synthesis sequence, but
via thermolysis (PtMo/Cthr ), the current densities slightly increase
and exhibit two small oxidation peaks at potentials of 170 and 380 vs.
SCE. These peaks correspond to Mo dissolution, possibly from PtMo
and Pt2 Mo3 surface phases, which have been found in Pt/Mo catalysts
synthesized by calcination and in Pt/Mo obtained by electrochemical
reduction on Pt (111) as well as in Pt/Mo electrodes sputter-deposited
on Au.29,30 The other small anodic peak appearing at 380 mV vs. SCE
(Figure 3b) is difficult to oxidize and it could be a Pt2 Mo3 phase or the
oxidation of MoO3 oxide present on the surface; however, this latter
possibility has not been corroborated or reported in the literature, in
studies on the oxidation of Mo-modified Pt catalysts. On reversing
the synthesis sequence by adding Mo on Pt/C, MoPt/C (Figure 3c),
higher capacitive currents are observed in both materials, compared
with those in Figure 3b. The highest oxidation peak current density at
0.170 mV is observed for the Mo/Pt/Cthr material, because the surface
could be rich in Mo.
In short, the differences observed in the electrochemical behavior
of the catalyst-sulfuric acid medium interface in different potential
regions of the hydrogen adsorption zone and the double layer region,
is indicative of an adsorption site distribution on three Pt polycrystalline surfaces, for materials obtained by bhr. Besides, it is shown
that stable molybdenum phases (MoO3 and (MoOx )(OH)ads ) on the
catalysts surfaces, in 0.5 M sulfuric acid, are present in Pt-Mo/Cbhr ,
Pt+Mo/Cbhr , PtMo/Cbrh and Pt+Mo/Cthr materials. Next is shown the
electrochemical behavior for methanol oxidation.
Figure 4 presents the methanol oxidation voltammograms for the
catalysts obtained by both methods. In the scan toward positive potentials, the anodic current density under the label D is due to the initial
adsorption and oxidation of the alcohol, with maxima at potentials
between 0.63 and 0.66 V vs. SCE. The scan continues up to 1.1 V,
and the anodic current increases again from 0.85 V, as the Pt particles
are oxidized. When the scan direction is reversed toward negative potentials, the anodic peak E is observed. This current density is due to
the oxidation of intermediates adsorbed on Pt, assisted by oxygenated
species from the surface of Pt itself, or from molybdenum phases
present on the catalyst surface (the complete oxidation of methanol to
CO2 requires the addition of an oxygen atom to the original alcohol
molecule). Considerably higher electrical currents for the catalysts
obtained by bhr method were observed, but lower potentials of peak
E were found for the materials synthesized by thr. The highest currents were obtained for the Pt/Mo/C bhr catalyst, despite the fact that
it is not the material with the highest effective Pt surface area (see
Table II), thus indicating greater selectivity of active sites on the Pt
surface.
Electrochemical performance of Pt/Mo/Cbrh electrode.Table III
shows the charge-transfer contribution involved in the methanol oxidation reaction for each material. The charge was calculated by incorporating the area under the curve of each voltammogram, obtained
in H2 SO4 (qacid ) and MeOH (qMeOH ) electrolytic media, in the potential interval from 400 mV up to 900 mVvs. SCE where MeOH
adsorption-oxidation processes take place in this type of catalyst.31

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Current density, A/m2

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)


3
2
1
0
10
8
6
4
2
0
8
6
4
2
0
8
6
4
2
0
8
6
4
2
0
-2

PtMo/Cthr

D'

E'

Pt+Mo/Cthr

Pt-Mo/Cthr

Pt/Cthr

MoPt/Cthr

(a)
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volt vs SCE, V
80

H137

E'

PtMo/Cbhr

D'

Figure 4. Methanol oxidation voltammetry in 1 M


methanol, 0.5 M H2SO4, at 50 mV/s. a) carbonyl thermolysis method. b) Borohydride reduction method.

40

Current density, A/m2

0
10

Pt+Mo/Cbhr

0
10

Pt-Mo/Cbhr

0
10

Pt/Cbhr

0
40
20

MoPt/Cbhr

(b)
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volt vs SCE, V

The charge (qnet ) is the difference between charges (qMeOH qacid )


and indicates the charge-transfer contribution (faradaic charge) of the
methanol oxidation reaction.

Table III. Charge-transfer contribution of methanol oxidation


reaction for each material.
Material
Pt/Mo/Cthr
Pt+Mo/Cthr
Pt-Mo/Cthr
Pt/Cthr
Mo/Pt/Cthr
Pt/Mo/Cbhr
Pt+Mo/Cbhr
Pt-Mo/Cbhr
Pt/Cbhr
Mo/Pt/Cbhr
charge

qacid (mC) qMeOH (mC) qnet (mC) Q MeOH (mC/m2)


0.66
0.4
1.13
1.83
1.13
0.58
0.23
0.21
0.12
0.71

1.23
2.73
6.79
6.01
6.02
5.18
1.95
0.27
0.98
2.02

0.57
2.33
5.66
4.18
4.89
4.59
1.73
0.06
0.86
1.31

0.04
0.45
0.37
0.33
0.32
4.1
0.67
0.06
0.45
1.16

density normalized with the active area of each material.

As can be observed in Table III, the charge-transfer contribution varies a lot among materials, although those synthesized by
the thr method show greater constancy, because they produce higher
charges. However, if these results are expressed as charge density
(Q = [mC/m2 ]), taking into account the effective area of each catalyst, the Pt/Mo/Cbhr material, synthesized by the chemical reduction
method is one that exhibits a contribution of 4.1 mC/m2 , thus being
superior to all other materials.
In order to show the relationship between the electrochemical performance of methanol oxidation and mean particle size, Figure 5
compares the ratios between direct scan peak currents of methanol
oxidation, D (Jf ) and the maximum oxidation peak current density of
the intermediates, E , from the reverse scan (Jb ) as a function of the
particle size. In addition, the ratios between differences in potentials
of the main direct scan oxidation processes (B - D ) and differences
in potentials of reverse scan processes (B - E ) in the potential interval from 400 mV to 900 mV vs. SCE are also compared. The ratio
between potential differences is an indicator of the mean energy required to carry out different oxidation and reduction processes in the
electrochemical system. The ratio (Jf /Jb ) is a measure of comparison
between the rates of two oxidation processes. Figure 5 exhibits the

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H138
2.0
1.6

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

Jf / Jb

thr
bhr

Pt/Mo/C

Pt-Mo/C
Pt+Mo/C

1.2
Pt+Mo/C

Mo/Pt/C

Pt-Mo/C

0.8
4

Mo/Pt/C

Pt/C

(a)

Pt/Mo/C

Pt/C
2

MeOH oxidation charge, QMeOH, (mC/cm )

Figure 5. The effect of particle size of synthesized materials on: a) ratio between Jf/Jb, b) methanol oxidation
charge and c) ratio of differences in the potential of oxidation and reduction processes of CVs from Figs. 2
and 4.

2
1
0
5
4
3
2
1
0

(b)
E oxi (PtOx-MeOH) / E red (PtOx-MeOH)

(c)

2.5

3.0

3.5

4.0

4.5

5.0

Particles size XRD (nm)

if the potential in the first scan is high, the formation of oxides at Pt


and Mo phases is faster and attains a higher conversion than in short
forward scans. Thus, in the reverse scan, peak E appears at lower
potentials since these oxides facilitate the oxidation of intermediates.
The reacting system may follow steps (1) and (2), below, and high Mo
concentrations have an effect on the catalyst activity.34,35
Pt(CO)ads + Mo(OH)ads CO2 + Pt + Mo + H+ + e

[1]

Pt(CO)ads + (MoOx ) (OH)ads CO2 + Pt + MoOx + H+ + e

[2]

Under e conditions with short forward scans, peak E shifts to more


positive potentials because with a lower oxide concentration at the
surface, the oxidation of intermediates requires a higher energy. The
highest current for peak E is obtained when the forward scan goes
up to 1000 mV, and the Jf /Jb ratio decreases to a value of 0.80; this
potential, e , provides the best synergy to perform reactions (1) and
(2). We propose this can be used as a criterion to measure the effective
activity: we observe the condition at which peak E yields the highest
current, and the material with which this peak is located at the lowest

10

Pt/Mo/Cbhr

Current density, A/m2

direct correlation between Jf /Jb , and the ratio between differences in


potentials of the scans of oxidation and reduction processes, as a function of particle size. However, the correlation of the above parameters
with the charge-transfer contribution of MeOH oxidation (QMeOH ) is
inverted with respect to the particle size. For the materials obtained by
the bhr method, the correlations are greater than for those obtained by
thr. The curves of the materials obtained via bhr clearly show that an
average particle size of 3.65 nm results in a minimum ratio value of the
energy required for oxidation and reduction processes, a maximum
oxidation charge of MeOH (4.1 mC/m2 ) and a minimum Jf /Jb (0.56).
In bimetallic and trimetallic Pt catalysts modified with Ru and Mo,
the Jf /Jb ratio decreases with the number of activation cycles and with
the decrease in the Ru atomic ratio in the catalyst; besides, low values
of Jf /Jb (<1) indicate low CO tolerance of the catalyst.32,33 However,
in Pt catalysts modified with Mo by brh, the material obtained using
the synthesis sequence PtMo/Cbhr , turns out to be better for MeOH
oxidation and the only one to have Jf /Jb <1 (0.85). According to the
above results, there is no clear explanation to the effect of Mo on CO
tolerance of the catalyst, especially since this system is so complex
and includes many variables: medium concentration, catalyst parameters, structure and size of the particles, activation and temperature, as
well as the scan rate used to obtain the CV. Regarding CO tolerance,
our group continues performing CO striping and catalyst optimization
studies.
In order to understand the trend between the relationship of oxidation peak currents of methanol, D (Jf ), and the maximum current
density of the oxidation peak of intermediates, E, (Jb ), a study on
inversion potentials, e in PtMo/Cbhr nanocatalyst was performed and
established the best performance. It is important to mention that the
voltammetric study of alcohol oxidation in PtMo/Cbrh was performed
by applying 35 voltammetric cycles and, in cycle 7, the steady state
of alcohol oxidation was achieved (results not shown).
Figure 6 shows the voltammetry curves of one of the catalysts
(Pt/Mo/Cbhr ) from the experiments with variation in the positive inversion potential, e , in order to obtain information on peak E, which
is associated with the oxidation of adsorbed intermediates. The potential scans were consecutively carried out up to 800, 900, 1000 and
1100 mV. It can be seen that peak D remains unchanged in all sweeps,
but peak E presents changes in the current and potential. When the
forward scan is performed up to 1100 mV vs. SCE, peak E in the
reverse scan is located at a low potential. As the forward scan is shortened, peak E shifts to more positive potentials. This happens because

0
-5
-10
e = 0.8 V

10

E'

e = 0.9 V

D'
e

e =1V
e = 1.1V

0
-0.2

0.0

0.2

0.6

0.8

1.2

Figure 6. Methanol oxidation voltammetry on Pt/Mo/C bhr, in 1 M methanol,


0.5 M H2SO4, at 50 mV/s, with varying inversion potential, e.

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)

H139

Figure 7. (a) Annular dark field TEM image showing bright dots well dispersed on the C support of gray contrast and b) HR-TEM image showing the
crystalline nature of nanoparticles. Inset FT. These micrographs correspond to
the Pt/Mo/Cbhr sample.

potential is the catalyst that requires the lowest activation energy


to oxidize adsorbed intermediates. Following this criterion, the best
catalyst among both series is Pt/Mo/C bhr .
In order to understand the high electrochemical performance of
the Pt/Mo/Cbhr catalyst, a detailed study through TEM technique was
conducted on this material. Bright dots on a gray contrast are observed
in the annular dark field TEM image of Figure 7a. According to the
Rutherford dispersion theory, the bright dots were generated by heavy
elements such as Pt and Mo and the gray contrast was generated by
a light element such as carbon. All metallic particles were well dispersed on the carbon support. The nanometric-scale particle size was
in the range of 1 to 10 nm. High resolution-transmission electron microscopy (HR-TEM) performed on the metallic nanoparticles revealed
their crystalline nature, see Figure 7b. Fourier transform (FT) crystallographic analysis evidenced the d-spacing of (111), (200) and (220)
crystalline planes, corresponding to the face-centered-cubic platinum
phase. A d-spacing corresponding to molybdenum oxide structures
was not observed. This could be because of the low Mo concentration in the sample. However, to elucidate the metallic Mo by FT is
difficult because both metals have the same space group, according
to the JCPDS card number 882331 for metallic Mo, and their lattice
parameters are very close (a = 0.3923 nm and Pt a=0.403 nm Mo).
Therefore, the molybdenum atoms detected by SEM-XEDS could be
present in the Pt structure as solid solution.
On the other hand, HR-TEM images of single metallic nanoparticles presented some very well defined geometric shapes and others
unshaped. This is illustrated in Figure 8. Triangular, square and hemispherical shapes are observed in Figure 8a8c and an unshaped particle
is displayed in Figure 8d. FT obtained from the lattice displayed in
the single nanoparticle is shown in the inset on each image. All white
dots observed in the FT had a d-spacing corresponding to the (111)

Figure 8. HRTEM of single nanoparticles showing the different shapes of


growth: a) Triangular, b) Square, (c) Hemispherical and triangular, and
(d) unshaped. Inset FT.

and (200) planes of the platinum structure. This result shows no direct
evidence of the Mo atoms. However, their presence was evidenced
by TEM-EDXS. The results obtained are displayed in the EDX spectra of Figure 9. The elemental chemical analysis performed on the
single nanoparticles evidenced the presence of Mo, Pt, Si, O and C
atoms while that performed on the Vulcan carbon support detected
Mo, Si, O and C atoms. The Cu peak corresponds to the copper grid.
The chemical analysis was performed on the catalyst sample that was
present on a holey of the formvar film. Therefore, the carbon detected
corresponds to the catalyst sample and not to the formvar film of the
Cu grid. As can be seen in the EDXS results, the single crystalline
nanoparticles were formed by Pt atoms. The presence of a Mo peak in
the analysis suggests that it is located on the Vulcan carbon support, as
was detected in the chemical analysis performed only on the Vulcan
carbon support. Therefore, the Mo atoms could be very well dispersed
on the carbon support at atomic level which is difficult to appreciate
in the HR-TEM images. Direct evidence of such Mo atom dispersion
on the catalytic support and possibly on the nanoparticles at very low
concentrations will be further investigated in detail in future.

Figure 9. XEDS spectra obtained from a (a) single nanoparticle and (b)Vulcan carbon support.

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H140

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)


shown), whereas the material containing Pt/C exhibits direct catalytic
hydrogenation activity by fast poisoning.40 The presence of MoOx
in the Pt/Mo/Cbhr formulation contributes to MeOH oxidation with
additional protons in the spillover mechanism.
Conclusions
Pt and Mo bimetallic electrocatalysts containing around 10.5 wt%
Pt and 2.4 wt% Mo on a Vulcan carbon support were obtained by
the thr and bhr methods. Pt crystalline particles in nanometric sizes
between 1 and 10 nm were very well dispersed on the Vulcan carbon
support where Mo was dispersed on an atomic level , too.
Redox couples due to changes in the oxidation state of Mo can be
seen in the electrocatlysts obtained by the thr method, but not in the bhr
reduction method. Higher electrochemical activity area was obtained
in bimetallic electrocatalysts than in monometallic Pt electrocatalyst.
The thr method yields smaller particles and larger active areas
than the bhr method. However, it is by the bhr reduction method that
the highest methanol oxidation charges were obtained. Therefore, the
Pt surface area is not the main factor to consider in explaining the
high anodic current obtained, or the catalytic mechanism of methanol
oxidation.
An aggregation sequence, in which Pt was deposited on the
monometallic Mo/C by borohydryde reduction, leads to the highest overall activity observed for the Pt/Mo/C bhr material because the
MoOx and Mox Cy in the vicinity of Pt participate in the spillover
mechanism.
Acknowledgments

Figure 10. Schematic representation of H2O activation over the catalytic systems without Mo (a) and in Mo-modified Pt nanocatalyst with low (b) and
optimal (c) Pt/MoOx/MoxCy ratios.

The authors thank Cecilia Saucedo for having obtained XRD patterns and I. Puente-Lee for XEDS measurements. Ugalde acknowledges the grant provided by Conacyt Mexico.
References

The effect of molybdenum on MeOH oxidation.It is generally suggested that Mo is able to assist CO electro-oxidation because molybdenum brings in oxygenated species needed for the removal of
CO adsorbed on platinum through the formation of Mo(OH)ads or
(MoOx )(OH)ads . Consequently, Mo-modified Pt catalyst ( in ex. PtRu-Mo) presents best catalytic activities for alcohol electro-oxidation,
which may be attributed to the electronic effect, bifunctional mechanism, and hydrogen spillover effect.3638
The effect of molybdenum is nowadays an issue open to discussion, but there are gas-phase studies applied to catalytic hydrogenation of toluene, such as that carried out in the gas phase by Zosimova
et al.36 Based on XRD, SEM, TEM and XEDS analyses performed
in this study, a spillover mechanism is proposed in which the MoOx
acts as a stabilizer of the proton that is previously dissociated on the
platinum surface. Our electrochemical results show that the best results toward MeOH oxidation correspond to the Pt/Mo/Cbhr catalyst,
which presents a slight slope in the electrical double layer, indicative
of the presence of Mo oxide phase at the surface; this presence is
corroborated by the increase in the reduction current on the reverse
scan. Likewise, HR-TEM results show the presence of metallic Pt
particles resulting in a mixture of Mo oxide and Pt. On the other hand,
due to the chosen aggregation sequence, a part of the Mo loading
may be incorporated into the support, thus forming an amorphous
Mox Cy phase. This is evidenced by the presence of Mo traces shown
by XEDS analysis in Figure 9b. Therefore, we propose a proton stabilization mechanism based on spillover model, which suggests that
Pt hydrolyzes the water molecule37 and the resulting proton is stabilized on the surface of MoOx and MoC.39 As Zosimova et al., we
suggest that the MoOx particles in the vicinity of Pt are the only
ones to participate in the spillover mechanism, as shown in Figure 10
. This mechanism is evidenced by the lack of electrochemical activity for MeOH oxidation in the Mo oxide on C sample (Mo/C, not

1. C. N. Hamelinck and A. P. C. Faaij, J. Power Sources, 111, 1, (2002).


2. E. Antolini, F. Colmati, and E. R. Gonzalez, Electrochem. Comm. 9, 398 (2007).
3. M. V. Martnez-Huerta, J. L. Rodrguez, N. Tsiouvaras, M. A. Pena, J. L. G. Fierro,
and E. Pastor, Chemistry of Materials 20, 4249 (2008).
4. M. Watanabe and S. Motoo, Journal of Electrolytical Chemistry 60, 267 (1975).
5. J. L. Gomez de la Fuente, M. V. Martnez-Huerta, S. Rojas, P. Hernandez-Fernandez,
P. Terreros, J. L. G. Fierro, and M. A. Pena, Applied Catalysis B: Environmental 88,
505 (2009).
6. J. G. Kim, J. Z. Zhyu, and J. R. Regalbuto, Journal of Catalysis, 139, 153 (1993).
7. A. C. C. Tseung and K. Y. Chen, Catalysis Today 38, 439 (1997).
8. Z. Jiang, W. Huang, H. Zhao, and Z. Zhang, Journal of Molecular Catalysis A:
Chemical 268, 213 (2007).
9. E. Antolini, J. R. C. Salgado, and E. R. Gonzalez, Applied Catalysis B: Environmental
63, 137 (2006).
10. A. J. Dickinson, L. P. L. Carrette, J. A. Collins, K. A. Friedrich, and U. Stimming,
Electrochimica Acta, 47, 3733 (2002).
11. W. H. Lizcano-Balbuena and E. R. Gonzalez, Electrochimica Acta 49, 1289 (2004).
12. F. Colmati, E. Antolini, and E. R. Gonzalez, Applied Catalysis B: Environmental 73,
106 (2007).
13. M. S. Hyun, S.-K. Kim, B. Lee, D. Peck, Y. Shul, and D. Jung, Catalysis Today 132,
138 (2008).
14. P. Ochal, J. L. Gomez de la Fuente, M. Tsypkin, and F. Seland, Journal of Electroanalytical Chemistry 655, 140 (2011).
15. J. E. Hu, Z. Liu, B. W. Eichhorn, and G. S. Jackson, International Journal of Hydrogen Energy 37, 11268 (2012).
16. O. Guillen-Villafuerte, R. Guil-Lopez, E. Nieto, G. Garca, J. L. Rodrguez, E. Pastor,
and J. L. G. Fierro, Int. J. Hydrogen Energy 37, 7171 (2012).
17. G. Longoni and P. Chini, J. of the American Chemical Society 96, 23 1976.
18. L. C. Ordon ez, P. Roquero, P. J. Sebastian, and J. Ramrez, Int. J. Hydrogen Energy
32, 3147 (2007).
19. O. Ugalde Reyes, P. Roquero, R. Hernandez Maya, A. L. Ocampo Flores, and
E. Sosa Hernandez, ECS Transactions 36(1) 21 (2011).
20. L. C. Ordon ez, P. Roquero, P. J. Sebastian, and J. Ramrez, Catalysis Today 46, 107
(2005).
21. F. Maillard, G.-Q. Lu, A. Wieckowski, and U. Stimming, J. Phys. Chem. B 109,
16230 (2005).
22. H. Liu, C. Song, L. Zhang, J. Zhang, H. Wang, and D. P. Wilkinson, J. Power Sources
155, 95 (2006).
23. P. Piela, C. Eickes, E. Brosha, F. Garzon, and P. Zelenay. Journal of The Electrochemical Society, 151(12) A2053 (2004).

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (3) H132-H141 (2015)


24. M. Umeda, M. Kokubo, M. Mohamedi, and I. Uchida, Electrochimica Acta 48, 1367
(2003).
25. D. M. Dos Anjos, K. B. Kokoh, J. M. Leger, A. R. De Andrade, P. Olivi, and
G. Tremiliosi-Filho, Journal of Applied Electrochemistry, 36, 1391 (2006).
26. M. Watanabe, K. Makita, H. Usami, and S. Motoo, J. Electroanal. Chem. 197, 195
(1986).
27. A. C. R. Aguiar and P. Olivi. Journal of Power Sources 195, 3485 (2010).
28. Jin-Bum Park and Su-Moon Park, J. Electroanal. Chem. 656, 243 (2011).
29. M. Topic, Z. Khumalo, and C. A. Pineda-Vargas, Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms
318 (Part A), 163 (2014).
30. G. Samjeske, H. Wang, T. Loffler, and H. Baltruschat, Electrochimica Acta 47, 3681
(2002).
31. Zeng Jianhuang, Shu Ting, Liao Shijun, and Liang Zhenxing, Chinese Journal of
Catalysis 32, 86 (2011).
32. Chien-Te Hsieh, Jia-Yi Lin, and Shu-Ying Yang, Physica E 41, 373 (2009).

H141

33. Sung-Hyeon Park, Sung-Jun Joo, and Hak-Sung Kima, Journal of The Electrochem.
Soc. 161, F405 (2014).
34. A. Lima, C. Coutanceau, J. M. Leger, and C. Lamy, J. Appl. Electrochem. 31, 379
(2001).
35. D. A. Stevens, J. M. Rouleau, R. E. Mar, R. T. Atanasoski, A. K. Schmoeckel,
M. K. Debe, and J. R. Dahn, J. Electrochem. Soc. 154, B1211 (2007).
36. Y. W. Song, S. H. Park, W. S. Han, J. M. Hong, and H. S. Kim, Mater. Lett. 65, 2510
(2011).
37. S. H. Park, H. M. Jung, S. S. Um, Y. W. Song, and H. S. Kim, Inter. J. Hydrogen
Energy 37, 12597 (2012).
38. N. Tsiouvaras, M. V. Martnez-Huerta, O. Paschos, U. Stimming, J. L. G. Fierro, and
M. A. Pena, Int. J. Hydrogen Energ. 35, 11478 (2010).
39. P. A. Zosimova, A. V. Smirnov, S. N. Nesterenko, V. V. Yuschenko, W. Sinkler,
J. Kocal, J. Holmgren, and I. I. Ivanova, J. Phys. Chem. C, 111, 14790 (2007).
40. P. A. Attwood, B. D. Mcnicol, and R. T. Short, J. of Appl. Electrochem. 10, 213
(1980).

Downloaded on 2015-01-10 to IP 187.233.26.155 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Você também pode gostar