You are on page 1of 9

Electrochimica Acta 52 (2006) 385–393

Pt-Co/C nanoparticles as electrocatalysts for oxygen reduction in


H2SO4 and H2SO4/CH3OH electrolytes
F.H.B. Lima, W.H. Lizcano-Valbuena 1 , E. Teixeira-Neto 2 , F.C. Nart,
E.R. Gonzalez, E.A. Ticianelli ∗
Instituto de Quı́mica de São Carlos, Universidade de São Paulo, CEP 13560-970, CP 780 São Carlos-SP, Brazil
Received 7 April 2006; received in revised form 21 May 2006; accepted 22 May 2006
Available online 30 June 2006

Abstract
The oxygen reduction reaction (ORR) was studied on carbon dispersed Pt and Pt-Co alloyed nanocatalysts with high contents of Co in H2 SO4
and H2 SO4 /CH3 OH solutions. The characterization techniques considered were transmission electron microscopy (TEM), X-ray diffraction (XRD)
and in situ X-ray absorption near edge structure (XANES). The electrochemical activity for the ORR was evaluated from steady state polarization
measurements, which were carried out in an ultra thin layer rotating disk electrode. The results showed that with the increase of Co content, the
nanoparticle size distributions become sharper and the mean particle diameters become smaller. XRD indicated low degree of alloy formation but
significant phase segregation of Co was observed only for Pt-Co/C 1:3 and 1:5 (Pt:Co atomic ratios). The electrochemical measurements indicated
that the four-electrons mechanism is mainly followed for the ORR on all materials and the electrocatalytic activities per gram of Pt is higher for
the catalysts with higher Co contents. This was explained based on the XANES results which evidenced a decrease of the coverage of oxygenated
Pt adsorbates due to the presence of Co. In the methanol-containing electrolyte, the Pt-Co/C 1:5 catalyst showed the highest performance. This
was attributed to its low activity for the methanol oxidation due to the smaller probability for presenting three Pt neighboring Pt active sites.
© 2006 Elsevier Ltd. All rights reserved.

Keywords: Oxygen reduction; Methanol crossover; Platinum alloys; Direct methanol fuel cell; Particle morphology

1. Introduction tion strength of oxygenated species. X-ray absorption near edge


structure (XANES) measurements for Pt-V/C, Pt-Cr/C and Pt-
Several studies have been carried out in the last years aiming Co/C alloys indicated lower Pt 5d band vacancy in higher
to increase the platinum electrocatalytic activity for the oxygen potentials due to the presence of the non-noble metal. This was
reduction reaction (ORR) [1–3]. Several authors have reported attributed to the strong electronic interaction between Pt and
that on Pt alloys, as for example Pt-V, Pt-Cr, Pt-Co, Pt-Fe, Pt-Ni, the non-noble metal, reducing the Pt reactivity for adsorbates,
etc., there is an increase in the kinetics of the ORR compared to and/or due to the presence of M-OH (M being the non-noble
pure Pt, and this fact has been attributed to changes in the Pt–Pt metal) neighbor species suppressing the Pt-OH formation, as
bond distance, number of Pt nearest neighbors, electron density also reported by other authors [1,2]. Steady-state polarization
of states in the Pt 5d band, and nature and coverage of surface curves have indicated a higher catalytic activity for the Pt-
oxide layers [3–10]. V/C material. The activity enhancement was attributed to faster
In our previous work [3], it was proposed that the oxygen Pt-O− electroreduction due to the lowering of the adsorption
reduction activity increases with the decrease in the adsorp- strength of adsorbed oxygen species caused by a stronger elec-
tronic interaction and by the presence of some surface vanadium
hydroxide or oxide.
∗ Corresponding author. Tel.: +55 16 3373 9945; fax: +55 16 3373 9952. Nørskov and co-workers [11–14] have shown that the cat-
E-mail address: edsont@iqsc.usp.br (E.A. Ticianelli). alytic reactivity of the metals can be rationalized in terms of the
1 Present address: Departamento de Quı́mica, Universidad del Valle, Calle 13

#100-00, Ciudad Universitaria “Melendez”, Cali, Colombia.


energy of the d-band center with respect to the Fermi level. The
2 Present address: Instituto de Quı́mica, Universidade Estadual de Campinas, Nørskov model is based on shifts of the d-band center, increas-
Unicamp, 13081-970 Campinas, SP, Brazil. ing or decreasing the reactivity of the metal catalysts. As the

0013-4686/$ – see front matter © 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2006.05.019
386 F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393

d-band center shifts up, a distinctive anti-bonding state appears the binary electrocatalysts for methanol oxidation, arising from
above the Fermi level. Since the anti-bonding states are above the a composition effect.
Fermi level, they are empty, and the bond becomes increasingly This paper aims to evaluate the electrocatalytic activity for
stronger as the number of empty anti-bonding states increases. It the ORR in H2 SO4 and H2 SO4 /CH3 OH solutions on ultra thin
has been also shown [13,14] that a metal monolayer on different layer electrodes, formed by high Co content Pt-Co alloy nanopar-
metal substrates may be subjected to a tensile or compressive ticles dispersed on carbon (Pt-Co/C) prepared with several Pt:Co
force, which is determined by the substrate lattice parameter. atomic ratios. Transmission electron microscopy (TEM), X-
This can cause a variation on the d-band broadness, leading to ray diffraction (XRD) and in situ X-ray absorption near edge
a shift of the d-band center to preserve the degree of the band structure analyses were carried out on the Pt-Co/C alloys to
filling. The shift of the d-band center has been also calculated for determine the particle diameter distribution and the structural
alloys with two or more metals, or for surface segregated phases and electronic properties of the catalysts. The electrochemical
of a given metal on a metal nanoparticle. For example, in the case techniques used were cyclic voltammetry and steady state polar-
of the alloying of Pt with an excess of Co, for which the model ization measurements, which were carried out with the standard
predicts a surface segregation of Pt [11], there is a down-shift rotating disk electrode technique. Mass-transport corrected Tafel
of the Pt 5 d-band center. This is mainly caused by the lattice plots were used for analyzing the ORR kinetics in methanol
mismatch and the strong electronic interaction between the Pt free electrolytes. TEM, XDR, XANES and the electrochemical
and Co atoms [12]. For the oxygen reduction [12,15,16], it has results were compared in order to correlate the particle morphol-
been shown that an up-shift of the d-band center, resulting in ogy with the kinetics of the ORR and the methanol tolerance of
empty anti-bonding states, leads to stronger Pt-O− adsorption. the electrocatalysts.
The opposite conducts to the formation of occupied anti-bonding
orbitals and, as a consequence, weaker Pt-O− adsorption. Thus, 2. Experimental
higher activity of the Pt alloys compared to that of pure Pt is
expected when the Pt 5 d band center shifts down, causing a The investigated Pt-Co/C electrocatalysts with different
decrease of the adsorption strength of oxygenated species, and Pt:Co atomic ratios (1:0, 3:1, 1:1, 1:3 and 1:5) were prepared
leading to faster electroreduction kinetics of the reaction inter- by chemical reduction of the precursors (H2 PtCl6 , Aldrich and
mediates. Co(NO3 )2 , Merck) using NaBH4 , followed by deposition onto
Direct methanol fuel cells (DMFC) are good candidates for a high surface area carbon powder (Vulcan XC-72R) [23]. The
application in transportation and mainly in portable electron- reduction process was carried out at 60 ◦ C by adding drop by
ics devices as power sources, and so, the ORR in the pres- drop the solution with the precursor and sodium borohydrid
ence of methanol has been intensively studied [17–20]. Besides (0.5 mol L−1 ) into a carbon powder (Vulcan XC-72) aqueous
the slow kinetics in both, anode and cathode, the effects of slurry, prepared by suspending the carbon powder in ultra pure
crossover of methanol from the anode to the cathode must be water (Millipore) and ultrasonically blending for 20 min. After
minimized, either by using electrolytes with lower methanol filtering, the composite powders were submitted to a thermal
permeability or by developing more efficient methanol tol- treatment conducted in a tubular oven (MAITEC) under H2
erant cathode catalysts. For the latter, two concepts can be atmosphere at 500 ◦ C for 1 h.
identified. First, the cathode catalysts in the DMFC should The working electrodes were composed of the metal/C cat-
show a high methanol tolerance, which means that the oxy- alysts deposited as a thin layer on a pyrolitic graphite disk,
gen reduction reaction will not be affected by the adsorption 5 mm diameter (0.196 cm2 ), of a rotating disk electrode. The
and oxidation of methanol. Secondly, in order to shift the ultra thin layers were prepared starting from an aqueous suspen-
mixed potential to more positive values, the catalyst should sion of 2.0 mg mL−1 of the metal/C produced by ultrasonically
show a higher exchange current density for oxygen reduc- dispersing the powder in pure water (Millipore) [24]. A 20 ␮L
tion. High tolerance is reported in the literature for non-noble aliquot of the dispersed suspension was pipetted onto the top of
metal catalysts based on chalcogenides [21] or metallopor- the pyrolitic graphite substrate surface. After the evaporation of
phyrins [22]. These catalysts have presented nearly the same water, in a low vacuum condition, 20 ␮L of a diluted Nafion solu-
current/potential behavior in the absence and in the presence tion (from 5%, Aldrich) were pippeted onto the electrode surface
of methanol. However, in methanol free solutions these materi- catalyst in order to attach the catalytic particles on the pyrolitic
als did not attain the activity of Pt for the ORR. Furthermore, graphite RDE electrode substrate and, after that, dried under vac-
the long time stability under fuel cell condition has still to be uum. Right after preparation, the electrodes were immersed into
improved. the desaerated electrolytes. A large area platinum screen served
Carbon supported Pt-Co/C alloys with low Co contents have as counter electrode and a reversible hydrogen electrode (RHE)
been tested as methanol-tolerant oxygen reduction catalysts in 0.5 mol L−1 H2 SO4 was used as reference. All the experi-
[18,20]. It has been shown [18] that Pt-Co/C, 4:1 and 3:1(atomic ments were carried out in 0.5 mol L−1 H2 SO4 or 0.5 mol L−1
ratios) alloys possess enhanced oxygen reduction activity com- H2 SO4 /0.1 mol L−1 CH3 OH electrolytes, prepared from high
pared to Pt/C in the presence of methanol both in sulphuric acid purity reagents (Merck) and water purified in a Milli-Q (Milli-
electrolyte in a half-electrochemical cell configuration and in a pore) system. Before the polarization experiments, the electrode
direct methanol fuel cell (DMFC). This high methanol tolerance potential was cycled several times between 0.05 and 0.9 V ver-
of the Pt-Co/C electrocatalysts was ascribed to the low activity of sus RHE in order to produce clean surfaces, using an Autolab
F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393 387

potentiostat (PGSTAT30). Steady state polarization measure- The computer program used for the analysis of the XAS
ments were recorded point-by-point in the potentiostatic mode, data was the WinXAS package [30]. The data analysis was
at several rotation rates. The experiments were conducted under done according to procedures described in detail in the liter-
controlled temperatures of 30 and 60 ± 0.1 ◦ C by using a Hakee- ature [31,32]. Briefly, the X-ray absorption near edge structure
K20 thermostat. spectra were first corrected for the background absorption by
The catalyst composition on different powder samples was fitting the pre-edge data (from −60 to −20 eV below the edge)
determined by energy dispersive X-ray analysis using an EDX- to a linear formula, followed by extrapolation and subtraction
LEO model 440. Morphological information on the catalysts from the data over the energy range of interest. Next, the spectra
was obtained by TEM using a JEM-3010 ARP microscope. The were calibrated for the edge position using the second deriva-
samples were prepared by ultrasonically treating the catalyst tive of the inflection points at the edge jump of the data from the
powders in isopropanol. A drop of the resulting dispersion was reference channel. Finally, the spectra were normalized, taking
placed on thin carbon films deposited on standard TEM copper as reference the inflection points of one of the extended X-ray
grids and dried in air. The images were acquired [25] by observ- absorption fine structure (EXAFS) oscillations.
ing many different areas of the samples, in order to assess its
average characteristics; at least five different areas, randomly 3. Results and discussion
chosen, were imaged under low magnification. The diameters
of the catalyst particles were measured from the TEM images The effect of the catalyst composition on the particle size was
using an Image-Pro Plus 4.0 software. At least 500 nanoparti- investigated by TEM measurements. Fig. 1 shows the obtained
cles of each sample were measured to build the size distribution particle diameter distribution histograms together with repre-
histograms. sentative TEM images of each catalyst. It is seen that the size
Physical properties such as the lattice parameter and the aver- distributions become sharper and the particles smaller as the Co
age particle size were estimated from X-ray diffraction (XRD- content in the catalyst increases. The diameter ranges and peak
RIGAKU model RU200B) measurements carried out in the 2θ relative abundances of each investigated material are shown in
range from 10◦ up to 100◦ and using Cu K␣ radiation (with a Table 1. The Pt/C catalyst has the largest particles (mean diam-
scan rate of 3◦ min−1 ). The lattice parameter (a) was calculated eter 4.8 nm) and the broader size distribution, with 2.9% of the
and refined using a least-square method [26]. The average parti- particles in the range of 15–46 nm. The presence of either large
cle size was estimated using the (1 1 1) peak of the Pt diffraction particles or the aggregation of small particles is evidenced by
pattern by using the Sherrer equation [27]. these TEM results for Pt/C. The measurements on the Pt-Co/C
1:1 catalyst indicated a mean particle diameter of 2.9 nm with
D = kλ/B cosθ (1) only 0.5% of the particles presenting diameters in the range of
15 nm to 19 nm, indicating a good particle spatial dispersion
where D is the average particle size in Å, k a coefficient taken without the formation of large particle aggregates. The Pt-Co/C
here as 0.9 [27], λ the wavelength of the X-rays used (1.5406 Å), 1:3 catalyst has even smaller particles (mean diameter 2.7 nm),
B the width of the diffraction peak at half height in radians, and and the size distribution histogram shows the existence of only
θ is the angle at the position of the peak maximum. 1.4% of the particles within the range of 15 nm to 40 nm. The
In situ XANES measurements were performed in the Pt L3 Pt-Co/C 1:5 catalyst has the sharpest size distribution histogram
absorption edge using a home-built spectroelectrochemical cell and the smallest particles (mean diameter 2.0 nm). Hence, 0.2%
[28]. The working electrodes consisted of pellets formed with of the particles were in the size range of 15 nm to 16 nm, so this
the dispersed catalysts agglutinated with Nafion (ca. 5 wt.%) material shows the smallest tendency for particle aggregation.
and containing 6 mg cm−2 (Pt loading). The counter electrode Although the results are not shown here, the Pt-Co/C 3:1 catalyst
was a Pt screen cut in the center in order to allow the free was studied in a previous work and the particle morphology has
passage of the X-ray beam. Prior to the experiments, the work- shown similar features when compared with Pt/C [18].
ing electrodes were soaked in the electrolyte for at least 48 h. Fig. 2 shows the X-ray diffraction patterns for the Pt-Co/C
XANES experiments were carried out at 0.3 and 0.9 V versus alloys. A broad reflection observed at 2θ = 25◦ is due to the car-
RHE, after cycling the electrodes in the range defined by these bon support. As indicated, all XRD patterns show the five main
potentials. Results presented here correspond to the average of at characteristic peaks of the face-centered cubic (fcc) crystalline
least two independent measurements. All experiments were con- structure of Pt, namely the planes (1 1 1), (2 0 0), (2 2 0), (3 1 1)
ducted at the X-ray absorption spectroscopy (XAS) beam line and (2 2 2). In all cases, the broad reflection peaks observed
in the Brazilian Synchrotron Light Laboratory (LNLS), Brazil. for all catalysts indicate that they are nanostructured materi-
The data acquisition system for XAS comprised three ionization als with small grain sizes. The values of the lattice parameters
detectors (incidence I0 , transmitted It and reference Ir ). The ref- for the Pt-Co/C catalysts are shown in Table 2. The reflections
erence channel was employed primarily for internal calibration of the Pt-Co/C samples are essentially at the same positions
of the edge positions by using a pure foil of the metals. Nitro- as those for Pt/C, indicating a negligible difference of the Pt
gen was used in the I0 , It and Ir chambers. Owing to the low unit cell lattice parameter. In Table 2, it is noted that the lattice
critical energy of the LNLS storage ring (2.08 keV), third-order parameters are only slightly smaller for the Pt-Co/C materials
harmonic contamination of the Si (1 1 1) monochromatic beam compared to Pt/C, and this evidences very low incorporation
is expected to be negligible above 5 eV [29]. of Co atoms into the Pt lattice, in contrast of what is observed
388 F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393

Fig. 1. TEM micrographs and particle size distribution histograms obtained from the micrographs of Pt and Pt-Co nanoparticles supported on Vulcan carbon XC 72
substrate, deposited on an on amorphous carbon film.
F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393 389

Table 1
Morphological information on the catalysts obtained by HRTEM: nominal atomic Co content, mean particle diameter with respective standard deviation, maximum
measured diameter and diameter range of peak relative abundance
Catalyst Nominal atomic Mean diameter Standard deviation Maximum Diameter range (nm) of peak
Co content (%) (nm) (±) (nm) diameter (nm) relative abundance (%)

Pt 0 4.8 4.2 45.6 2.5–3.0; 15


Pt-Co/C 1:1 50 2.9 2.0 18.4 1.5–2.0; 25
Pt-Co/C 1:3 75 2.7 3.0 39.2 1.5–2.0; 30
Pt-Co/C 1:5 83 2.0 1.4 15.2 1.0–1.5; 30

alloy (0.3831 nm), assuming a linear dependence with the Co


atomic fraction. The values of the calculated χCo for the Pt-
Co/C electrocatalysts are also presented in the Table 2. As can
be observed, only a small amount of Co really alloyed with
Pt was obtained for all Pt-Co/C catalysts. This means that the
degree of alloying is low, evidencing a considerable amount of
Co segregated phases. Part of this segregated phase is evidenced
by the diffraction pattern shoulders at 43◦ and 59◦ , while the
rest of the segregated Co atoms must be forming a preponderant
amorphous oxide phase, not detected by XRD. The crystallite
sizes were estimated using the Sherrer equation and the results
are also shown in Table 2. It can be noted a decrease in the crys-
tallite size when the Co content increases, in agreement to the
TEM results.
The electronic properties of the platinum alloys were investi-
gated by X-ray absorption near edge structure. Fig. 3 shows the
results obtained at the Pt L3 edge for the Pt-Co/C 3:1, Pt-Co/C
1:3 and Pt/C electrocatalysts in acid medium at two electrode
Fig. 2. X-ray diffraction (XRD) patterns for the Pt and Pt-Co/C alloys at different
atomic ratios. potentials. The result for a thin Pt foil is included for compari-
son. The absorption at the Pt L3 edge (11.564 eV) corresponds to
for high temperatures heat treated Pt-Co/C materials [33]. The 2p3/2 –5d electronic transitions and the magnitude of the absorp-
presence of some amount of a segregated Co3 O4 phase is evi- tion hump or white line located at ca. 5 eV is directly related
denced by the appearance of the diffraction pattern shoulders at to the occupancy of the 5d electronic states, the higher is the
43◦ and 59◦ for Pt:Co 1:3 and 1:5. For the other materials, the hump the lower is the occupancy and vice-verse. Fig. 3a com-
presence of cobalt oxides cannot be discarded, but if so, they pares the results obtained for the Pt-Co/C 3:1 catalyst with that
may be present in very small amounts or forming amorphous for Pt/C. It seen that the white line magnitude increases with the
phases. increase of the electrode potential from 0.0 to 0.90 V for both
Assuming that the dependence of the lattice parameter on Co catalysts. This phenomenon is attributed to the emptying of the
content is the same for supported and unsupported Pt-Co alloys, Pt 5d band, in agreement with the presence of an electron with-
the Co atomic fraction of carbon supported Pt-Co, χCo , can be drawing effect of the oxygen present in a well-known surface
obtained from Vegard’s law [34]: oxide layer formed above 0.80 V on the catalyst particle surface
[5]. For Pt-Co/C 3:1 a slight lower increase in the white line is
a − a0
χCo = (2) observed, compared to Pt/C. On the other hand, an important
k aspect is observed in Fig. 3b, where it is seen that the increase of
where a0 is the lattice parameter of pure carbon supported Pt, the white line, caused by the oxide formation on Pt, is much less
0.3917 nm, and k = 0.0368 nm is a constant, obtained from the pronounced for the Pt-Co/C 1:3 compared to Pt/C. These results
lattice parameters of unsupported Pt (0.3923 nm) and Pt3 Co indicate smaller Pt-oxide formation for the Pt-Co/C material.

Table 2
Structural characteristics by XRD analysis of carbon-supported Pt and PtCo electrocatalysts prepared with different atomic ratios
Catalyst EDX Pt:Co Lattice parameter (nm) Mean interatomic distance (nm) Particle size (nm, XRD) % of Co in the alloy

Pt/C – 0.3917 0.2769 5.0 –


Pt-Co/C 3:1 76.86:23.14 0.3905 0.2761 3.8 3.2
Pt-Co/C 1:1 51.22:48.78 0.3904 0.2760 3.4 3.7
Pt:Co/C 1:3 23.87:76.13 0.3880 0.2743 2.7 10.0
Pt:Co/C 1:5 20.12: 79.88 0.3895 0.2754 2.6 6.0
390 F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393

Fig. 4. Cyclic voltammograms for the Pt/C and Pt-Co/C alloys in N2 saturated
0.5 mol L−1 H2 SO4 electrolyte at 20 mV s−1 and 30 ◦ C. The currents are nor-
malized by mass of Pt in the catalyst layer.

with the neighboring Pt atoms. The redox pair observed for Pt/C
at ca. 0.5 V refers to some carbon feature, as seen previously in
acid media [36]. The CV profiles for the different Co contents
are in apparent contradiction with the tendency of increasing the
catalyst area observed in Fig. 1 because a decrease of the currents
is observed with the increase of the Co content. In fact, these
results give some support for the proposition that the reduction
of the Pt area observed in the CVs is caused by the presence of
some Co atoms on the catalyst particle surface, which decreases
the number of Pt active sites.
Fig. 5 compares the disk polarization results for the ORR
at 1600 rpm for the different electrocatalysts, in methanol-free
electrolytes at 30 ◦ C. It can be noted on all catalysts that the
ORR is diffusion-controlled, when the potential is lower than
0.6 V versus RHE, and is under mixed diffusion-kinetic control
between 0.6 and 0.85 V. Also, it is seen that the half-wave poten-

Fig. 3. In situ XANES spectra at the Pt L3 edge for the Pt-Co/C materials com-
pared to pure Pt/C and Pt foil at 0.30 and 0.90 V vs. RHE in H2 SO4 0.5 mol L−1 .
(a) Pt-Co/C 3:1 and (b) Pt-Co/C 1:3.

Also, these results are in agreement to previous XANES studies


on Pt alloys in acid and in alkaline media [5,38].
Fig. 4 shows the cyclic voltammograms (CV) obtained at a
scan rate of 20 mV s−1 for the home-made Pt/C and Pt-Co/C cat-
alysts in 0.5 mol L−1 H2 SO4 , with the currents normalized per
mass of Pt in the catalyst layer. Results show the typical behavior
regarding the hydrogen and the oxide regions on Pt/C and Pt-
Co/C catalysts in acid solutions [35]. A stable response was seen
after two cycles and no current of Co oxidation or dissolution
was observed for potentials up to 950 mV. This indicates that the
particle surface is covered by a Pt-rich layer, and/or that the Co Fig. 5. Steady-state polarization curves for the ORR on Pt/C and Pt-Co/C alloys
atoms on the particle surface are stabilized by the interactions in O2 saturated 0.5 mol L−1 H2 SO4 electrolyte at 30 ◦ C. ω = 1600 rpm.
F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393 391

Fig. 6. Mass-transport corrected Tafel plots for the ORR on Pt/C and Pt-Co/C
alloys, normalized per mass of Pt in the catalyst layer, in O2 saturated 0.5 mol L−1
H2 SO4 electrolyte at 30 ◦ C. ω = 1600 rpm.

tials are very similar for all catalysts, even for Pt-Co/C 1:5. The
limiting current densities observed for all Pt-Co/C catalysts are
close to that for Pt/C, in which the ORR follows a four-electron
mechanism [36]. This indicates the occurrence of mainly a four-
electron ORR mechanism on all Pt-Co/C alloys, as also seen in
other works [37,38].
The electrocatalytic activities for the ORR on the Pt-Co/C
alloys were compared through mass-transport corrected Tafel
plots [36], obtained at 30 ◦ C, with the currents normalized by
the mass of platinum in the catalytic layer, as presented in Fig. 6.
Two different Tafel linear regions were observed with slopes near
to 60 and 120 mV/dec for low and high overpotentials, respec-
tively, in agreement to previous works [36]. This is explained
in terms of the coverage of adsorbed oxygen, which follows
a Temkin isotherm (high coverage) at low overpotentials and a Fig. 7. Steady-state polarization curves for the ORR on Pt/C and Pt-Co/C
Langmuir isotherm (low coverage) at higher overpotentials. The alloys in O2 saturated 0.1 mol L−1 CH3 OH/0.5 mol L−1 H2 SO4 electrolyte. (a)
results in Fig. 6 show similar catalytic activity for Pt/C and Pt- T = 30 ◦ C and (b) T = 60 ◦ C. ω = 1600 rpm.
Co/C alloys at low overpotentials (activation region), while at
high overpotential a better performance is seen for the Pt-Co/C on methanol-containing electrolyte obtained at 30 ◦ C (Fig. 7a)
materials with high Co contents. This is an interesting result showed that Pt/C has higher activity compared to those of the
because it implies a much better utilization of the Pt catalyst, Pt-Co/C catalysts, as can be seen through the more positive
especially in the case of Pt-Co/C 1:5. It is important to mention half-wave potentials. However, at 60 ◦ C (Fig. 7b), the situation
that the Tafel plots with the currents normalized by the Pt CV changes and higher methanol tolerances are presented by the Pt-
active surface area (not shown) presented the same trend for the Co/C 1:3 and 1:5 catalysts compared to Pt/C, Pt-Co/C 3:1 and
catalyst performances as those normalized by the mass of Pt. 1:1. An important point is the high methanol oxidation activity
This evidences a good proportionality between the Pt content or low methanol tolerance observed during the ORR for Pt-Co/C
and the CV surface area. 3:1 and 1:1, which will be discussed below.
With the aim of evaluating the ORR activity of these mate- Fig. 8 shows the polarization curves for methanol oxidation
rials in a methanol containing-electrolyte, polarization curves on Pt/C and Pt-Co/C catalysts in nitrogen saturated 0.5 mol L−1
were obtained at 30 and 60 ◦ C in an O2 saturated 0.5 mol L−1 H2 SO4 /0.1 mol L−1 CH3 OH solutions at 30 and 60 ◦ C. It is
H2 SO4 /0.1 mol L−1 CH3 OH solution and the results are pre- found that the methanol oxidation currents are higher for the
sented in Fig. 7. When the polarization curves for the ORR in Pt-Co/C 3:1 and 1:1 catalysts at 30 and 60 ◦ C, and these are fol-
methanol-containing electrolyte curves are compared to those lowed by Pt/C and by Pt-Co/C 1:3 and 1:5. In fact, a comparison
in pure 0.5 mol L−1 H2 SO4 , it is noted that all catalysts present between the results in Figs. 7 and 8 indicates that, for the Pt-Co/C
increased overpontentials for the ORR due to the simultane- alloys, a higher activity for the methanol oxidation corresponds
ous occurrence of the oxygen reduction and methanol oxi- to a lower activity for the ORR in the presence of methanol,
dation reactions [17]. The polarization curves for the ORR independently of the electrode potential. As a consequence,
392 F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393

On the other hand, the higher methanol oxidation activity


or low methanol tolerance during the ORR observed for Pt-
Co/C 3:1 and 1:1 (atoms) can be attributed to a more favorable
PtCo surface arrangement and/or a more appropriated atomic
Pt:Co ratio that provides Co oxides at lower potentials, on the
basis of occurrence of the bi-functional mechanism for methanol
oxidation, and/or the alcohol dehydrogenation promoted by the
superficial Co atoms, as proposed before, in the same conditions
of this work, but for Pt-Ni alloys [42]. Thus, it seems that a higher
methanol oxidation on Pt-Co/C catalysts is achieved on particles
with lower degree of alloying and/or with segregated phase of
the non-noble metal. This is supported by a previous work [18],
where low methanol oxidation activity has been observed for
a Pt-Co/C 3:1 alloy. In contrast to the present work, the Pt-
Co/C 3:1 catalyst in Ref. [18] showed higher degree of alloying
between Pt and Co.
The results for the ORR (methanol-free electrolyte) with the
currents normalized per mass of Pt (Fig. 6) indicate a superior
activity for Pt-Co/C alloys with high Co contents, particularly at
high overpotentials. This fact cannot be associated to an increase
of the Pt active area because the CV results show an opposite
trend, but the results can be understood following the Ham-
mer and Nørskov [11] model. In this case, it is recognized that
the Pt atoms in the Pt-Co/C alloys are subjected to a com-
pressive strain due to the lower Co lattice parameter and also
subjected to a high electronic interaction with the Co atoms [12].
These two effects conduct to a Pt 5 d-band broadening and, as
a result, to a down-shift on the Pt 5d band center, leading to a
decrease of the adsorption strength of oxygenated adsorbates.
Thus, the Pt–O− bonding becomes weaker, and its electrore-
duction becomes more facile, improving the ORR kinetics. This
proposal is supported by the XANES results in Fig. 3, which
shows a lower white line at high electrode potentials for the
Pt-Co/C alloys compared to Pt/C. This indicates lower Pt-O−
interaction for the Pt atoms in the alloys, in agreement to the
Fig. 8. Methanol oxidation curves on Pt/C and Pt-Co/C alloys in N2 satu-
Nørskov model. Finally, if the particle surface contains some Co
rated 0.1 mol L−1 CH3 OH/0.5 mol L−1 H2 SO4 electrolyte. (a) T = 30 ◦ C and
(b) T = 60 ◦ C. atoms, it is important to also mention the effect of the “neigh-
boring Co oxide”, as proposed previously [2,16]. According to
this, the activity enhancement of Pt alloys compared to Pt may
the higher methanol tolerance during the course of the ORR is be due to OH repulsion between Pt-OH and neighbor non-noble
observed for Pt-Co/C 1:3 and 1:5 at 60 ◦ C. Based on the present metal hydroxides or oxides, decreasing the OH coverage on Pt
results, one would expect that this is a long term effect because and thus increasing the number of free Pt active sites.
a high stability is evidenced by the cyclic voltammetric results.
As has been proposed, methanol oxidation requires at least 4. Conclusions
three neighboring Pt atoms in the proper crystallographic
arrangement to activate the chemisorption of methanol [39–41]. The particle diameter distribution determined by TEM has
For the Pt-Co/C catalysts investigated here, the Co content shown that, increasing of Co content in the Pt-Co/C catalysts,
enrichment leads to the reduction of its mean particle diameter the nanoparticles diameter distribution becomes sharper and the
and a reduction of the number of neighboring Pt atoms. Addi- mean particle diameter becomes smaller. The XRD measure-
tion of Co to Pt-based catalysts initially causes an increase of ments indicated that the lattice parameters are slightly smaller
the electrocatalytic properties for the methanol oxidation, but for for the Pt–Co/C materials compared to Pt/C, indicating a con-
high Co contents, the effect is reversed. One possibility is that traction of the Pt unit cell lattice parameter due to alloy forma-
the lower kinetics of methanol oxidation on Pt-Co/C 1:3 and 1:5 tion. However, these changes were not as pronounced as for high
is due to the reduced probability of finding three neighboring Pt temperatures heat treated Pt–Co/C materials. Also, the XRD
active sites in a proper crystallographic arrangement, to perform analysis has indicated the presence of a Co segregated phase in
the methanol oxidation reaction, compared to Pt/C or to Pt-Co/C the form of Co3 O4 mainly for Pt-Co/C 1:3 and 1:5. XANES mea-
3:1 and 1:1. surements showed that the increase of the white line, caused by
F.H.B. Lima et al. / Electrochimica Acta 52 (2006) 385–393 393

the oxide formation on Pt at high potentials, is less pronounced [13] B. Hammer, J.K. Nørskov, Adv. Catal. 45 (2000) 71.
for the Pt-Co/C alloys compared to Pt/C. These results indi- [14] J. Greeley, J.K. Nørskov, M. Mavrikakis, Annu. Rev. Phys. Chem. 53 (2002)
cate smaller Pt-oxide formation or lower Pt–O− bond strength 319.
[15] J. Zhang, M.B. Vukmirovic, Y. Xu, M. Mavrikakis, R.R. Adzic, Ange-
in the Pt alloys compared to Pt/C. The electrochemical mea- wandte 44/14 (2005) 2132.
surements indicated that the four-electrons mechanism is the [16] J. Zhang, M.B. Vukmirovic, K. Sasaki, A.U. Nilekar, M. Mavrikakis, R.R.
main pathway for the ORR on the Pt alloys. The electrocatalytic Adzic, J. Am. Chem. Soc. 127 (2005) 12480.
activity for this reaction in pure H2 SO4 electrolyte is higher for [17] W. Vielstich, V.A. Pagnin, F.H.B. Lima, E.A. Ticianelli, J. Electrochem.
Soc. 148 (2001) 502.
the Pt-Co/C materials with high Co contents. This was mainly
[18] J.R.C. Salgado, E. Antolini, E.R. Gonzalez, Appl. Catal. B Environ. 57
explained on the basis of a Pt d-band down-shift, which conducts (2004) 281.
to a weaker Pt–O− bond strength and its faster electroreduction. [19] H. Yang, C. Coutanceau, J.-M. Leger, N. Alonso-Vante, C. Lamy, J. Elec-
In the methanol-containing electrolyte, the Pt-Co/C 1:5 catalyst troanal. Chem. 576 (2005) 305.
showed the highest performance. This higher methanol toler- [20] J.-F. Drillet, A.a.J. Friedemann, R. Kotz, B. Schnyder, V.M. Schmidt, Elec-
trochim. Acta 47 (2002) 1983.
ance was attributed to the lower probability for this catalyst to
[21] N. Alonso, H. Tributsch, Nature 323 (1996) 431.
present three active Pt neighboring atoms, which decrease the [22] R. Jiang, D. Chu, J. Electrochem. Soc. 147 (2000) 4605.
activity for methanol oxidation, marginally affecting the oxygen [23] L. Xiong, A.M. Kannan, A. Manthiram, Electrochem. Commun. 4 (2002)
reduction kinetics. 898.
[24] T.J. Schmidt, H.A. Gasteiger, G.D. Stäb, P.M. Urban, D.M. Kolb, R.J.
Behm, J. Electrochem. Soc. 145 (1998) 2354.
Acknowledgments [25] E. Teixeira-Neto, K. Bergamaski, J.F. Gomes, F.H.B. Lima, W.H. Lizcano-
Valbuena, E.A. Ticianelli, E.R. Gonzalez, F.C. Nart, Braz. J. Morphol. Sci.
The authors thank the Fundação de Amparo à Pesquisa 22A (2005) 21.
do Estado de São Paulo (FAPESP), the Conselho Nacional [26] Y.P. Mascarenhas, J.M.V. Pinheiro, Programa para Calculo de Parametro
de Rede Pelo Metodo de Minimos Quadrados, SBPC, 1985.
de Desenvolvimento Cientı́fico e Tecnológico (CNPq), the
[27] A.R. West, Solid State Chemistry and Its Applications, Wiley, New York,
Coordenação de Aperfeiçoamento de Pessoal de Nı́vel Supe- 1984.
rior (CAPES) and the Brazilian Synchrotron Light Laboratory [28] J. McBreen, W.E. O’Grady, K.I. Pandya, R.W. Roffman, D.E. Sayers,
(LNLS), Brazil, for financial assistance. ETN thanks LME- Langumuir 3 (1987) 428.
LNLS for the use of the JEM-3010 ARP microscope. [29] H. Tolentino, J.C. Cezar, D.Z. Cruz, V. Compagnon-Caillol, E. Tamura,
M.C. Alves, J. Synchrotron Radiat. 5 (1998) 521.
[30] T. Ressler, J. Phys. IV C2 7 (1997) 269.
References [31] K.I. Pandya, R.W. Roffman, J. McBreen, W.E. O’Grady, J. Electrochem.
Soc. 137 (1990) 383.
[1] J. Zhang, Y. Mo, M.B. Vukimirovic, R. Klie, K. Sasaki, R.R. Adzic, J. [32] J.B.A.C. van Zon, D.C. Konigsberger, H.F.J. Van’t Blik, D.E. Sayers, J.
Phys. Chem. B 108 (2004) 10955. Chem. Phys. 82 (1985) 5742.
[2] U.A. Paulus, A. Wokaun, G.G. Sherer, T.J. Shmidt, V. Stamenkovic, V. [33] Y. Kiros, J. Electrochem. Soc. 143 (1996) 2152.
Radmilovic, N.M. Markovich, P.N. Ross, J. Phys. Chem. B 106 (2002) [34] J.R.C. Salgado, E. Antolini, E.R. Gonzalez, J. Phys. Chem. B 108 (2004)
4181. 17767.
[3] F.H.B. Lima, M.J. Giz, E.A. Ticianelli, J. Braz. Chem. Soc. 16 (2005) 328. [35] U.A. Paulus, A. Wokaum, G.G. Sherer, T.J. Schmidt, V. Stamenkovic, N.M.
[4] M. Min, J. Cho, K. Cho, H. Kim, Electrochim. Acta 45 (2000) 4211. Markovic, P.N. Ross, Electrochim. Acta 47 (2002) 3787.
[5] J. McBreen, S. Mukerjee, J. Electrochem. Soc. 142 (1995) 3399. [36] J. Perez, E.R. Gonzalez, E.A. Ticianelli, Electrochim. Acta 44 (1998) 1329.
[6] S. Mukerjee, S. Srinivasan, M.P. Soriaga, J. McBreen, J. Electrochem. Soc. [37] R.R. Adzic, in: J. Lipkowski, P.N. Ross (Eds.), Electrocatalysis, vol. 5,
142 (1995) 1409. VCH, New York, 1998, p. 197.
[7] A.S. Arico, A.K. Shukla, H. Kim, S. Park, M. Min, V. Antonucci, Appl. [38] F.H.B. Lima, E.A. Ticianelli, Electrochim. Acta 49 (2004) 4091.
Surf. Sci. 172 (2001) 33. [39] H.A. Gasteiger, N.M. Markovic, P.N. Ross Jr., E.J. Cairns, J. Phys. Chem.
[8] A.K. Shukla, M. Neergat, P. Bera, V. Jayaram, M.S. Hegde, J. Electroanal. 98 (1994) 617.
Chem. 504 (2001) 111. [40] N.M. Markovic, P.N. Ross, Surf. Sci. Rep. 45 (2002) 117.
[9] T. Toda, H. Igarashi, M. Watanabe, J. Electroanal. Chem. 460 (1999) 258. [41] E.A. Batista, G.R.P. Malpass, A.J. Motheo, T. Iwasita, J. Electroanal. Chem.
[10] M. Min, J. Cho, K. Cho, H. Kim, Electrochim. Acta 45 (2000) 4211. 571 (2004) 273.
[11] B. Hammer, J.K. Nørskov, Surf. Sci. 343 (1995) 211. [42] K.W. Park, J.H. Choi, B.K. Kwon, S.A. Lee, Y.E. Sung, H.Y. Ha,
[12] J.R. Kitchin, J.K. Nørskov, M.A. Barteau, G. Chen, J. Chem. Phys. 120 S.A. Hong, H. Kim, A. Wieckowski, J. Phys. Chem. B 106 (2002)
(2004) 10240. 1869.

You might also like