You are on page 1of 8

Electrochimica Acta 56 (2011) 1611–1618

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Effect of Mo addition on the electrocatalytic activity of Pt–Sn–Mo/C for direct


ethanol fuel cells
Eungje Lee, Arun Murthy, Arumugam Manthiram ∗
Electrochemical Energy Laboratory & Materials Science and Engineering Program, University of Texas at Austin, Austin, TX 78712, USA

a r t i c l e i n f o a b s t r a c t

Article history: Carbon-supported Pt–Sn–Mo electrocatalysts have been synthesized by a polyol reduction method and
Received 11 August 2010 characterized for ethanol electro-oxidation reaction (EOR). While the percent loading of the synthesized
Received in revised form 20 October 2010 nanoparticles on the carbon support is higher than 35%, energy dispersive spectroscopy (EDS) reveals
Accepted 29 October 2010
that the Mo contents in the nanoparticle catalysts are lower than the nominal value, indicating incom-
Available online 5 November 2010
plete reduction of the Mo precursor. X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD)
analyses reveal that the Sn and Mo exist as oxide phases at the surface layers of the nanoparticles and the
Keywords:
degree of alloying is very low. The electrochemical properties of the electrocatalysts have been evaluated
Fuel cells
Ethanol electro-oxidation reaction
by cyclic voltammetry (CV) and chronoamperometry. The catalytic activity for EOR decreases in the order
Electrocatalysts PtSnMo0.6 /C > PtSnMo0.4 /C > PtSn/C. Single cell direct ethanol fuel cell (DEFC) tests also confirm that the
Platinum-based nanoalloys PtSnMo0.6 /C anode catalyst exhibit better performance than the PtSn/C anode catalyst. An analysis of
Polyol synthesis the electrochemical data suggests that the incorporation of Mo to Pt–Sn enhances further the catalytic
activity for EOR.
© 2010 Elsevier Ltd. All rights reserved.

1. Introduction electro-oxidation reaction (MOR), has been greatly improved by


the addition of a third metal M to give the ternary electrocata-
Direct ethanol fuel cells (DEFCs) are an attractive option for lysts Pt–Ru–M. Among the number of ternary compositions tested,
power generation as ethanol offers higher theoretical energy Pt–Ru–Mo has been repeatedly reported to have enhanced cat-
density (8.0 kWh kg−1 ) and is less toxic compared to methanol alytic activity for MOR compared to Pt–Ru [8–13]. In this regard,
(6.1 kWh kg−1 ) [1]. Also, ethanol can be easily produced in large Pt–Ru–Mo has also been applied for DEFC, and it exhibits improved
quantities by chemical or bio-processes [2,3]. However, the slow catalytic activity for EOR compared to that of Pt–Ru [9,14]. How-
kinetics of the ethanol electro-oxidation reaction (EOR) on Pt anode ever, regarding the catalytic activity for EOR, the best binary catalyst
catalyst diminishes the overall performance of the DEFC system is Pt–Sn and it would be valuable to develop a ternary anode cat-
and hampers its commercialization [4]. Therefore, there have been alyst for DEFC based on Pt–Sn. Indeed, it has been found in other
considerable efforts to develop alternative anode electrocatalysts studies that the catalytic activity of Pt–Ru–Mo for EOR is inferior to
that offer high catalytic activity while lowering the cost. For exam- that of Pt–Sn [6,15].
ple, alloying of Pt with other elements such as Ti, Mo, Ru, Rh, To our knowledge, there has been no report on the catalytic
Pd, Sn, W and Ce has been studied, and among them, Sn has activity of Pt–Sn–Mo ternary electrocatalyst for EOR. Accordingly,
shown the best promoting effect on EOR [5–7]. Nevertheless, it we present here the synthesis and characterization of carbon-
is still an ongoing task to improve further the performance of supported Pt–Sn–Mo and its potential as an alternative anode
the Pt–Sn anode catalysts to the level suitable for commercializa- catalyst for DEFC, in comparison to Pt–Sn. The Pt–Sn–Mo and Pt–Sn
tion. electrocatalysts are prepared by the polyol synthesis method to
It has been a good starting point in DEFC research to benchmark obtain supported nanoparticles with high % loading on carbon.
the results of DMFC research since DMFC shares similar challenges To study the effect of Mo addition to Pt–Sn, the atomic ratio of
with DEFC, e.g., the sluggish kinetics of the catalytic oxidation Pt:Sn was fixed at 1:1 for all of the samples tested. The prepared
reaction at the anode. In DMFC research, the catalytic activity of catalysts are characterized by transmission electron microscopy
Pt–Ru, which is the best binary anode catalyst for the methanol (TEM), energy dispersive spectroscopy (EDS), thermogravimetriy
analysis (TGA), X-ray photoelectron spectroscopy (XPS), and X-
ray diffraction (XRD). The catalytic performances are evaluated by
∗ Corresponding author. cyclic voltammetry (CV), COad stripping voltammetry, chronoam-
E-mail address: rmanth@mail.utexas.edu (A. Manthiram). perometry, and single cell DEFC test.

0013-4686/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2010.10.086
1612 E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618

2. Experimental methods alytic oxidation of pre-adsorbed CO was assessed by COad stripping


voltammetry. After the cleaning scans in N2 -purged H2 SO4 solu-
The carbon-supported Pt–Sn and Pt–Sn–Mo electrocatalysts tion, the electrolyte was purged with 5% CO–95% Ar for 30 min while
with 40 wt.% metal loading were synthesized by the polyol method holding the electrode potential at 0.1 V to form a COad layer on the
as described below [16]. 0.4 g of NaOH (Fisher Chemical) was dis- surface of the catalyst. Subsequently, the purge gas was switched
solved in 1 mL of deionized (DI) water and added into 100 mL to N2 to remove the dissolved CO gas from the electrolyte solu-
of ethylene glycol (EG, Acros Organics). Required amounts of tion. After the N2 -purging for 30 min, the COad was stripped by
H2 PtCl6 ·6H2 O (Strem Chemicals Inc.), SnCl2 ·2H2 O (Acros Organ- subsequent potential scans between 0.02 and 1.2 V at a scan rate of
ics), and (NH4 )6 Mo7 O24 ·4H2 O (Alfa Aesar) were dissolved in the 0.02 V s−1 . The catalytic activity for EOR was compared by CV and
NaOH-EG solution and refluxed at 130 ◦ C for 2 h in flowing N2 gas. chronoamperometry conducted in N2 -purged 0.5 M H2 SO4 + 1 M
The color of the solution turned into dark brown indicating the ethanol electrolyte. Prior to the above experiments, the electrode
formation of colloidal nanoparticles. 120 mg of as-received carbon was cleaned by cycling 10 times between 0.2 and 0.8 V in 0.5 M
black (Vulcan XC-72R, Cabot Corp.) without any further treatment H2 SO4 and transferred to the ethanol containing electrolyte. All of
was dispersed separately in 500 mL of dilute isopropyl alcohol (IPA) the electrochemical experiments were conducted at room temper-
solution (IPA:DI water = 1:1 by volume) by ultrasonication and vig- ature.
orous stirring. The prepared nanoparticle colloid was added into The performances of the single cell DEFC systems employing
the carbon dispersion and a few drops of 2 M H2 SO4 solution were PtSn/C and PtSnMo0.6 /C as the anode catalyst were compared using
added until the pH of the mixture reached 2. The resulting dis- commercial 60 wt.% Pt/C (Alfa Aesar) and Nafion 115 membrane
persion was stirred vigorously for 2 h by heating at 70 ◦ C in order (Electrochem Inc.), respectively, as the cathode catalyst and pro-
to facilitate the attachment of the colloidal nanoparticles onto the ton exchange membrane. The electrodes consisting of gas-diffusion
carbon surface. After the stirring, the mixture was cooled down to and catalyst layers were prepared as below. The catalyst ink pre-
room temperature for filtration and the filtered residue was dried pared with the required amount of the catalyst, IPA, DI water, and
overnight in a vacuum oven at 90 ◦ C. 5 wt.% Nafion solution by ultrasonication was sprayed onto a gas-
The atomic ratios of Pt, Sn, and Mo were examined by energy diffusion layer (BASF), followed by drying in air at 90 ◦ C. The catalyst
dispersive spectroscopy (EDS, Oxford instruments) attached to a metal loading was kept at 1.0 and 0.7 mg cm−2 , respectively, for
scanning electron microscope (SEM, JEOL-JSM5610). The weight the anode and the cathode. On top of the catalyst layer, 40 mg of
percentage of the nanoparticles on carbon was measured by ther- 5 wt.% Nafion solution was sprayed additionally for better adhesion
mogravimetric analysis (TGA, Perkin Elmer). In the TGA analysis, between the membrane and the catalyst layer. The membrane-
a pre-measured amount of the carbon-supported electrocatalyst electrode assembly (MEA) was fabricated by hot pressing the anode
was heated up to 800 ◦ C in flowing air to burn out the carbon and cathode onto the membrane at 120 ◦ C for 2 min. The fuel cell
support completely, and the weight of the nanoparticle residue tests were carried out with a single-cell hardware (Scribner) with
was used to obtain the % nanoparticle loading on carbon. A scan- serpentine flow-field pattern (5 cm2 active area) and feeding 1 M
ning transmission electron microscopy (STEM, Hitachi S-5500) ethanol solution (flow rate: 2.5 mL min−1 ) and humidified oxygen
system was employed to obtain the size distribution and dispersion (flow rate: 200 mL min−1 , backpressure: 16 psi). The steady state
morphology of the supported nanoparticles. X-ray photoelectron fuel cell polarization data were collected after operating the fuel
spectroscopic (XPS, Kratos Analytical) data were acquired using cell for two consecutive days including the period of shut down
monochromatic Al K␣ X-ray source to assess the oxidation states overnight.
of Pt, Sn, and Mo on the surface layers of the samples. The spectra
were fitted with a Gaussian–Lorentzian function (G:L = 85:15) and
a Shirley background function using the XPSPEAK 4.1 software. The 3. Results and discussion
crystallographic phase analysis was carried out with X-ray diffrac-
tion (XRD, Philips). The lattice parameter values were obtained Fig. 1 shows the TEM images and nanoparticle size distribu-
from the position of the (2 2 0) peak fitted with the pseudo-Voigt tions of PtSn/C, PtSnMo0.4 /C, and PtSnMo0.6 /C. The images reveal
function (R-factor: ∼3). The (2 2 0) peak position was calibrated by the presence of both dispersed nanoparticles as well as agglom-
using the (4 4 0) peak of a Si internal standard that was mixed with erated nanoparticles. It appears that some of the finely dispersed
the sample powder. colloidal nanoparticles become agglomerated when they are non-
The electrochemical characterizations were conducted with homogeneously attached onto the carbon surface (Fig. 1d). The
a potentiostat (CH Instruments) in a conventional single- particle size distribution presented in Fig. 1 indicates that both
compartment three-electrode cell having a platinum foil counter the mean particle size and the size dispersion decrease as the Mo
electrode. A Hg/HgSO4 (MMS) electrode in a saturated K2 SO4 solu- content increases. The chemical composition and % loading of the
tion served as a reference electrode, but all the potentials reported nanoparticles supported on carbon were measured by EDS and
in this paper are referenced with respect to the normal hydrogen TGA (Table 1). The observed Pt:Sn ratios for all of the prepared
electrode (NHE). A glassy carbon (3 mm dia.) electrode was pol- samples are 1:1, which is consistent with the Pt:Sn ratio in the ini-
ished to a mirror-like finish with 0.05 ␮m alumina media (Buehler) tial reaction mixture. However, the EDS results of PtSnMo0.4 /C and
and coated with a thin layer of the catalyst to serve as a working PtSnMo0.6 /C reveal that the Mo contents in the samples are lower
electrode. In a typical experiment, the catalyst ink was prepared by than the nominal values, indicating the loss of Mo component due
dispersing 2 mg of the catalyst in a mixture of DI water, ethanol, to the incomplete reduction of Mo precursors during the synthesis.
and 5 wt.% Nafion solution (Electrochem Inc.) by ultrasonic vibra- The observed Pt:Mo atomic ratio in PtSnMo0.4 /C is 1:0.4, while the
tion. When a dark homogeneous dispersion was formed, 2 ␮L of the nominal ratio in the initial reaction mixture is 1:0.5. The Mo loss
ink was dropped onto the glassy carbon electrode and dried in air is more apparent in the PtSnMo0.6 /C, where the observed Pt:Mo
to give an effective catalyst loading of 140 ␮g cm−2 . atomic ratio is only 1:0.6 compared to the nominal ratio of 1:1.
The electrochemical surface characteristics were examined by The % loading of the nanoparticles in PtSnMo0.4 /C and PtSnMo0.6 /C
cyclic voltammetry (CV) in N2 -purged 0.5 M H2 SO4 electrolyte. The were, respectively, 39 and 37 wt.%, which are slightly less than the
electrode potential was scanned between 0.02 and 0.8 V for 10 nominal 40 wt.% possibly due to the Mo loss. However, it cannot be
cycles at a scan rate of 0.05 V s−1 in order to clean the surface, ruled out that some portion of the reduced nanoparticles might not
and the 10th scan was used for characterization. The electrocat- have attached onto the carbon surface and might have been washed
E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618 1613

Fig. 1. TEM images and nanoparticle size distributions of (a) PtSn/C, (b) PtSnMo0.4 /C, and (c) PtSnMo0.6 /C, and (d) a magnified image of the nanoparticle agglomeration in
PtSnMo0.4 /C.

Table 1
Compositional and morphological characterization of PtSn/C, PtSnMo0.4 /C, and PtSnMo0.6 /C.

Electrocatalyst Atomic ratio (Pt:Sn:Mo) Atomic ratio (Pt:Sn:Mo) Metal loading from Particle size from
based on the precursor based on EDS analysis TGA data/wt.% TEM/nm

PtSn/C 1.0:1.0 1.0:1.0 36 3.62 ± 2.31


PtSnMo0.4 /C 1.0:1.0:0.5 1.0:1.0:0.4 39 3.57 ± 1.74
PtSnMo0.6 /C 1.0:1.0:1.0 1.0:1.0:0.6 37 3.16 ± 1.39
1614 E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618

Fig. 2. X-ray photoelectron spectra of (a) Pt 4f in PtSn/C, (b) Sn 3d in PtSn/C, (c) Pt 4f in PtSnMo0.6 /C, (d) Sn 3d in PtSnMo0.6 /C, and (e) Mo 3d in PtSnMo0.6 /C.

away during the synthesis since the % loading of the PtSn nanopar- electron spectra and the binding energies (BEs) with corresponding
ticles is also less than the 40 wt.% despite that the Pt:Sn ratio is same relative intensities are presented in Fig. 2 and Table 2. Fig. 2a shows
as the nominal value. The lower % loading of PtSn nanoparticle may the Pt 4f region of the spectrum for PtSn/C. In accordance with the
be because the surface charge of PtSn nanoparticles is less attractive literature [18,19], the Pt 4f7/2 and 4f5/2 peaks are deconvoluted into
to the carbon surface compared to that of Pt–Sn–Mo nanoparticles three different Pt chemical states. The peaks with the lowest bind-
in the reaction medium with a pH of 2 [17]. Considering the above ing energies (BE of Pt 4f7/2 = 71.4 eV and BE of Pt 4f5/2 = 74.7 eV)
analysis, the samples are denoted by their observed compositions are assigned to the metallic Pt(0), while the second peaks (BE of Pt
and the catalytic activities were normalized by the actual amount 4f7/2 = 72.4 eV and BE of Pt 4f5/2 = 75.7 eV) are ascribed to oxidized
of Pt, for further discussion in this paper. Pt(II) such as PtO or Pt(OH)2 . The peaks with the highest binding
The oxidation states of Pt, Sn, and Mo in the surface layers of energies (BE of Pt 4f7/2 = 73.9 eV and BE of Pt 4f5/2 = 77.2 eV) are
the catalyst nanoparticles were examined by XPS. The X-ray photo- associated with the Pt(IV) oxide such as PtO2 . The slight shift of the

Table 2
Binding energies and relative intensities of the PtSn/C and PtSnMo0.6 /C catalysts.

PtSn/C PtSnMo0.6 /C

B.E. (eV) Relative intensity (%) B.E. (eV) Relative intensity (%)

Pt (0) 71.4 51.4 71.4 48.6


Pt (II) 72.4 30.5 72.3 37.0
Pt (IV) 73.9 18.1 74.0 14.4
Sn (0) 485.5 3.3 – –
Sn (IV) 487.1 96.7 487.0 100
Mo (VI) – – 232.8 100
E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618 1615

Pt(0) peaks to higher binding energy in comparison to that of bulk


Pt (BE of Pt 4f7/2 = 70.5 eV) [19] is due to the Pt–carbon interaction
and the effect of small particles [18–20].
Similarly, the Pt 4f spectrum of PtSnMo0.6 /C is deconvoluted
into three components with respective binding energies (BE of Pt
4f7/2 ) of 71.4, 72.3, and 74.0 eV (Fig. 2c). The addition of Mo does
not shift the Pt(0) peaks at all and induces only 0.1 eV changes
in the Pt(II) and Pt(IV) peaks. This negligible shift in the binding
energy of Pt suggests no significant electronic modification on Pt in
PtSnMo0.6 /C due to the addition of Mo. In Fig. 2b, the Sn 3d spec-
trum of PtSn/C is deconvoluted into the 3d5/2 and 3d3/2 peaks of
Sn(0) (BE of Sn 3d5/2 = 485.5 eV and BE of Sn 3d3/2 = 493.9 eV) and
Sn(IV) oxide (BE of Sn 3d5/2 = 487.1 eV and BE of Sn 3d3/2 = 495.5 eV)
[21]. Even though the intensity of the Sn(0) peaks are low, we were
not able to fit the overall Sn 3d spectrum of PtSn/C precisely without
including the Sn(0) peaks. On the other hand, the Sn 3d spectrum of
PtSnMo0.6 /C could be fitted well with a single component of Sn(IV)
oxide (BE of Sn 3d5/2 = 487.0 eV and BE of Sn 3d3/2 = 495.4 eV), as
shown in Fig. 2d, implying reduced amount of metallic Sn(0) in the
ternary sample. The difference in the binding energy of Sn 3d5/2
between PtSn/C and PtSnMo0.6 /C is only 0.1 eV, indicating negli-
gible electronic modification on Sn due to the addition of Mo. In
Fig. 2e, the Mo in the surface layers of PtSnMo0.6 /C exists as Mo(VI)
oxide (BE of Mo 3d5/2 = 232.8 eV and BE of Mo 3d3/2 = 235.9 eV)
without any metallic Mo(0) signal.
Fig. 3a shows the XRD patterns of the PtSn/C, PtSnMo0.4 /C, and
PtSnMo0.6 /C catalysts, and all the samples exhibit reflections char-
acteristics of the face centered cubic (fcc) Pt structure (space group:
Fm–3m). The weak reflections (marked with circles) observed at
around 34◦ and 52◦ are attributed to the (1 0 1) and (2 1 1) reflec-
tions of tetragonal SnO2 phase (Powder Diffraction File Card No.
41-1445). To obtain accurate lattice parameter values, a more pre-
cise scan with a step size of 0.02◦ was carried out in the range of
60◦ –74◦ after mixing the catalyst samples with Si powder as an
internal standard (Fig. 3b). The lattice parameter values obtained
from the positions of the (2 2 0) peak are presented in Fig. 3c and
Table 3. The observed lattice parameter a = 0.3981 nm of the PtSn/C
catalyst is close to the values reported for PtSn/C with similar
compositions and similar synthesis methods [6,22]. The Sn atomic
fraction (XSn ) in the Pt–Sn alloy was obtained by the relationship,
 (a − a ) 
0
XSn = XS (1)
(aS − a0 )
where aS and XS are, respectively, the lattice parameter (0.4002 nm)
of and the Sn atomic fraction (0.25) in the Pt3 Sn/C ETEK catalyst, Fig. 3. (a) X-ray power diffraction patterns of PtSn/C, PtSnMo0.4 /C, and PtSnMo0.6 /C
and a0 is the lattice parameter (0.3916 nm) of the Pt/C ETEK cata- with the vertical lines at the bottom representing the reflections corresponding to
fcc Pt, (b) (2 2 0) peak of the fcc Pt-alloys in reference to the (4 0 0) peak position
lyst [23]. Also, the degree of alloying between Pt and Sn (Snal /Sntot ),
of an Si internal standard, and (c) dependence of the fcc lattice parameter of the
which indicates the ratio of alloyed Sn (Snal ) to total Sn in the cat- Pt–Sn–Mox /C samples on Mo content.
alyst (Sntot ), is given by
Snal XSn PtEDS
= (2) Sn (rPt = 138.5 pm, rSn = 151 pm, and rMo = 139 pm) [24] or (2) the
Sntot (1 − XSn )SnEDS
decrease in the degree of alloying between Pt and Sn caused by
where PtEDS and SnEDS are, respectively, the experimentally deter- the presence of a non-alloyed Mo. The decrease in the degree of
mined Pt and Sn contents by EDS analysis. In Table 3, the degree alloying of the binary base alloy in the presence of a third non-
of alloying of Sn with Pt in PtSn/C is 23.4%, which is higher than alloyed element has been observed before in other ternary catalysts
the value expected from the XPS results that indicated only a trace [11,22,25]. Considering that the XPS spectrum of Mo shows only an
amount of Sn in the metallic Sn(0) state in the sample. While the for- un-alloyed MoO3 signal without metallic Mo(0) in PtSnMo0.6 /C and
mation of core–shell type nanoparticles, comprising of Pt(Ox )–SnO2 the metallic Sn(0) signal observed in PtSn/C disappears on going to
at the surface and Pt–Sn alloy at the core, can be suggested when PtSnMo0.6 /C, the decrease in the lattice parameter on adding Mo
the shallow depth resolution of XPS is considered, further detailed may be explained better by the second reason above, i.e., due to
investigation will be needed to unambiguously clarify the structure the decrease in the degree of alloying between Pt and Sn. Assum-
of the nanoparticles. ing no incorporation of Mo into the Pt lattice occurs, the XSn and
The lattice parameters of PtSnMo0.4 /C (0.3961 nm) and Snal /Sntot values for the PtSnMo0.4 /C and PtSnMo0.6 /C catalysts are
PtSnMo0.6 /C (0.3959 nm) are smaller than that of PtSn/C, which obtained using Eqs. (1) and (2). In Table 3, the degree of alloying
can be attributed to either (1) the incorporation of Mo into the decreases with increasing Mo content and reaches as low as 14.5%
lattice since the atomic radius r of Mo is smaller than that of in PtSnMo0.6 /C.
1616 E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618

Table 3
XRD analysis of PtSn/C, PtSnMo0.4 /C, and PtSnMo0.6 /C.

Electrocatalyst (2 2 0) position (◦ ) Lattice parametera , Sn atomic Degree of alloying


a (nm) fractionb , XSn of Snb , Snal (%)

PtSn/C 66.36(4) 0.3981(2) 0.190 23.4


PtSnMo0.4 /C 66.73(3) 0.3961(1) 0.132 15.2
PtSnMo0.6 /C 66.79(5) 0.3959(3) 0.126 14.5
a
Calculated from the (2 2 0) peak position fitted by the pseudo-Voigt function.
b
Calculated from Vegard’s law assuming no alloying of Mo with Pt.

Fig. 4 compares the electrochemical surface characteristics of lines in Fig. 5b and c). The electrochemical surface areas (ECSACO )
the prepared catalysts by the cyclic voltammograms recorded in obtained from the charge between the COad stripping and the fol-
0.5 M H2 SO4 electrolyte. The voltammogram of PtSn/C shows well- lowing curves are 49, 41, and 44 m2 gPt −1 , respectively, for PtSn/C,
defined hydrogen adsorption/desorption peaks characteristic of PtSnMo0.4 /C, and PtSnMo0.6 /C. Considering the smallest particle
polycrystalline Pt surface, indicating the low degree of alloying size value of PtSnMo0.6 /C among the prepared samples as mea-
with Sn at the surface of the nanoparticles, which is consistent sured by TEM, the lower ECSACO of PtSnMo0.6 /C in comparison to
with the XPS (Fig. 2b) and XRD (Table 3) analyses. However in the that of PtSn/C implies surface segregation of the Mo component in
Mo-containing ternary catalysts, the Pt-like surface characteristics the ternary nanoparticles, resulting in the blockage of the Pt surface
of the hydrogen adsorption/desorption region are suppressed and sites.
only a broad peak is present. Also, an increase in the double layer The catalytic activity of the prepared catalysts is compared by
capacitance is observed with increasing Mo content. Similar elec- cyclic voltammetry recorded in 0.5 M H2 SO4 + 1 M ethanol elec-
trochemical features in the hydrogen adsorption/desorption and trolyte. Fig. 6a shows the mass activity of the catalysts based on the
double layer regions have been observed for the Pt–RuO2 catalysts 10th anodic sweeps in the cyclic voltammograms normalized by
also and were ascribed to the formation of oxygenated species at the weight of Pt in the electrodes. To contrast the effect of Mo addi-
low potentials on the surface of RuO2 [26,27]. A similar explana- tion to PtSn on the activity enhancement, the data for Pt/C, which
tion may be applied to the Mo-containing ternary catalysts as the was prepared by conventional chemical reduction method [30], is
XPS data have indicated the presence of MoO3 . An additional peak also presented in Fig. 6. The mass activity of the catalysts for EOR
observed at 0.4 V is most likely due to the redox reaction of Mo decreases in the order of PtSnMo0.6 /C > PtSnMo0.4 /C > PtSn/C > Pt/C.
between Mo(VI) and Mo(IV), and this process has not been reported The PtSnMo0.6 /C catalyst shows the lowest onset potential and the
to pose a serious concern regarding the stability of the Pt–Mo cat- highest current over the entire potential range of the test. The spe-
alysts in the literature [28]. cific activities of the catalysts for EOR are also obtained using the
Fig. 5 shows the COad stripping voltammograms of the pre- ECSACO values and presented in Fig. 6b, showing the highest specific
pared catalysts. A typical COad stripping curve on PtSn/C is shown activity for PtSnMo0.6 /C as well. Thus, the higher catalytic activity of
in Fig. 5a, which is characterized by a low onset potential (0.3 V) the ternary catalysts is not only due to the morphological improve-
for COad oxidation and multiple waves in the COad stripping curve. ment (as seen in Fig. 1) but also due to the enhancement in intrinsic
The low onset potential on the PtSn/C catalyst is ascribed to the activity.
facile formation of oxygenated species on Sn sites in comparison
to Pt. According to the bifunctional mechanism, these oxygenated
species assist the removal of COad at lower potentials. Moreover,
the oxidation reaction of COad on PtSn/C occurs over a large poten-
tial range, resulting in an oxidation curve with multiple waves,
which may be attributed to a wide distribution of CO adsorption
energy [29]. The COad stripping curves of the PtSnMo0.4 /C and
PtSnMo0.6 /C catalysts reveal even lower onset potential (0.15 V)
and a clearly resolved pre-peak at 0.4 V as seen in Fig. 5b and c.
This may be related to the redox reaction of Mo since the position
of the redox peak in the following curve (dotted lines in Fig. 5b and
c) corresponds to that of the pre-peak in the stripping curve (solid

Fig. 5. COad stripping voltammograms of (a) PtSn/C, (b) PtSnMo0.4 /C, and (c)
PtSnMo0.6 /C. The solid and dotted lines represent, respectively, the stripping and
Fig. 4. Cyclic voltammograms recorded in 0.5 M H2 SO4 with a scan rate of 0.05 V s−1 . the following curves.
E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618 1617

Fig. 8. Comparison of the performances in DEFC of the PtSn/C and PtSnMo0.6 /C


anode catalysts (metal loading: 1.0 mg cm−2 ). Cathode catalyst: 60 wt.% Pt/C (metal
loading: 0.7 mg cm−2 ); membrane: Nafion 115; cell temperature: 65 and 90 ◦ C;
anode feed: 1.0 M ethanol solution at 2.5 mL min−1 ; cathode feed: humidified O2
at 200 mL min−1 with a backpressure of 16 psi.

The single cell DEFC performances of the PtSn/C and PtSnMo0.6 /C


anode catalysts were evaluated at the cell temperatures of 65 and
90 ◦ C (Fig. 8). In agreement with the CV and chronoamperome-
try data, the PtSnMo0.6 /C anode catalyst exhibits higher single cell
performance than PtSn/C in the DEFC tests. For example, the cell
with the PtSnMo0.6 /C anode shows open circuit potentials (OCP)
of 0.57 and 0.70 V, respectively, at 65 and 90 ◦ C, which are higher
than those exhibited by the cell with the PtSn/C anode (0.54 and
0.64 V, respectively, at 65 and 90 ◦ C). The cell with the PtSnMo0.6 /C
anode also exhibits higher maximum output power density values
Fig. 6. Anodic sweep curves of cyclic voltammograms, comparing (a) mass activ- of 14.3 and 25.7 mW cm−2 , respectively, at 65 and 90 ◦ C compared
ity and (b) specific activity of the catalysts for EOR (electrolyte = 0.5 M H2 SO4 + 1 M to 6.7 and 17.8 mW cm−2 , respectively, at 65 and 90 ◦ C for the cell
ethanol, scan rate = 0.02 V s−1 ).
with the PtSn/C anode.In general, the promoting effect of a second
element on the alcohol electro-oxidation reaction on Pt has been
explained by the bifunctional effect, electronic modification, and
In Fig. 7, the chronoamperometric profiles obtained at the elec- geometrical effect [4,5,31,32]. In the theory of electronic modifi-
trode potential of 0.5 V compares the quasi-steady state activities cation effect, the promoting effect of a second element is ascribed
of the synthesized catalysts for EOR. A potential step from the open to the modification of the electronic band structure of Pt and sub-
circuit potential (OCP) to 0.8 V was applied for 1 s in ethanol con- sequent reduction in the binding strength of intermediate species
taining electrolyte and subsequently 0.02 V was applied for 2 s to on Pt. In the geometrical consideration, the activity enhancement
reduce any adsorbed oxides/hydroxides produced on the electrode is attributed to the modification of the Pt–Pt bond distance, which
at 0.8 V. Then, a target potential of 0.5 V was applied for 1800 s facilitates the breakage of bonds such as the C–C bond in ethanol. In
to achieve a quasi-steady state current density profile. While the our study, however, XPS did not provide any critical evidence for the
PtSnMo0.6 /C shows the highest chronoamperometric current value electronic modification of PtSn by Mo addition, and the observed
for the tested time period, all of the tested samples represent similar lattice parameter change is not large enough for the geometrical
rates of current decay. effect to fully explain the huge enhancement in the catalytic activ-
ity. Furthermore, it is not likely that the observed lattice parameter
changes will have an effect on the Pt–Pt distance of the atoms at
the nanoparticle surface since both Sn and Mo are found to exist as
unalloyed oxide phases at the surface. Therefore, the bifunctional
mechanism may be the origin of the enhanced EOR activity of the
Pt–Sn–Mo/C catalyst on adding Mo. While the bifunctional mecha-
nism of Pt–Sn (or Pt–SnO2 ) on the alcohol oxidation has been well
established in many reports [4,31,33–35], the bifunctional mecha-
nism of Pt–Mo (or Pt–MoOx ) on the oxidation of alcohol or CO has
been suggested in recent reports as follows [7,36]:

Mo + H2 O → Mo(OH) + H+ + e− (3)
+ −
Pt–COad + Mo(OH) → Pt + Mo + CO2 + H + e (4)

or

MoOx + H2 O → (MoOx )–OHad + H+ + e− (5)


Fig. 7. Chronoamperometric profiles of the prepared catalysts for EOR recorded at
0.5 V in 0.5 M H2 SO4 + 1 M ethanol electrolyte. Pt–COad + (MoOx )–OHad → Pt + MoOx + CO2 + H+ + e− (6)
1618 E. Lee et al. / Electrochimica Acta 56 (2011) 1611–1618

In reactions (3) or (5), Mo or MoOx promotes the water acti- References


vation, resulting in the generation of OH species to facilitate the
oxidation of intermediates species such as COad as evidenced by a [1] W.J. Zhou, S.Q. Song, W.Z. Li, Z.H. Zhou, G.Q. Sun, Q. Xin, S. Douvartzides, P.
Tsiakaras, J. Power Sources 140 (2005) 50.
low onset potential in the COad stripping voltammograms. [2] F. Vigier, S. Rousseau, C. Coutanceau, J.M. Léger, C. Lamy, Top. Catal. 40 (2006)
Even though the bifunctional effect, which provides an expla- 111.
nation for the promoting effect in the binary Pt–Sn and Pt–Mo [3] J. Riberio, D.M. dos Anjos, K.B. Kokoh, C. Coutanceau, J.M. Léger, P. Olivi, A.R. de
Andrade, G. Tremiliosi-Filho, Electrochim. Acta 52 (2007) 6997.
catalysts, may also be applicable for the ternary Pt–Sn–Mo catalyst, [4] F. Vigier, C. Coutanceau, F. Hahn, E.M. Belgsir, C. Lamy, J. Electroanal. Chem. 563
it is not clear yet why the ternary catalysts with large fractional (2004) 81.
amount of the promoting elements (e.g., Pt:(Sn + Mo) = 1:1.6 in [5] E. Antolini, J. Power Sources 170 (2007) 1.
[6] W.J. Zhou, W.Z. Li, S.Q. Song, Z.H. Zhou, L.H. Jiang, G.Q. Sun, Q. Xin, K. Poulianitis,
PtSnMo0.6 /C) have better catalytic activity than PtSn/C (Pt:Sn = 1:1), S. Kontou, P. Tsiakaras, J. Power Sources 131 (2004) 217.
while Pt–Sn with large fractional amount of Sn does not. For exam- [7] D. dos Anjos, K. Kokoh, J.M. Léger, A.R. Andrade, P. Olivi, G. Tremiliosi-Filho, J.
ple, it has been reported by a few groups that the EOR activity of Appl. Electrochem. 36 (2006) 1391.
[8] M. Götz, H. Wendt, Electrochim. Acta 43 (1998) 3637.
PtSnx decreases when x is larger than 1 [15,23,37]. Therefore, a sim-
[9] A.O. Neto, E.G. Franco, E. Arico, M. Linardi, E.R. Gonzalez, J. Eur. Ceram. Soc. 23
ple combination of the two types of the bifunctional mechanisms (2003) 2987.
(Pt–Sn and Pt–Mo) may not fully describe the EOR mechanism for [10] A. Lima, C. Coutanceau, J.M. Léger, C. Lamy, J. Appl. Electrochem. 31 (2001) 379.
the Pt–Sn–Mo catalysts, and hydrogen spill-over mechanism can [11] Z. Jusys, T.J. Schmidt, L. Dubau, K. Lasch, L. Jörissen, J. Garche, R.J. Behm, J. Power
Sources 105 (2002) 297.
also be considered since the MoOx can easily form a bronze phase [12] K. Lasch, L. Jörissen, J. Garche, J. Power Sources 84 (1999) 225.
with hydrogen. According to the literature [38,39], hydrogen on [13] X. Zhang, F. Zhang, K.Y. Chan, J. Mater. Sci. 39 (2004) 5845.
the surface of Pt sites could migrate over to the MoOx sites to form [14] Z.B. Wang, G.P. Yin, Y.G. Lin, J. Power Sources 170 (2007) 242.
[15] W.J. Zhou, Z.H. Zhou, S.Q. Song, W.Z. Li, G.Q. Sun, P. Tsiakaras, Q. Xin, Appl. Catal.
Hx MoO3 , liberating the Pt active sites, which otherwise would have B: Environ. 46 (2003) 273.
been poisoned by hydrogen during the electro-oxidation of alco- [16] K.-S. Lee, H.-Y. Park, Y.-H. Cho, I.-S. Park, S.J. Yoo, Y.-E. Sung, J. Power Sources
hol. Therefore, the enhanced catalytic activity of PtSnMo0.4 /C and 195 (2010) 1031.
[17] H.-S. Oh, J.-G. Oh, H. Kim, J. Power Sources 183 (2008) 600.
PtSnMo0.6 /C may be due to the cooperative result from the bifunc- [18] V. Alderucci, L. Pino, P.L. Antonucci, W. Roh, J. Cho, H. Kim, D.L. Cocke, V.
tional effect of Sn and Mo and the hydrogen spillover effect of Mo. Antonucci, Mater. Chem. Phys. 41 (1995) 9.
[19] Y. Zhou, M. Nakashima, J.M. White, J. Phys. Chem. 92 (1988) 812.
[20] C. Roth, M. Goetz, H. Fuess, J. Appl. Electrochem. 31 (2001) 793.
4. Conclusions [21] Z.L. Liu, L. Hong, S.W. Tay, Mater. Chem. Phys. 105 (2007) 222.
[22] E. Antolini, F. Colmati, E.R. Gonzalez, Electrochem. Commun. 9 (2007) 398.
Carbon-supported Pt–Sn–Mo electrocatalysts have been syn- [23] F. Colmati, E. Antolini, E.R. Gonzalez, Appl. Catal. B: Environ. 73 (2007) 106.
[24] J.A. Dean (Ed.), Lange’s Handbook of Chemistry, 15th ed., McGraw-Hill, New
thesized by the polyol method. While EDS reveals that the Mo York, 1999.
content is lower than the nominal value, most of the Sn and Mo [25] R. Venkataraman, H.R. Kunz, J.M. Fenton, J. Electrochem. Soc. 150 (2003) A278.
are found as un-alloyed oxides on the surface of nanoparticles [26] L. Cao, F. Scheiba, C. Roth, F. Schweiger, C. Cremers, U. Stimming, H. Fuess, L.
Chen, W. Zhu, X. Qiu, Angew. Chem. Int. Ed. 45 (2006) 5315.
as revealed by XRD and XPS analyses. The CV and COad stripping
[27] F. Peng, C. Zhou, H. Wang, H. Yu, J. Liang, J. Yang, Catal. Commun. 10 (2009) 533.
voltammetry of the ternary Pt–Sn–Mo catalysts clearly show the [28] S. Mukerjee, S.J. Lee, E.A. Ticianelli, J. McBreen, B.N. Grgur, N.M. Marković, P.N.
surface characteristics of Pt modified by Sn and Mo. Among the Ross, J.R. Giallombardo, E.S. De Castro, Electrochem. Solid-State Lett. 2 (1999)
catalysts investigated, the PtSnMo0.6 /C exhibits the highest EOR 12.
[29] X. Li, X. Qiu, H. Yuan, L. Chen, W. Zhu, J. Power Sources 184 (2008) 353.
activity, followed by the PtSnMo0.4 /C and PtSn/C. The single cell [30] E. Lee, I.-S. Park, A. Manthiram, J. Phys. Chem. C 114 (2010) 10634.
DEFC test also confirms the better performance of the PtSnMo0.6 /C [31] T. Frelink, W. Visscher, J.A.R. Vanveen, Surf. Sci. 335 (1995) 353.
catalyst in comparison to PtSn/C. From the electrochemical tests [32] V. Ponec, Appl. Catal. A: Gen. 222 (2001) 31.
[33] C. Lamy, A. Lima, V. LeRhun, F. Delime, C. Coutanceau, J.M. Léger, J. Power
and single cell test, the carbon-supported Pt–Sn–Mo catalysts offer Sources 105 (2002) 283.
the potential to be considered as an alternative anode catalyst for [34] G.R. Salazar-Banda, H.B. Suffredini, L.A. Avaca, S.A.S. Machado, Mater. Chem.
DEFC. Phys. 117 (2009) 434.
[35] E.M. Crabb, R. Marshall, D. Thompsett, J. Electrochem. Soc. 147 (2000) 4440.
[36] Y. Wang, E.R. Fachini, G. Cruz, Y. Zhu, Y. Ishikawa, J.A. Colucci, C.R. Cabrera, J.
Acknowledgements Electrochem. Soc. 148 (2001) C222.
[37] J.H. Kim, S.M. Choi, S.H. Nam, M.H. Seo, S.H. Choi, W.B. Kim, Appl. Catal. B:
Environ. 82 (2008) 89.
Financial support by the Office of Naval Research MURI grant No. [38] H. Zhang, Y. Wang, E.R. Fachini, C.R. Cabrera, Electrochem. Solid-State Lett. 2
N00014-07-1-0758 and Welch Foundation grant F-1254 is grate- (1999) 437.
fully acknowledged. [39] Z.-B. Wang, G.-P. Yin, Y.-G. Lin, J. Power Sources 170 (2007) 242.

You might also like