Anda di halaman 1dari 11

Chevrel phases and chalcogenides

N. Alonso-Vante
Volume 2, Part 5, pp 534543
in
Handbook of Fuel Cells Fundamentals, Technology and Applications
(ISBN: 0-471-49926-9)
Edited by
Wolf Vielstich
Arnold Lamm
Hubert A. Gasteiger
John Wiley & Sons, Ltd, Chichester, 2003

Chapter 36
Chevrel phases and chalcogenides
N. Alonso-Vante
Laboratory of Electrocatalysis, University of Poitiers, Poitiers, France

1 INTRODUCTION
The interest in developing material with electrocatalytic
activity toward multi-electron charge transfer has increased
due to the technical need to convert the chemical energy
stored in fuel cells (FCs) to electricity, efficiently and at
low temperature. One key problem of such systems is
the development of cathodes with a high rate and selectivity for oxygen reduction reaction (ORR), either for
H2 /air, in a polymer electrolyte membrane (PEMFC), or
for CH3 OH/air, in a direct methanol FC (DMFC) system.
It is, however, recognized that platinum and the platinumgroup transition metals are the most active materials for
the ORR process. Platinum and platinum-based materials
are the main components for the anode and cathode of the
technically most advanced FC existing to date. However,
the understanding of the underlying electrochemical interfacial phenomena, at the anode and/or cathode, is far from
being understood. A plea has been addressed to perform
such studies in a review of oxygen electrochemistry by
Tarasevich et al.[1] Since then more insight into the ORR
has been gained thanks to electrochemical measurements
on well-defined surfaces. In fact, in order to probe that
the kinetics of the ORR can be affected by the adsorption of species in solution and on the surface orientation,
data for the ORR have been generated on Pt(hkl ),[2] as
well as on epitaxially grown Pd layer onto a Pt (111)
surface.[3] On the other hand, epitaxial ultra-thin layers of
Pd onto Au(100) and (111) showed beneficial effects for
ORR[4] although Au has a low activity. Aside from the
numerous works reported on polycrystalline electrodes (see
Ref. [5]) the reactivity of Pt/M alloys (M = Fe, Pd, Ni,

Cr, Mn, and Co) has been addressed by various groups


and constitutes, up to now, a major area of research.[6]
This latter approach is intended to enhance the catalytic
activity via the modification of the electronic structure
(d-band) of Pt. The increase of the d-band vacancy per
atom, as a function of the electron affinity of the alloying element, has indeed been probed via X-ray absorption
spectroscopy measurements by Mukerjee et al.[6d] In this
work, the authors correlated the catalytic current of the
oxygen electrode with the coordination distance of platinum RPt/Pt (and d-band vacancy) of various Pt/M alloys.

A volcano-type plot with a maximum at ca. RPt/Pt = 2.72 A


(and 0.37) was the result. These values correspond to
the optimized condition for Pt/Cr alloy. The volcano-type
plot also shows the role of the adsorption of the species
(H, OH) present in the electrolyte. Furthermore, the dband vacancy parameter decreases with the increasing size
of platinum nanodivided material (particle size effect).[7]
Such electronic and geometric factors are also modulated
by the applied electrode potential and species present in
the electrochemical cell. These parameters must also be
operative on Pt nanograins,[8] high surface Pt/Vulcan carbon catalysts,[9] or Pt particles covered with perfluorinated ionomer[10] or embedded into a conducting polymer
matrix.[11]
The process of ORR selectivity is not sustained on
platinum-based cathodes when methanol is present in the
electrolyte, or in a DMFC due to the cross-over effect. For
this reason, electrocatalysts based on a nonplatinum basis
have been developed. Ruthenium has also been examined
for the ORR in massif,[12] or in a nanodivided form[13]

Handbook of Fuel Cells Fundamentals, Technology and Applications, Edited by Wolf Vielstich, Hubert A. Gasteiger, Arnold Lamm.
Volume 2: Electrocatalysis. 2003 John Wiley & Sons, Ltd. ISBN: 0-471-49926-9.

Chevrel phases and chalcogenides 535


in alkaline and in acid media, respectively. The reaction proceeds primarily by a direct four-electron pathway.
However, its kinetic performance strongly depends on the
oxygen coordinated to the surface atoms (formation of
Rux Oy ) on the massif as well as on the nanodivided
material.
Much attention has also been given to the search for
catalysts based on Co/N4 chelates[14] or Fe/N4 chelates[15]
applied to a graphite base. The best among such chelates
are the porphyrins. These catalysts are more effective in an
alkaline medium than in a medium with acid electrolytes.
Curiously enough, these substances are more stable and
give better catalysis toward the ORR process when heattreated (600800 C) in He or Ar atmospheres.[15] Pyrolysis
actually destroys the organic structure leading to the formation of metallic particles deposited onto the carbon matrix
which remains protected by graphite shells.[15b] Studies,
under the same conditions, on the Fe-porphyrin molecule
showed a selectivity toward the ORR in the presence of
methanol.[15e] However, such a system is apparently not
completely reproducible.
Chevrel phases based on metal clusters constitute another
class of materials which have also been tested for ORR[16]
and selectivity.[16d] Here again electronic and geometric factors are responsible for electrocatalysis. In this connection,
novel materials were tailored within a cluster-like structure and further developed in powder[17] or colloidal
form.[18] This chapter presents the underlying background
and current developments of methanol tolerant electrocatalysts, based on ruthenium cluster materials for the
ORR.
Mo4Ru2Se8

2 THE CHEVREL PHASES:


ELECTROCATALYSIS FOR THE ORR
Ternary molybdenum chalcogenide compounds were first
reported in 1971 by Chevrel, Sergent and Pringent,[19] and
therefore they are often referred to as Chevrel phases. Their
general formulae are: Mx Mo6 X8 (M = Fe, Cu, Ag, Co,
Pb, rare earth, etc. and X = S, Se or Te). Their remarkable superconducting properties with large critical magnetic
fields attracted enormous interest in the 1970s.[20] This
interest has spread leading to the investigation of other
properties, such as their electronic band structure,[21, 22]
crystal structure,[20, 23] as a cathode for Li-batteries,[24]
hydrodesulfurization catalysis,[25] and electrocatalysis.[16]
The ternary phases are closely related to the binary chalcogenide structure: Mo6 X8 (X = S, Se and Te), consisting
of [Mo6 ] octahedron clusters surrounded by eight chalcogens arranged in a distorted cube (see Figure 1a). These
compounds are thermally produced by solid-state reactions.
Depending on the substance to be prepared, the procedure differs among laboratories, but in general the starting
materials are, e.g., molybdenum dichalcogenide and molybdenum metal, all in powder form. The reaction is then
performed in vacuum evacuated, sealed, silica tubes heated
at initial temperatures of ca. 500 C and then subjected to
reaction at temperatures 900 C for several days.

2.1 Characteristics of the material


As depicted in Figure 1(b), the Chevrel phase possesses a
three-dimensional framework due to the interconnectivity

Electrolyte

Se
Mo, Ru

R = 2.659

b
O2+ 4H+
4e

c
2H2O

(a) R = 2.598

R = 2.710

(b)

Figure 1. (a) Cluster unit based on Mo4 Ru2 Se8 . The intracluster distances RM M and RM Se are shown together as the schematic
interfacial representation of the process of charge transfer for ORR. (b) A stereo plot illustrating the interconnectivity of the cluster
units, either Mo6 Se8 or Mo4 Ru2 Se8 . The ternary compounds (Mx Mo6 Se8 ) are formed by hosting metal atoms M in the cavities marked
by triangles. The plane a b c indicates the surface (11, 2, 0).

536 Part 5: The oxygen reduction/evolution reaction


of Mo6 X8 cluster units. The linkage of the Mo6 X8 units
that allow
leads to short MoMo distances (3.13.6 A)
for weak metalmetal intercluster interactions. The free
cavities that can be filled by metal atoms (see triangles
in Figure 1b), in which metals can be hosted (insertion),
give rise to the ternary compound Mx Mo6 X8 (0 < x < 4).
On the other hand, two or four molybdenum atoms can be
replaced by ruthenium or rhenium (substitution) giving rise
to the so-called pseudo-ternary or pseudo-binary or mixed
cluster compounds, e.g., (Mo4 Ru2 )X8 and (Mo2 Re4 )X8 .
Both insertion and/or substitution (tuned by changing the
nature and concentration of the metal atoms) lead to a contraction of the cluster which then becomes more regular.
This is concomitant with the increase of the valence electron
concentration (VEC), thus the number of electrons available
for metalmetal bonding. This property confers the structure a degree of electronic and geometric flexibility, with an
interesting impact on their physical properties investigated
so far.
According to the results of molecular orbital (Mo) calculations in the Mo6 X8 cluster units, all bonding states are
occupied by 24 e per Mo6 -cluster unit (4 e per molybdenum atom). Band structure calculations[21, 22] reveal the
existence of a band gap (Eg ) in that particular electron
count. Electron counting rules are based on a simplified ionic-covalent model.[26] The chalcogen X can be
assigned a formal oxidation state of 2, i.e., 16 negative
charges for the Mo6 -cluster unit which are balanced by the
valence electrons of the metal atoms (e.g., Mo). Therefore, for Mo6 X8 the remaining electrons in the intracluster
bonding are: (6 6) 16 = 20 or 3.33 e per molybdenum atom. For the pseudo-ternary cluster (rutheniumcontaining), Mo4 Ru2 Se8 : (4 6) + (2 8) 16 = 24, i.e.,
4 e per cluster. This latter is a semiconductor (Eg =
1.3 eV) as initially reported some years ago.[16b] This
magic of number 4 is also met in the ternary compound Cu4 Mo6 S8 , but its semiconducting nature has not
yet been confirmed. These electrons are essentially derived
from the transition metal d-states[22] and available at the
Fermi level, EF . The increased density of states at EF is
attained with a VEC of 4 e per cluster, as demonstrated
by valence band spectra measurements.[16b, 27] The delocalization of electrons in the cluster offers a pool of charges
that can be engaged in an electrocatalytic reaction with a
minimum electronic relaxation (loss). In this sense, Chevrel
phases can be considered as models for the investigation of
charge transfer catalysis based on metal center-clusters (see
below).
From the point of view of material science, it has
been demonstrated that the molybdenum cluster can be
partially substituted by Ru, Re and Rh, whereas a complete substitution by other metals (e.g., W) has failed.

However, a substitution with rhenium based cluster compounds, Re6 X8 L8 , has been reported: [X = Se; L = Cl2 ,
Br2 ;[28] X = S and L = Cl2 .][29] The semiconducting nature
(Eg ) of these d-band state clusters was assessed to be 1.42,
0.85 and 1.65 eV, respectively. Novel compounds based on
Chevrel phase structures have also been reported recently:
Ti2 Nb6 O12 ,[30] and Nbx Ru6x Te8 .[23c] Furthermore, a new
synthesis route to prepare nickel Chevrel phases with grain
sizes of 200300 nm has recently been reported.[25e]

2.2 The electrochemistry of ORR on Chevrel


phases
Electrocatalysis of the ORR on some Chevrel phases
was reported for the first time in 1986 by Alonso-Vante
and Tributsch.[16a] Most of these pioneer works[16] were
performed on sintered powder materials synthesized as
described briefly above. The faradaic yield measured via
the rotating ring disk electrode (RRDE) on rutheniumcontaining clusters in 0.5 M H2 SO4 revealed that the main
channel is the 4 e charge transfer process (96%), thus leading to the formation of water, in accordance to the model
depicted in Figure 2(a).[16b] Moreover, Figure 2(b) contrasts the electrocatalytic activity of Mo6 X8 and Mo4 Ru2 X8
(X = Se and Te) clearly, illustrating the electronics, the
geometry and the chemistry at the interface.
All these will be the basis for the discussion on electrocatalysis on cluster-like materials. The molybdenum metal
(body centered cubic (bcc) structure) is shown for the
sake of comparison. As observed, a significant improvement for ORR was obtained going from Mo to Mo6 X8 and
Mo4 Ru2 X8 (X = Se and Te). The first remarkable thing
(independent of the chalcogen nature) is the fact that the
overpotential (read at a current density of 102 mA cm2
in Figure 2b) for ORR on the base metallic cluster (Mo6 )
is decreased by ca. 0.4 V in comparison to the metal.
Secondly, when the catalytic center is enriched by electrons
(from 3.3 to 4 e per metal atom in the cluster) the overpotential is further decreased (ca. 0.76 V for Mo4 Ru2 Se8
and 0.6 V for Mo4 Ru2 Te8 ) again in comparison to the Mo
metal, at 102 mA cm2 . The increasing phenomenon of
electrocatalytic activity could be explained in terms of geometric/electronic factors through electrochemical measurements in a whole family of Chevrel phase compounds,[16d]
which is supported by the density of states (DOS) at the
Fermi level (see Figure 2c for the Se compounds). However, the question is: where does the difference in the
overpotential (e.g., between Te- and Se-containing material) come from? The origin of this phenomenon is due
to the fact that the open circuit potential becomes limited
by the presence of a mixed-potential (corrosion potential).
The corrosion process is complex and can only be partially

Chevrel phases and chalcogenides 537


k1
O2

O2,ad

k2

H2O2,ad

k3

H2O

k4
(a)

Energy (eV)

103

i (mA cm2)

(1)
(2a)

102

(2b)
101

(3b)

(b)

Mo4Ru2Se8

5
Mo6Se8
10

(3a)
15

100
101
0.0

EF

H2O2

0.2

0.4

0.6

0.8

20

1.0

E (V(RHE))

(c)

10

15

DOS (a.u.)

Figure 2. (a) The kinetic model proposed by Damjanovic et al.[5a] for the molecular oxygen reduction. (b) Mass transfer corrected
current versus the electrode potential for ORR measurements of Chevrel phase compounds: Curves: (1) Mo metal; (2a) Mo6 Se8 ; (2b)
Mo6 Te8 ; (3a) Mo4 Ru2 Se8 ; (3b) Mo4 Ru2 Te8 . (Adapted from Ref. [16b, c, e]) (c) Valence band spectra (showing the DOS in the Mo and
Ru 4d bands in the energy interval from 0 to 9 eV) of Mo6 Se8 and Mo4 Ru2 Se8 . (Adapted from Ref. [27].)

represented, in a first approach, due to the anodic dissolution of molybdenum in water:


Mo6x Mx X8 Mo6x Mx X8 + zMo3+ ze
(1a)
zMo3+ + yH2 O Moz Oy + 3ze + 2yH+

(1b)

The formation of MoO3 (z = 1; y = 3) according to


equation (1b) takes place as revealed by XPS surface
analysis.[16b] Moreover, such process can be present in all
molybdenum-containing compounds and does not explain
the interfacial behavior difference (e.g., between Te and Se
in Mo4 Ru2 X8 ). Curves (3a) and (3b) in Figure 2(b) show
a similar Tafel slope in the region of activation, indicating
that, from a mechanistic point of view, similar catalytic
centers are involved for electrocatalysis. Here, it seems
evident that the chalcogen nature also has an important
implication in electrocatalysis. The results of studies of
DOS for Te-containing clusters are lacking. Tellurium
must be strongly involved in the electronic structure of
Mo4 Ru2 Te8 in order to reduce the DOS at the Fermi level.
Therefore, we assume that its position must be between that
of Mo6 Se8 and that of Mo4 Ru2 Se8 (see Figure 2c). This
idea is supported by ultraviolet photoelectron spectroscopy
experiments performed on Mo2 Re4 Se8 ,[27] where DOS
remains lower than Mo4 Ru2 Se8 but higher than Mo6 Se8 .
However, we cannot rule out a corrosion of chalcogens
parallel to that of molybdenum. It should be noted that

oxidation of Te2 to Te is more difficult than that of Se2


to Se (E = 0.14 V).[31]
Figure 3 (an extension to Figure 2a) summarizes the
electrocatalytic activity of Chevrel phases in terms of
the overpotential, (at 102 mA cm2 ) as a function
of the VEC per cluster unit. The partial substitution of
molybdenum atoms (increase in the number of electrons
in the cluster) confirms again that electron delocalization
favors the electrocatalysis of the ORR in the pseudobinary series (Mo6x Mx )Te8 . On the other hand, on the
ternary series Mx Mo6 Se8 , although the VEC increases, the
bimetallic interaction (M/Mo) does not seem to be efficient.
On the binary series Mo6 X8 , the trend in the overpotential
is as follows: Te < Se < S. As discussed above, all these
generated data stress that efficient electrocatalysis for
ORR can be performed on cluster materials in which a
reservoir for electrons must be available at the Fermi
level, i.e., DOS should be significant. In this connection,
Chevrel phases served as a model to understand the
interplay between electronic and geometric factors. The
cluster material with the highest electrocatalytic activity
is Mo4 Ru2 Se8 . Its currentpotential characteristic has the
same shape as that of platinum electrodes. The overpotential
(at 102 mA cm2 ), compared to massif platinum, is still
high (see arrow in Figure 3).
This aspect is the main limitation due to reaction with
water (equation (1a,b)) of the Chevrel phases. However,
from a fundamental material science point of view, this
class of model materials allows us to gain some insight

538 Part 5: The oxygen reduction/evolution reaction

300
Pt
Rh

450

) Ru

) Re

Se

Pt

(mV)

600

Os

750
900

Mo6

Cu

Ag

In
Pd

1050

Te

Mo
1200
19

20

21

22

23

24

25

Valence electron counting (VEC)

Figure 3. Dependence of the overpotential, at a current density of


102 mA cm2 , as a function of valence electron counting (VEC)
per cluster for Chevrel phases. Binary series: Mo6 X8 (X = S, Se
and Te); Ternary series Mx Mo6 S8 (Mx = Ag, Cu2 , Pd and In);
pseudo-binary series Mo6x Mx X8 (for X = Te; Mx = Pt0.2 , Os0.6 ,
Rh1 , Ru2 ; X = Se; Mx = Ru2 , Re4 ). Mo and Pt metals are shown
for comparison. (Adapted from Ref. [16b, d, e].)

into the interfacial process. In this respect, edge X-ray


absorption fine structure (EXAFS) data have been generated in an in situ electrochemical environment.[32] The
analysis on Ru and Mo edges as a function of applied
electrode potential revealed that ruthenium centers were
preferentially modified by the presence of oxygen in the
electrolyte. Furthermore, the coordination distance variation RMo/Mo between argon and oxygen of molybdenum RMo/Mo , in the potential range of electrocatalytic
current, was not significant. The coordination distance
change, RRu/Ru , however, for RRu/Ru , was evident and
attained a value of ca. 1%. The RRu/Se followed the
same trend as that of RRu/Ru , thus, assessing a certain
dynamics of the cluster unit for ORR and the identification of the electrocatalytic center: the ruthenium. However,
the nature of the surface complex species generated during electrocatalysis at the electrocatalytic centers is still
unknown.

3 NOVEL CLUSTER-LIKE
CHALCOGENIDE MATERIALS
Based on the knowledge gained on Chevrel cluster materials our strategy was to tailor electrocatalytic materials in
mild working conditions (T 473 K), with the components
already involved in the mixed cluster phase (Mo4 Ru2 Se8 ).
As briefly mentioned above, Chevrel phases are synthesized
at temperatures T 1173 K. For this reason, we have taken
into account the reactivity of some commercially available transition metal carbonyl compounds, such as, e.g.,

Mo(CO)6 , Ru3 (CO)12 and Os3 (CO)12 . These compounds


decompose in mild conditions (T 453 K), i.e., at the boiling points of some organic solvents. The use of carbonyl
compounds for the synthesis of dichalcogenide transition
metal semiconducting materials, such as FeS2 , was very
informative.[33] However, since Chevrel phases lay at the
interface between the molecular cluster compounds[34] and
metals, the concept behind tailoring electrocatalyst compounds was to obtain a material in which metalmetal
bonds are mostly involved, i.e., forming cluster-like materials. It was first thought that such a condition could be
fulfilled using a molecular cluster: Ru3 (CO)12 and hence in
decomposing this complex in an organic solvent, the free
CO sites could be available to be coordinated to a ligand
(e.g., a chalcogen X previously dissolved). The strategy
of using a higher metal to chalcogen ratio led, in fact, to
the synthesis of, e.g., Rux Sey cluster-like compounds active
for ORR. This was reported for the first time in 1991.[35]
This material was obtained in powder or in thin layer
form (Figure 4a,b). Very recently it has been recognized
that the chemical precursor was not the initial carbonyl
complex [Ru3 (CO)12 ] reacting with the chalcogen.[36] It
turned out that the main chemical precursor with high
nuclearity, Ru4 Se2 (CO)11 (depicted in Figure 4c), is formed
in situ in the reaction vessel which led to the novel cluster
based metallic compound embedded in selenium, Rux Sey
(with x 2, y 1). This molecular cluster compound was
detected in monitoring, as a function of the reaction synthesis, the intermediates formed via 13 C NMR and FT-infrared
spectroscopies.[37] Such a molecular cluster was previously
reported by Layer et al.[38] These authors used the pyrolysis at 185 C under vacuum of Ru3 (CO)12 and PhSeSePh
during 19 h to synthesize Ru4 Se2 (CO)11 . Our method of
fabricating this complex was found to be easier.
The essential characteristics of ruthenium based materials, in powder and in the colloidal form are their high
dispersiveness in the nanoscale range (Figure 4a) and the
simplicity of manufacturing them.[17, 18, 36]
Tailoring the material was also possible using X = S,
Te and Se.[17] Their electroactivity with respect to the
molecular ORR and their chemical stability are higher than
those of Chevrel phase materials.[39, 40] A typical ORR
experiment via the rotating disk electrode (RDE) on Rux Sey
thin layers is depicted in Figure 5.
Summing up, the characteristics of this type of novel
compounds are: (i) a high monodispersiveness; (ii) for the
colloidal solution, the presence of a chemical stabilizer
during synthesis does not perturbate the chemical nature
of the novel compound, therefore, a similar stoichiometry (Rux Sey , in which x 2 and y 1) as revealed by
Rutherford back scattering (RBS) was obtained;[39] (iii) this
colloidal material is easy to deposit, e.g., by dipping, on

Chevrel phases and chalcogenides 539

(a)

50nm
50nm

Ru
C
Se
O
(c)
0.32 m
(b)

Figure 4. (a) TEM picture depicting the morphology of Rux Sey


material issued from the synthesis in xylene. (b) A typical SEM
picture of Rux Sey layers deposited onto a glassy carbon (GC)
substrate. (c) Molecular structure of Ru4 Se2 CO11 . This is the
chemical precursor formed at the beginning of the reaction
synthesis between Ru3 CO12 and dissolved selenium in xylene.

4 THE SELECTIVITY TOWARD


OXYGEN REDUCTION IN PRESENCE
OF METHANOL AND A COMPARISON
WITH PLATINUM

102

i (mA cm2)

Although the structure of Rux Sey has not yet been established, such local structure measurements (via EXAFS)
were very informative in order to assess the cluster-like and
the nanocrystalline nature of these compounds. This fact is
not recognizable using the X-ray diffraction technique due
to the broad diffraction peaks.[41a] These observations have
been unfortunately attributed to an amorphous state of the
material by some authors,[42] who use the same synthesis
procedure developed by us.
An EXAFS in situ electrochemical study on Rux Sey ,
however, showed that, on the one hand, the local structure
is distorted during electrocatalysis. On the other hand, such
distortion is accompanied by an increase of the disorder
due to the interaction of the molecular oxygen with the
catalytic center leading to an amorphous-like state. This
phenomenon is similar when naked ruthenium clusters
are exposed to oxygen from air.[13] These observations
strongly suggest that the intermediate surface complex
during electrocatalysis is based on an oxide-like species.

101

4.1 The selectivity of ruthenium-containing


Chevrel mixed phase
100

0.0

0.2

0.4

0.6

0.8

1.0

E (V(RHE))

Figure 5. (a) Mass transfer corrected catalytic current on applied


electrode potential of Rux Sey thin layers deposited onto GC
substrates for the ORR in 0.5 M H2 SO4 in absence () and
in presence of 0.1 M CH3 OH (). The mass transfer correction
was done using the relation [i = iL iobs /(iL iobs )], where i is the
kinetic or catalytic current, iL is the diffusion limiting current and
iobs the observed current from RDE measurements.

conducting substrates (e.g., glassy carbon, SnO2 :F); (iv)


they are selective for the ORR in the presence of methanol,
(see Section 4). EXAFS studies on this type of novel compounds have been also performed in ex situ and in situ
environments.[41] The average structural parameters: coordination distance, R, and coordination number (N ) for
rutheniumselenium and rutheniumruthenium are, respec (2.4) and 2.64 A
(3.4). The metalmetal coortively, 2.43 A
dination distance is of the same order of magnitude as that
of the metalmetal distance in the Chevrel phase clusters.

In an earlier publication,[16d] it was reported that the mixed


cluster compound, Mo4 Ru2 Se8 , sustained an electrocatalytic current for the ORR, without degradation, in the
presence of 13.5 M CH3 OH. The fabrication of a rudimentary liquid FC with a separator containing Pt/Sn as
an anode and Mo4 Ru2 Se8 as a cathode (both in massif
form) delivered an open circuit potential of 0.45 V and,
at short-circuit, a current of 170 A cm2 . With the present
knowledge that the electrocatalytic center of the mixed cluster phase lies on the ruthenium centers,[32] it was expected
that the novel ruthenium-based compound would possess
a similar electrochemical behavior, as verified by in situ
EXAFS electrochemical studies.

4.2 The selectivity for ORR of novel materials


Using the same line of reasoning, with the rutheniumcontaining Chevrel cluster compound, the nanostructured
cluster-like materials Rux Sey were also found to be highly
selective towards molecular oxygen reduction in the presence of methanol[43] (see Figure 5). This property is interesting in the development of selective cathodes for FCs

540 Part 5: The oxygen reduction/evolution reaction

105

i (A cm2)

(see Section 4.3). The reason for this is probably due to


the d-states of the ruthenium centers favorably involved for
water coordination leading to rutheniumwater complexes
formed at potentials more negative than on platinum. In
fact, the differential mass spectrometry revealed that platinum can maintain the electrochemical channel (or catalytic
sites) for ORR in the presence of methanol.[44] This channel is not perturbated by the process of methanol oxidation
(CO2 production). However, the cell potential decreases
because of these two competing reactions (mixed-potential
formation), and therefore this constitutes a drawback for
the platinum cathode when methanol crosses over the electrolyte membrane in a DMFC.
To test the tolerance to methanol on the novel materials, half-cell measurements were performed using the
gas-diffusion electrode (GDE). For the fabrication of GDE,
two approaches were used. Firstly, the synthesized powder catalyst (20% w/o) was dry-mixed with carbon (XC-72
Vulcan-loaded 30% PTFE), dried at 100 C resulting in
Rux Sey /C(mix). This latter was thereafter deposited by
rolling onto porous carbon paper (Toray). The catalyst loading was 1 mg cm2 and electrodes were cut into disks (1-cm
diameter). Second, at the beginning of the synthesis reaction, XC-72 Vulcan was added to the reactor vessel in order
to obtain a mixture of Rux Sey (18% w/o). The resulting
product was labeled Rux Sey /C.
Figure 6 depicts the half-cell ORR measurement in
a double jacket cell, in galvanostatic mode, using
Rux Sey /C(mix) in 2 M H2 SO4 at 40 C. Using the same
arrangement, Pt/C (30% w/o) from ETEK was also used to
prepare GDE (as described above) onto carbon paper with
the same amount of catalyst loading. A gentle flow of pure
oxygen was applied to the rear part of the electrode (porous
carbon paper side) at atmospheric pressure. Figure 6 also
depicts the currentpotential characteristic of Pt/C in the
absence and in presence of 1 M CH3 OH. It looks clear
that methanol depolarizes the Pt/C system. At a current
density of 105 A cm2 the potential is shifted 0.14 V in
the negative direction, whereas Rux Sey /C(mix) remains
selective, and thus more performant than Pt/C at lower
current densities.
In this half-cell experiment, Pt/C, in the presence of
methanol, reaches the same performance of Rux Sey /C(mix)
at 103 A cm2 . These curves were not corrected by an
ohmic drop. With respect to these results, it was evident
that an optimization procedure was necessary. In this
connection, the supported chalcogenide catalyst (Rux Sey /C)
was also measured, and the kinetic for ORR was done
using the RDE technique and again compared with Pt/C.
The electrodes were prepared as follows: an aliquot (from
a sonicated catalyst suspension with water) was deposited
onto glassy carbon disks in order to have an equivalent

104

103

102
0.5

Rux Sey /C(mix)

0.6

Pt/C

0.7

0.8

0.9

1.0

E (V(RHE))

Figure 6. The ORR on GDEs measured in a half cell at normal


pressure and at a temperature of 40 C. The comparison is made
between a Pt/C, ETEK (, ) and an electrode of loading
1 mg cm2 was prepared by mixing XC-72 Vulcan and Rux Sey
powder to a carbon paper (Toray) substrate (, ). The electrolyte
was 2 M H2 SO4 (, ) and H2 SO4 + 0.1 M CH3 OH (, ). As
observed, the response of Rux Sey was identical with or without
methanol. The hysteresis observed on each curve indicates that
the galvanostatic measurements were performed from lower to
higher currents and vice versa.

of 14 g cm2 of metal catalyst. To attach the catalyst


particles a diluted Nafion solution was deposited onto
the electrode surface. Figure 7 contrasts such measurement
at 25 C in oxygen-saturated 0.5 M H2 SO4 . The current
is reported in milliamps per milligram of metal catalysts.
However, as depicted in the TEM picture (Figure 7a) of
the chalcogenide supported catalyst (Rux Sey /C), it can be
seen that distribution of the catalyst agglomerates and the
particles is not uniform. Nevertheless, it appears that in this
condition the ORR follows the same kinetics.
The Tafel slope is similar (70 mV dec1 ). Moreover, the
currentpotential characteristics of Rux Sey /C are shifted
by 0.1 V more negatively than on Pt/C. One arrives at
the conclusion that there is still a margin to optimize the
distribution of the supported chalcogenide catalyst particles.

4.3 The performance of cluster-like materials as


cathodes in a DMFC
Rux Sey /C (cf. Figure 7a) was used for the fabrication of
gas-diffusion cathode electrodes. The GDE was prepared
by applying the catalyzed carbons (mixture containing the
supported chalcogenide, Rux Sey /C and Nafion solution) to
a diffusion layer onto carbon cloth (from ETEK ) to attain a
catalyst loading of 1 mg cm2 of Rux Sey /C (18% w/w). The
anode was a Pt/Ru/C (ETEK ) with a loading of 1 mg cm2
(30% w/w). For the sake of comparison in a DMFC, a

Chevrel phases and chalcogenides 541

Rux Sey

101

Carbon

i (mA mg1)

100

Pt/C

Rux Sey /C

101

70 mVdec1

102

103
0.6

100nm
(a)

0.7

0.8

0.9

1.0

E (V(RHE))

(b)

Figure 7. (a) TEM picture issued of an in situ synthesis of Rux Sey in presence of XC-72 Vulcan (Rux Sey /C). (b) Mass transfer corrected
catalytic current of Rux Sey /C in comparison to Pt/C from ETEK in oxygen saturated 0.5 M H2 SO4 electrolyte at 25 C. Each electrode
was prepared with 14 g metal (Pt or Ru) cm2 . The mass transfer correction was done using the relation [i = iL iobs /(iL iobs )], where
i is the kinetic or catalytic current, iL is the diffusion limiting current and iobs the observed current from RDE measurements.

selectivity (due to the presence of ruthenium atoms) for


the ORR of these catalysts allowed them to determine, in
real FC conditions, the steady state concentration of the
fuel.[45a]

0.7

40

0.6

35

25
0.4
20
0.3

15

0.2

10

0.1

0.0

50

100

150

200

250

P (mW cm2)

30

0.5

E (V)

cathode Pt/C (ETEK ) with an equivalent catalyst loading


was used, 1 mg cm2 (30% w/w). Membrane electrode
assembly (MEA) was done using a membrane layer of
Nafion 117. This MEA was hot pressed at 130 C over
90 s and at a pressure of 6.89 MPa.
The performance of both systems is compared in a single
DMFC. Figure 8 shows the current densitypotential and
the powerpotential curves of Rux Sey /C and Pt/CMEAs
at the temperature of 90 C. Such a performance can
be improved by increasing the temperature and catalyst
loadings. The following observations can be noted. First,
the chalcogenide material is selective for ORR in the
presence of methanol, as verified with half-cell measurements. Second, in spite of the high activity of the Pt/C
cathodes (cf. Figure 7) a similar electrochemical behavior is observed between Rux Sey /C and Pt/C in the FC
(Figure 8).
Because of the high tolerance of the chalcogenide material to methanol, this measurement proves that platinum is
depolarized by the methanol crossed over the electrolyte
Nafion 117 membrane. With respect to the in situ synthesis, the novel material is far from being optimized. Therefore, the perspectives for further improvement are open
as recently pointed out by the Newcastle group.[45] These
authors produced metal-centered catalysts (Mx Ruy Sz ), via
carbonyl compounds, using sulfur as a ligand. The high

0
300

j (mA cm 2)
Figure 8. The polarization and the power characteristics of
a liquid-feed DMFC operating at 90 C. , MEA of 20 wt%
Pt/C (1 mg Pt cm2 ) and 20 wt% Pt/Ru/C (1 mg Pt/Ru cm2 ); ,
MEA of 18.6 wt% Rux Sey /C (1 mg Rux Sey cm2 ) and 20 wt%
Pt/Ru/C (1 mg Pt/Ru cm2 ). For both systems PCH3 OH = 1.7 bar;
and PO2 = 1.5 bar.

542 Part 5: The oxygen reduction/evolution reaction

5 CONCLUSIONS
This chapter states the role of cluster-like materials from
micro- to nano-size in the reduction of molecular oxygen
in an acid medium. The Chevrel phase cluster compounds
were used as a starting model. These compounds served
as a basis to understand the complex interplay of the
electrocatalytic process of ORR. They also served as a
basis in understanding the concept of electron reservoirs
created in the cluster unit for electrocatalysis. All this
knowledge is conveyed at designing novel materials with
interesting physicalchemical properties vis-`a-vis the same
electrochemical process. One of these properties, necessary
for electrocatalysis, can be assigned to the ability of
the catalytic center to provide coordination chemistry for
reactivity and selectivity. Cluster-like materials seem to
fulfill this expectation. The evidence for this is provided
in half-cells and FCs using methanol as the fuel. In this
connection, in tailoring clusters, in a nanodivided form, one
seeks to mimic the way in which real clusters are arranged
in complex molecular structures. Therefore atoms should be
arranged in such a way as to fulfill the structure (electronic
and geometric) designed for a selective multielectron charge
transfer reaction.

ACKNOWLEDGEMENT
The author would like to thank his students and co-workers
of his former institute (Hahn-Meitner-Institut-Berlin) and
present laboratory.

REFERENCES
1. M. R. Tarasevich, A. Sadkowski and E. Yeager, Kinetics
and mechanism of electrode processes, in Comprehensive
treatise of electrochemistry, B. E. Conway, J. OM. Bockris,
E. Yeager, S. U. M. Kahn and R. E. White (Eds), Plenum
Press, New York, Vol. 7, p. 301 (1983).
2. (a) B. E. Conway and G. Jerkiewicz, J. Electroanal. Chem.,
339, 123 (1992); (b) V. Stamenkovic and N. A. Markovic,
Langmuir, 17, 2388 (2001); (c) N. A. Markovic, H. A.
Gasteiger and Ph. N. Ross, J. Phys. Chem., 100, 6715 (1996).
3. V. Climent, N. M. Markovic and P. N. Ross, J. Phys. Chem.
B, 104, 3116 (2000).
4. H. Naohara, S. Ye and K. Uosaki, Electrochim. Acta, 45,
3305 (2000).
5. (a) J. C. Huang, R. K. Sen and E. Yeager, J. Electrochem.
Soc., 126, 786 (1979); (b) K.-L. Hsueh, D.-T. Chin and
S. Srinivasan, J. Electroanal. Chem., 153, 79 (1983); (c)
K.-L. Hsueh, E. R. Gonzalez and S. Srinivasan, Electrochim.
Acta, 28, 691 (1983); (d) S.-M. Park, S. Ho, S. Aruliah,

M. F. Weber, W. A. Ward, R. D. Venter and S. Srinivasan,


J. Electrochem. Soc., 133, 1641 (1986); (e) S. J. Clouser,
J. C. Huang and E. Yeager, J. Appl. Electrochem., 23, 597
(1993); (f) S. M. Kaska, S. Sarangapani and J. Giner, J. Electrochem. Soc., 136, 75 (1989).
6. (a) M. T. Paffett, J. G. Beery and S. Gottesfeld, J. Electrochem. Soc., 135, 1431 (1988); (b) P. Stonehart, J. Appl.
Electrochem., 22, 995 (1992); (c) J. A. Poirier and G. E.
Stoner, J. Electrochem. Soc., 142, 1127 (1995); (d) S. Mukerjee, S. Srinivasan, M. P. Soriaga and J. McBreen, J. Electrochem. Soc., 142, 1409 (1995); (e) T. Toda, H. Igarashi and
M. Watanabe, J. Electroanal. Chem., 460, 258 (1999).
7. S. Mukerjee and J. McBreen, J. Electroanal. Chem., 448, 163
(1998).
8. J. A. Poirier and G. E. Stoner, J. Electrochem. Soc., 141, 425
(1994).
9. U. A. Paulus, T. J. Schmidt, H. A. Gasteiger and R. J. Behm,
J. Electroanal. Chem., 495, 134 (2001).
10. T. Okada, J. Dale, Y. Ayato, O. A. Asjornsen, M. Yuasa and
I. Sekine, Langmuir, 15, 8490 (1999).
11. C. Coutanceau, M. J. Croissant, T. Napporn and C. Lamy,
Electrochim. Acta, 46, 579 (2000).
12. J. Prakash and H. Joachim, Electrochim. Acta, 45, 2289
(2000).
13. V. le Rhun, E. Garnier, S. Pronier and N. Alonso-Vante,
Electrochem. Commun., 2, 475 (2000).
14. (a) F. C. Anson, C. Chi and B. Steigert, Acc. Chem. Res., 30,
437 (1997); (b) E. Song, C. Shi and F. C. Anson, Langmuir,
14, 4315 (1998); (c) B. Steiger and F. C. Anson, Inorg.
Chem., 39, 4579 (2000); (d) A. L. Bouwkamp-Wijnoltz,
W. Visscher, J. A. R. van Veen and S. C. Tang, Electrochim.
Acta, 45, 379 (1999); (e) P. Gouerec and M. Savy,
Electrochim. Acta, 44, 2653 (1999); (f) O. El. Mouahid,
C. Coutanceau, E. M. Belgsir, P. Crouigneau, J. M. Leger
and C. Lamy, J. Electroanal. Chem., 426, 117 (1997).
15. (a) A. L. Bouwkamp-Wijnoltz, W. Visscher and J. A. R. van
Veen, Electrochim. Acta, 43, 3141 (1998); (b) S. L. J.
Gojkovic, S. Gupta and R. F. Savinell, J. Electrochem.
Soc., 145, 3493 (1998); (c) S. L. J. Gojkovic, S. Gupta
and R. F. Savinell, J. Electroanal. Chem., 462, 63 (1999);
(d) S. L. J. Gojkovic, S. Gupta and R. F. Savinell, Electrochim. Acta, 45, 889 (1999); (e) S. Gupta, D. Tryk,
S. K. Zecevic, W. Alfred, D. Guo and R. F. Savinell, J. Appl.
Electrochem., 28, 673 (1998).
16. (a) N. Alonso-Vante and H. Tributsch, Nature (London),
323, 431 (1986); (b) N. Alonso-Vante, W. Jaegermann,
H. Tributsch, W. Honle and K. Yvon, J. Am. Chem. Soc., 109,
3251 (1987); (c) N. Alonso-Vante, B. Schubert, H. Tributsch
and A. Perrin, J. Catal., 112, 384 (1988); (d) N. AlonsoVante, B. Schubert and H. Tributsch, Mater. Chem. Phys., 22,
281 (1989); (e) C. Fischer, N. Alonso-Vante, S. Fiechter and
H. Tributsch, J. Appl. Electrochem., 25, 1004 (1995).
17. N. Alonso-Vante, J. Chim. Phys., 93, 702 (1996).
18. N. Alonso-Vante, German patent DE 196 44 628 C2, October
17, 1996.
19. R. Chevrel, M. Sergent and J. Pringent, J. Solid State Chem.,
3, 515 (1971).

Chevrel phases and chalcogenides 543


20. (a) . Fischer and M. B. Maple (Eds), Superconductivity of
Ternary Compounds, Springer-Verlag, Heidelberg (1982);
(b) R. Chevrel, M. Hirrien and M. Sergent, Polyhedron, 5,
87 (1986).
21. O. K. Andersen, W. Klose and H. Nohl, Phys. Rev. B, 17,
1209 (1978).
22. T. Hughbanks and R. Hoffmann, J. Am. Chem. Soc., 105,
1150 (1983).
23. (a) F. Kubel and K. Yvon, Acta Crystallogr., C43, 1655
(1987); (b) W. Honle, H. Flack and K. Yvon, J. Solid State
Chem., 49, 157 (1983); (c) J. Neuhausen, E. W. Finckh and
W. Tremel, Inorg. Chem., 35, 5622 (1996); (d) F. Le Berre,
C. Hamard, O. Pena and A. Wojakowski, Inorg. Chem., 39,
1100 (2000).
24. (a) W. Wakihara, T. Uchida, K. Suzuki and M. Taniguchi,
Electrochim. Acta, 34, 867 (1989); (b) S. Yamaguchi,
T. Uchida and M. Wakihira, J. Electrochem. Soc., 138, 687
(1991); (c) L. Guohua, H. Ikuta, T. Uchida and M. Wakihara,
J. Power Sources, 54, 519 (1995).
25. (a) K. F. McCarty and G. L. Schrader, Ind. Eng. Chem. Res.
Dev., 23, 519 (1984); (b) K. F. McCarty, J. W. Anderegg
and G. L. Schrader, J. Catal., 93, 375 (1985); (c) S. J.
Hilsenbeck, R. E. McCarley, R. K. Thompson, L. C. Flanagan and G. L. Schrader, J. Mol. Catal. A, 122, 13 (1997);
(d) J. W. Benson, G. L. Scrader and R. J. Angelici,
J. Mol. Catal. A, 96, 283 (1995); (e) V. Harel-Michaud,
G. Pesnel-Leroux, L. Burel, R. Chevrel, G. Geantet, M. Cattenot and M. Vrinat, J. Alloys Compd., 317318, 195
(2001).
26. K. Yvon, Curr. Topics Mater. Sci., 3, 54 (1978).
27. W. Jaegermann, C. Pettenkoffer, N. Alonso-Vante, Th.
Schwarzlose and H. Tributsch, Ber. Bunsenges. Phys. Chem.,
94, 513 (1990).
28. (a) L. Leduc, A. Perrin and M. Sergent, Acta Crystallogr.,
C39, 1503 (1983); (b) N. L. Speziali, H. Berger, G. Leicht,
R. Sanjines, G. Chapuis and F. Levy, Mater. Res. Bull., 23,
1597 (1988).
29. (a) C. Fischer, N. Alonso-Vante, S. Fiechter, H. Tributsch,
G. Reck and B. Schulz, J. Alloys Compd., 178, 305
(1992); (b) C. Fischer, S. Fiechter, H. Tributsch, G. Reck
and B. Schulz, Ber. Bunsenges. Phys. Chem., 96, 1653
(1992).
30. E. V. Anokhina, M. W. Essig, C. S. Day and A. Lachgar, J.
Am. Chem. Soc., 121, 6827 (1999).
31. R. C. Weast (Ed), CRC Handbook of Chemistry and
Physics, 65th edition, CRC Press, Boca Raton, FL,
1984.

32. N. Alonso-Vante, M. Fieber-Erdmann, H. Rossner, E. HolubKrappe, Ch. Giorgetti, A. Tadjeddine, E. Dartyge, A. Fontaine
and R. Frahm, J. Phys. IV, 7, C2887 (1997).
33. G. Chatzitheodorou, German patent DE 3802236 A1, January
22, 1988.
34. M. Mingo and D. Wales, Introduction to Cluster Chemistry,
Prentice Hall, Englewood Cliffs, NJ, 1990.
35. N. Alonso-Vante, M. Giersig and H. Tributsch, J. Electrochem. Soc., 138, 639 (1991).
36. V. Le Rhun and N. Alonso-Vante, J. New Mater. Electrochem. Syst., 2, 475 (2000).
37. V. Le Rhun, Ph.D. Dissertation, University of Poitiers, France
(2001).
38. T. M. Layer, J. Lewis, A. Martin, P. R. Raithby and W.-T.
Wong, J. Chem. Soc. Dalton Trans., 3411 (1992).
39. O. Solorza-Feria, K. Ellmer, M. Giersig and N. AlonsoVante, Electrochim. Acta, 39, 1647 (1994).
40. N. Alonso-Vante, H. Tributsch and O. Solorza-Feria, Electrochim. Acta, 40, 567 (1995).
41. (a) N. Alonso-Vante, M. Fieber-Erdmann, P. Borthen,
H. Strehblow and E. Holub-Krappe, Electrochim. Acta, 45,
4227 (2000); (b) N. Alonso-Vante, P. Borthen and M. FieberErdmann, in Proceedings of the 2nd International Symposium on New Materials for Fuel Cells and Modern Battery
Systems II, O. Savadogo and P. R. Roberge (Eds), Presented
at Editions de lEcole Polytechnique de Montreal, Montreal,
Canada, p. 654 (1997); (c) I. V. Malakov, S. Nikitenko,
E. Savinova, D. Kochubey and N. Alonso-Vante, J. Phys.
Chem. B, 106, 1670 (2002).
42. (a) S. D. Ramirez-Raya, O. Solorza-Feria, E. Ordonez-Regil,
M. Benaissa and S. M. Fernandez-Valverde, Nanostruct.
Mater., 10, 1337 (1998); (b) O. Solorza-Feria, S. CitalanCigarroa, R. Rivera-Noriega and S. M. Fernandez-Valverde,
Electrochem. Commun., 1, 585 (1999); (c) P. J. Sebastian,
Int. J. Hydrogen, 25, 255 (2000).
43. N. Alonso-Vante, in Proceedings of the 1st International Symposium of Fuel Cell Systems, O. Savadogo,
P. R. Roberge and T. N. Veziroglu (Eds), Presented at Editions de lEcole Polytechnique de Montreal, Montreal,
Canada, p. 658 (1995).
44. N. Alonso-Vante, P. Bogdanoff and H. Tributsch, J. Catal.,
190, 240 (2000).
45. (a) R. W. Reeve, P. A. Christensen, A. J. Dickinson,
A. Hamnett and K. Scott, Electrochim. Acta, 45, 4237
(2000); (b) R. W. Reeve, P. A. Christensen, A. Hamnett and
S. C. Roy, J. Electrochem. Soc., 145, 3463 (1998).

Anda mungkin juga menyukai