Está en la página 1de 34

INTERMETALLIC COMPOUNDS IN CATALYSIS

Introduction
Metal-based catalysts are used in many industrially relevant reactions, for example, hydrogenations or the synthesis of methanol (1). Often the catalysts comprise
a noble metal, and a second metal is added to improve the catalytic properties.
These bimetallic catalysts are dealt with in great detail in another article of this
Encyclopedia, showing the importance of these systems (2). An ideal bimetallic
catalyst consists of supported particles, in which the metals are well mixed and
adopt the crystal structure of one of the two constituent metals. From a metallurgical point of view, the particles form a solid solution or a substitutional alloy, in
which the crystallographic sites are occupied in a random manner. Here, the term
substitutional alloy specifies a subclass of the generally adopted meaning of the
term alloy in the catalysis community: . . . any metallic system containing two or
more components, irrespective of their intimacy of mixing or the precise manner
in which their atoms are disposed (3). These substitutional alloys are realized
by a wide range of different bimetallic transition metal mixtures, for example,
CuNi or AgPd. The formation of substitutional alloys allows tuning the electronic structure, approximated by the rigid-band model (4). Keeping the crystal
structure, while modifying the electronic structure, results in chemical and physical properties that lie between the properties of the participating elements. The
advantageous catalytic properties of such catalystsin comparison to elemental
catalystshas been realized since the 1960s (5).
Besides the bimetallic systems forming substitutional alloys, there are also
combinations that do not mix as well as systems in which the elements are forming chemical compounds comprising crystal structures that are more complex
than the structures of the constituent elements. Such an intermetallic compound
is a chemical compound consisting of two or more metallic elements, which exhibits a crystal structure different from the constituting elements. Another requirement is that the realized crystal structure is at least partly ordered. In addition, the term compound implies the single-phase nature of an intermetallic
compound. With this definition, the intermetallic compounds are another subclass of the alloys as defined above (6,7). As shown in the following, the changes
in the crystal structure results in strongly modified electronic structure and,
thus, in catalytic properties that are very different from the elements forming
the compoundjust as a hydrocarbon compound possesses properties that are
completely different from elemental carbon or hydrogen.
1
c John Wiley & Sons, Inc. All rights reserved.
Encyclopedia of Catalysis. Copyright 

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 1. Phase diagram of the system Al-Cu (211). Reprinted with permission of ASM
International . All rights reserved. www.asminternational.org.

Changes upon Compound Formation


Intermetallic compounds are usually not among the first choice of novel catalytic
materials. Besides being widely unknown, they are normally synthesized by a
heat treatment of the appropriate amounts of the metallic elements in inert atmosphere, resulting in limited specific surface areas. Only very recently nanostructured intermetallic compounds became available by industrially feasible routes,
making these interesting materials available to the catalytic community (8,9).
Since the formed intermetallic compounds are thermodynamically more stable than the elements, the chemical reaction is accompanied by an exothermic
reaction. This reaction results in (1) a change of crystal structure, (2) a modified
electronic structure, and (3) a transfer of charge. This influences the physical and
chemical properties drastically, and often covalent bonds are formed between the
atoms. An example to illustrate the changes upon mixing metals is the AlCu
system in Figure 1. Both elements crystallize in the Cu-type of structure (cubic closed packed). Aluminum can dissolve small amounts of Cu (denoted (Al) in
Fig. 1), keeping the close-packed atom arrangement, while Cu can take up 20
at% of aluminum, leading in both cases to the formation of substitutional alloys.
Within these regions, the rigid-band approach can be applied as an approximation, since there is no change in the crystal structure. As in the case of other

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 2. Crystal structure of CuAl2 . The interpenetrating 63 Al-nets (grey bonds) are interconnected by three centre Al-Cu-Al bonds (grey triangles). In addition, the Cu atoms
form chains parallel to the c-axis by homoatomic bonds (black). Three centre bonds are
only shown for the central Cu-chain.

substitutional alloys such as CuNi or AgPd, alloy formation allows a finetuning of the electronic structure by increasing or decreasing the number of electrons
per atom in the unit cell. As a result, the adsorption and thus the catalytic properties of alloys can be influenced.
At compositions between those of the substitutional alloys, a number of ordered intermetallic compounds with different crystal structures are formed between Al and Cu. While the substitutional alloys still hold four atoms in their
unit cell (Pearson symbol cF4), the intermetallic compounds realize more complex structures, holding between six atoms per unit cell in 2 -Cu1.5 Al (Ni2 In type
of structure, hP6 (10) and 156 atoms in the case of 1 -Cu33 Al17 (hR156 (11)). Out
of the numerous intermetallic compounds in this system, -CuAl2 has been investigated in depth by experimental as well as quantum chemical methods (12) and
serves here as an example to show the changes that take place upon compound
formation. In contrast to the substitutional alloys, CuAl2 realizes a tetragonal
crystal structure with 12 atoms per cell (CuAl2 type of structure, tI12) in which
each copper atom is surrounded by two Cu and eight Al atoms and each aluminum atom has four Cu and three Al neighbors (Fig. 2). As revealed by quantum chemical calculations, the aluminum atoms form three homoatomic bonds
and build covalently connected and interpenetrating 63 nets. The Cu atoms are located in tetragonal antiprismatic cavities, interconnect the 63 nets by eight threecenter Al Cu Al bonds and form Cu-chains along [001] by homoatomic bonds.
As a result of the realized crystal structure, the electronic structure is changed
dramatically and the resulting bonding situation determines the physical and

INTERMETALLIC COMPOUNDS IN CATALYSIS

chemical properties. In contrast to Cu and Al, CuAl2 is no longer ductile, but brittle as glassa property which is often observed for (partly) covalently bonded
intermetallic compounds. In addition, localization of the electrons in the covalent
bonds reduces the number of charge carriers and thus the electric resistivity is
significantly increased to 7.6 cm (13) at 295 K compared to 295K,Cu = 1.55
cm and 295K,Al = 2.73 cm. From the quantum chemical calculation follows
further a partially negatively charged Cu, as expected from the electronegativity
difference after Allred-Rochow. The strongly modified electronic structure of the
compounds, especially the shifting of the d-band with respect to the Fermi level,
is the most powerful method to tailor adsorption and reactivity of the surface.
As can be seen in Figure 1, intermetallic compounds often possess a homogeneity range in which their composition is changing, but not their crystal
structure. Homogeneity ranges represent another possibility to tune the properties of the compounds, similar to the formation of alloys. Often the electronic
structure can also be tuned by adding a third element while keeping the crystal
structure. This allowsas in the case of the bimetallic alloy catalyststo adjust
the adsorption and, thus, the catalytic properties. Since intermetallic compounds
possess crystal structures that are very different from the elemental structures
as well as strongly altered electronic structures, it is not possible to treat them
as intermediate between the constituting elements. They rather should be noticed as new elements concerning their potential in catalysis. For example, the
structural complexity of intermetallic compounds with motifs of transition metal
atoms or clusters surrounded by elements with strongly differing chemical properties allows in a unique way to realize geometric concepts, such as the isolation
of active sites (14). The strong localization of the electrons in the covalent bonds,
creates centers of strong Lewis basicity providing effective chemisorption of reactants without having to rely on metaloxygen entities, which are typical for
supported catalysts.
Besides being a driver for the realized and partly very complex structures
resulting in quasicrystalline intermetallic compounds in the extreme case (15)
the covalent bonding is very beneficial to the stability of the compounds under
catalytic reaction conditions. If the compounds are stable in situ, the observed
catalytic properties, such as activity, stability, and especially the selectivity, can
directly be connected to the crystal and electronic structure of the compounds.
This enables a knowledge-based development, which is much more efficient than
the widespread try-and-error exploration. However, special care has to be taken,
when intermetallic compounds are used in hydrogen containing or oxidizing atmosphere, for example, steam reforming of alcohols. Depending on the parameters, intermetallic compounds might be oxidized or form hydrides. An example
is the hydride-forming intermetallic compound La(Ni,Co)5 , which is the active
phase in nickelmetalhydride batteries. Upon hydride formation or any other
transformation, the crystal and electronic structure change dramatically and in
the case of the transformation of LaNi5 to LaNi5 H7 , the changes have been investigated in detail (16). Hydrogen uptake results in a doubling of the unit cell
as well as in a pronounced change in the chemical bonding from being metal-like
toward being more ionic after the hydrogen uptake. In addition, the electronic
density of states (DOS) changes significantly, on the one hand by increasing the
Fermi level in comparison to the unoccupied La states and on the other hand

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 3. Electronic DOS of LaNi5 (a) and LaNi5 H7 (b). The hydride formation changes the
electronic structure by opening a gap at around 4 eV as well as by reducing the distance
between the unoccupied La states and the Fermi-level. Reprinted from Sol. State Sci., 11,
A. F. Al Alam, S. F. Matar, M. Nakhl, N. Quaini, Investigations of Changes in Crystal
and Electronic Structures by Hydrogen within LaNi5 from First-Principles, 1098-1106,
Copyright 2009, with permission from Elsevier.

by opening a gap at around 4 eV (Fig. 3). As seen in the following section, hydride formation can either be beneficial-eg, for total hydrogenation reactions or
detrimental if the substrate should only be partially hydrogenated.

INTERMETALLIC COMPOUNDS IN CATALYSIS

Another benefit of the structural rigidity of many intermetallic compounds


is that it prevents the dissolution of reactants underneath the surface. This prevents the formation of subsurface compounds, such as hydrides and carbides,
which have been identified as being the catalytically active phase in elemental Pd hydrogenation catalysts as dealt within more detail below. Embedding
the centers of reactivity in a dense atomic matrix like in the intermetallic compounds provides mechanical and structural stability and excellent thermal properties preventing the subsurface chemistry.

Applications of Intermetallic Compounds in Catalysis


Because of the hurdles in the synthesis and their structural complexity, intermetallic compounds as catalysts may sound like an exotic applicationperhaps
with only academic interest. Intermetallic compounds do, however, offer a unique
combination of valuable properties that outweighs by far the difficulties in their
catalytic applicationa picture which is corroborated by a careful literature
search. Besides the well-known examples of the Raney-catalysts, there are about
1200 reports, where intermetallic compounds are involved in catalytic processes.
On the one hand, they are used as precursors for high surface area catalysts,
on the other hand they are commonly observed after catalytic processes where
hydrogen is involved as a result of reactions between the supported noble metal
and the partly reduced support. Also, very rarely, the catalytic properties of wellcharacterized intermetallic compounds have been investigated.
The earliest use of unsupported intermetallic compounds in catalysis was

reported by Rienacker
and co-workers in the 1930s, who applied different intermetallic compounds as catalysts in the dehydrogenation of formic acid (1719).
Since then intermetallic compounds played an important role in heterogeneous
catalysisoften without being recognized as such. Especially if reactions or pretreatments of supported metal catalysts are performed in reducing atmosphere
at high temperatures intermetallic compounds can be formed and change the
properties of the catalyst drastically. Unfortunately, according to the above definition of alloy, the metallic phases present on the catalysts are often referred to
as alloys, hindering the recognition of the involved intermetallic compounds.
The usage of intermetallic compounds can thus be divided into the use as
precursors, the (un)intended formation on the support, as well as their unsupported use. In each of the following sections, different approaches and reactions
are selected from the literature mirroring the different ideas and uses rather
than aiming at a full review.

Intermetallic Compounds as Precursor Materials.


Decomposition of Intermetallic Compounds. In 1925, Raney invented the
technique of decomposing intermetallic compounds such as NiAl, Ni2 Al3 , NiAl3 ,
Ni3 Al, or NiSi in a controlled manner (2022). By treatment of the conventionally
synthesized, that is from the elements by melting, and ground intermetallic compounds with NaOH solution, the silicon or aluminum is dissolved, leaving finely
divided elemental nickel behind. The high surface area and the easy and cheap
synthesis made the resulting catalyst industrially important and well known,
especially for the hardening of fats to produce, for example, margarine (23).

INTERMETALLIC COMPOUNDS IN CATALYSIS

Recently, these catalysts were also applied to produce hydrogen from methanol
(24). The same technique was used later to obtain other transition metals such
as Pt, Co, Cu, or Au with high specific surface areas. Hydrolysis of LiPt2 or NaPt2
leads to colloidal Pt-based hydrogenation catalysts, whereas leaching RECo5
(RE = rare earth element) with 1,2-diiodoethane leads to high surface area cobalt
catalysts (25,26). Cu-Raney catalysts for the synthesis of methanol from synthesis gas (CO2 , CO, and H2 in varying fractions) can be obtained by treating
AlCuZn alloys, containing the intermetallic compounds CuAl2 or Cu(Zn)Al2 ,
with NaOH solution (27). The discovery of gold as a heterogeneous catalyst by
Haruta and co-workers (28) also triggered the synthesis of a Raney-type Au catalyst, which was obtained by strong leaching of the intermetallic compound AuCu3
with 50% HNO3 aqueous solution (29). Generally, it is thought that size effects
can influence the catalytic activity of gold catalysts and that they are only active for oxidation reactions below a critical particle size (30). This is contradictory to the results obtained with the Raney-Au catalyst, which showed the same
specific surface activity as Harutas Au/TiO2 for CO oxidation despite being not
nanoparticulate. To increase the specific activity further, a promising strategy
is not to leach bulk compounds, but to start with nanoparticulate intermetallic
compounds. This has been demonstrated by Bonnemann and co-workers, who
synthesized nanoparticulate NiAl intermetallic compounds and then leached
them with NaOH (31). The resulting catalysts showed the same leaching behavior
as conventionally prepared Raney catalysts, but possessed a two to three times
higher activity for the hydrogenation of butyronitrile.
Besides hydrogenation reactions, Raney-type catalysts derived from intermetallic compounds can also be used in electrocatalysis as hydrogen evolution
electrodes. Ni-based electrodes were produced by leaching NiAl3 with an aqueous NaOH solution (32). An improvement in the electrodes in form of a lower
overvoltage could be achieved by not leaching NiAl3 but (Nix Moy Tiz )Al3 , which in
addition showed a higher corrosion resistance (33).
Leaving intermetallic compounds with translational periodicity also structurally very complex quasicrystalline compounds have been investigated as
suitable starting materials for the synthesis of Raney-type catalysts. Tsai and
co-workers investigated the leaching behavior of a series of quasicrystalline compounds in the AlCuFe system (3437). After wet milling in ethanol, the compounds were leached with Na2 CO3 , leading to elemental Cu-particles supported
on the remaining unleached quasicrystalline core. The obtained Cu-based catalysts were tested in methanol steam reforming, which resulted in higher activity
than for conventionally synthesized Raney catalysts. This is thought to be due to
either residual iron, which stabilizes the small Cu-particles, or the contact to the
quasicrystalline surface, which is acting as support. Usually oxides are used as
supports and since they are electrically insulating, the band structure of the supported particles is not influenced. In the case of an underlying conducting material, the Cu particles can be influenced by the band structure of the support, thus
changing the catalytic properties. Since methanol steam reforming might become
an important building block for a hydrogen-based energy infrastructure in the
future, the role of intermetallic compounds in this reaction has been reviewed in
depth very recently (38).

INTERMETALLIC COMPOUNDS IN CATALYSIS

Another elegant approach is to synthesize the active species as well as the


support in one step from a single intermetallic compound by controlled oxidation
or total oxidation and subsequent controlled reduction. This approach was first
pursued by Wallace in 1976-1977 by accident. He was investigating intermetallic
compounds with a composition of RETM 5 and RE2 TM 17 (RE = Th, Sm, Gd, La,
Er, or Ce; TM = Ni, Fe or Co), which are known to form hydrides (7). The idea
was to use the activated hydrogen atoms in the hydride to perform hydrogenation
reactions. Hydridic hydrogen atoms are usually highly active (eg, in PdHx (39),
since the hydrogenhydrogen bond is already broken. In the first publication, the
formation of hydrocarbons from H2 /CO mixtures by the FischerTropsch process
was investigated (40). The intermetallic compounds were introduced as fine powder in the reactor and treated with hydrogen. Interestingly, the compounds activated with time on stream, and a first idea was an increase in active surface area
under reaction conditions. Shortly after this publication, the real cause for the
activation was revealed (41). The intermetallic compounds not only formed hydrides, but were partly oxidized by the oxygen, which is contained in the CO. This
resulted in the formation of very active, finely divided Ni, Co, or Fe particles supported on a rare earth oxide. Depending on the transition metal used, the main
FischerTropsch product was different. Nickel-based catalysts produced mainly
methane, whereas copper-based catalysts led to the formation of methanol.
Simultaneously, the synthesis of ammonia from the elements was also tested
over RE-TM intermetallic compounds by the same group (42). Here, 36 intermetallic compounds of rare earth elements and the transition metals Fe, Co, and
Ru were evaluated. In the case of the ammonia synthesis, the rare earth component is transformed to the corresponding nitride and the transition metal is finely
divided thereon. Some of the compounds showed an even higher specific activity
than commercial catalysts used at that time. Later, the catalysts based on the
intermetallic compound CeNi5x Cox (x = 05) were oxidized in a controlled way
prior to the catalytic characterization in the FischerTropsch reaction (43). The
reasons to oxidize the compound before use were twofold. On the one hand, the
material can be handled in air afterwards, whereas the intermetallic compound
itself is not stable in air. Second, the oxidation can be better controlled and reproduced if it is not performed in situ. A review on the early work on these catalysts
is available by Wallace (44).
Besides the synthesis of hydrocarbon and ammonia, this type of catalysts
has also been applied by different groups for the synthesis of methanol (4548).
Recently, catalysts derived from RE-TM intermetallic compounds had a revival
for the oxidative dehydrogenation of 2-propanol to acetone and propene (49).
Here, as well as in earlier work, the oxidation and reduction behavior of the intermetallic compounds and their decomposition products is studied in detail to
optimize the synthesis procedure (49).
Intermetallic compounds also play a role in the development of photocatalysts for water splitting. Using TiSi2 as a photocatalyst, water can be split in
hydrogen and oxygen (50). During the reaction with water, TiO2 and SiO2 are
formed on the surface by oxidation. This leads to a material that can split the
water using light and also stores the oxygen. As a further advantage, the latter is
only released when the catalyst is heated to 100 C, thus enabling the separation
of the catalytically generated H2 and O2 . Up to now, it is not clarified how the

INTERMETALLIC COMPOUNDS IN CATALYSIS

at least three components in the catalyst interact and which of them are needed
for water splitting. TiSi2 itself might play a part in the reaction since strong deactivation is observed when the reaction is performed in the gas phase under
elevated pressure (51).
Formation of Hydrides. The hydrogen storage compounds of the type RETM already mentioned can also be used in a different way as catalysts. Instead of
oxidizing or nitriding the rare earth component, the hydride itself can be used for
hydrogenation reactions as aimed for by W. E. Wallace in the first place. To avoid
the oxidation of the intermetallic compound, any oxygen has to be avoided, which
even in the form of CO or CO2 can oxidize the RE-TM intermetallic compounds.
To form the catalytically active hydride, either the compounds are treated with
hydrogen before use, or the hydride is formed in situ. The resulting hydrides
are very active for hydrogenation of unsaturated hydrocarbon compounds. As the
most active element for hydrogenations, that is palladium, which can also form
a very active hydride, the unsaturated hydrocarbons are transformed to their
fully saturated counterparts. As a consequence, these compounds are not suited
for semihydrogenations, for example, of C C triple bonds to double bonds. On
the other hand, the highly active hydrides allow the hydrogenation of very stable hydrocarbons, for example, benzene, or the hydrogenolysis of hydrocarbons.
First reports of the use of the hydrides of intermetallic compounds based on Ti
and Ni for the transformation of 1-hexene to n- and iso-hexane, as well as the
hydrogenolysis of n-heptane and n-hexane go back to 1975 (52,53). Within the
next 15 years, different groups demonstrated the use of the hydrides of LaNi5
(54,55), LaCo5 (55,56), CeCo5 (55,57), PrCo5 (55), and SmCo5 (55) as catalysts
in a series of hydrogenation and hydrogenolyses reactions. In parallel, the suitability of hydrides based on intermetallic compounds without rare earth metal,
such as ZrNi (58,59), ZrCo (60), ZrFe (61), and TiCu (61) in these reactions, was
demonstrated successfully. Besides the hydrogenation of unsaturated C C bonds,
this type of catalyst proved to be useful for the synthesis of alcohols from aldehydes (62) and the synthesis of aniline from nitrobenzene (63,64). A series of
compounds, based on intermetallic compounds of a transition metal and a main
group metal, have been successfully tested as hydrogenation catalysts recently.
Mg2 CoH5 , Mg2 NiH4 , Mg2 FeH6 , and MgCuH2 are able to hydrogenate ethylene,
n-butenes, as well as 1,3-butadiene (65). Also, this class of compounds can activate C O bonds, enabling the synthesis of 2-propanol by hydrogenation of acetone (66). Besides hydrogenation reactions, Oesterreicher and Spada also studied
the formation of water from hydrogen and oxygen. The reaction can already be
catalyzed at room temperature if the hydride of LaNi5 is used as catalyst (67).
Furthermore, the electrocatalytic properties of the hydrides were explored. Besides hydrogen production by water splitting on RETM5 electrodes (68) also more
complex reactions like the reaction of water and nitrobenzene to phenylhydroxylamine can be catalyzed by ZrNi electrodes (69). More examples are collected in
a recent review (70).
Another very interesting approach to use the hydride formation ability of
the rare earth based intermetallic compounds was demonstrated by using them
as hydrogen absorbers in dehydrogenation reactions. In this way, the intermetallic compound is transformed into a stable hydride during the reaction. Propane
can be cracked and dehydrogenated into methane and carbon over LaNi5 and

10

INTERMETALLIC COMPOUNDS IN CATALYSIS

LaCo5 (71). However, dehydrogenation can be diminished if alcohols are used


as reactants. The dehydrogenation of 2-propanol over Nd2 Co7 or Sm2 Co7 yields
acetone (72,73), whereas methanol and ethanol result in the formation of carbon monoxide and acetaldehyde, respectively (73). In both cases, ErFe2 or DyFe2
can also be used as catalysts, whereas in the case of methanol Zr2 Ni has been
proven to be active after a pretreatment with HF, NaOH, and hydrogen (74). By
the pretreatment, Zr2 Ni is partially destroyed, leading to the formation of small
Ni-clusters and fluorides. During the last part of the pretreatment, the remaining
Zr2 Ni is transformed into its hydride. It was shown that the nickel clusters are
vital for the catalytic process and that the removal of the stored hydride as well
as oxidation of the nickel clusters leads to a loss of catalytic activity. Since all
the above-mentioned dehydrogenation reactions were performed in batch operation, the amount of hydrogen to be stored as hydride was limited. Unfortunately,
no data are available on the behavior of these systems if the intermetallic compounds are completely transformed into their hydrides, thus are saturated with
hydrogen. Also, the intermetallic compounds in the above reactions are not used
as catalysts in the classical sense, since they are transformed into the hydride
and the material as well as its properties will change during the reaction.
The materials mentioned in this section have in common that they were
derived from intermetallic compounds, which acted as precursors for the active
species. The selective etching of the compounds leads to Raney-type catalysts,
whereas the controlled oxidation results in catalysts that consist of supported
transition metals on oxides. Both systems have the common advantage that the
transition metal of the intermetallic compound can be obtained in a finely divided
state, resulting in a high specific activity. However, the catalytic properties such
as selectivity and deactivation behavior often resemble to those of the underlying
transition elements. A different approach is the use of intermetallic compounds
with the ability to form hydrides. But also in this case, the catalytically active
species is not the intermetallic compound, but its hydride, which possesses different properties, for example, the highly active hydridic hydrogen atoms as well as
a different crystal and electronic structure.

(Un)intended Formation of Supported Intermetallic Compounds.


By far the most common way intermetallic compounds take part in catalysis is
by being formed during synthesis or under reaction conditions. There are two
ways the intermetallic compounds can be formed on conventionally supported
catalysts. On the one hand, this can be initiated by the reaction of the supporting material with the catalytically active transition metal; on the other hand,
the presence of more than one metal on the support can lead to the formation
of intermetallic compounds. While these processes can cause the unintended formation of intermetallic compounds, two methods for the targeted synthesis of
supported intermetallic compounds were also developed. The first is the so-called
controlled surface reaction (CSR), comprising the reaction of the supported metal
with a highly reactive gas-phase species, for example, organometallic compounds
(75,76) or Zn-vapor (77). The second one is an industrially feasible synthesis of
single-phase supported intermetallic compounds by careful decomposition of a
single-phase hydrotalcite material (8,9).
Conventionally, catalysts are often synthesized by impregnation methods.
After impregnation, drying, and calcination, the material is reduced in hydrogen

INTERMETALLIC COMPOUNDS IN CATALYSIS

11

to obtain the active transition metal in a finely divided form. If the transition
metal, now in its elemental state, is able to activate hydrogen, for example, by
adsorption or by hydride formation, intermetallic compounds can be formed by
partial reduction of the support in reducing atmospherethe so-called reactive
metalsupport interaction (RMSI) (38). Supports, which can be reduced and are
quite often used in heterogeneous catalysis, comprise CeO2 , Ga2 O3 , GeO2 , ZnO,
and SnO. On SiO2 , Al2 O3 , or metal free supports like carbon or boron nitride,
the formation of intermetallic species is not to be expected unless the temperatures during reduction are unusually high (78,79) or a second metal is present
as a promoter. The formation of intermetallic compounds is not restricted to the
synthesis of the catalysts. Also under reductive reaction conditions, intermetallic compounds can be formed and are then observed after the reaction. Usually,
drastic changes in the catalytic properties result from the formation of the intermetallic compounds, for example, a pronounced change in selectivity or activity.
Sometimes, the term strong metal support interaction (SMSI) is used for the
phenomenon of intermetallic compound formation, which is not correct, since an
SMSI effect must be reversible by definition (8082). This is not true for the intermetallic compounds formed, since they usually stay intact during the reaction
and a strongly oxidizing treatment is necessary to decompose them.
By far the most reports deal with the partial reduction of ZnO, used as support, and subsequent formation of Zn-containing intermetallic compounds. In the
case of Pd/ZnO, a catalyst used for the decomposition (83), the partial oxidation
(8489) and the steam reforming of methanol (9094), the interaction between
the supported Pd particles and the partial reduced support can lead to a mixture
of ZnPd, Zn3 Pd2 , and ZnPd2 . When this catalyst is used for the hydrogenation of
unsaturated hydrocarbons, such as butenes, isoprene, or crotonaldehyde, only the
formation of ZnPd is observed (9597). The same holds for reductive treatments
of Pd/ZnO at temperatures above 400 C (Fig. 4). The supported catalyst Pd/ZnO
is not the usual case, since normally a mixture of intermetallic compounds and
the originally supported transition metal is observed.
PtZn is another intermetallic compound frequently observed after synthesis or reaction. The underlying catalyst, that is Pt/ZnO, has been shown to catalyze the FischerTropsch reaction (98), the decomposition of methanol (83) as
well as the water gas shift reaction (99). In the case of Pt/CeO2 and Pt/GeO2 ,
used as catalysts for the hydrogenation of ,-unsaturated aldehydes, the formation of CePt5 (100) and PtGe as well as PtGe2 (101), respectively, has been
reported, leading to a higher selectivity towards the unsaturated alcohol. Applying Ni/CeO2 for the hydrogenation of benzene results in the occurrence of the
intermetallic compounds CeNi and CeNi2 (102), whereas the decomposition of
methanol over Cu/ZnO can result in the formation of Cu0.61 Zn0.39 (103). Interestingly, even in quite oxidizing environments the intermetallic compounds can play
a role in catalysis. This has been demonstrated with the use of Pt/SnO2 , on which
PtSn2 and PtSn4 are formed during the synthesis, as a CO-resistant catalyst for
fuel cells where hydrogen and oxygen are reacted to water (104).
Another way is the formation of intermetallic compounds by the so-called
promotion or modification of supported catalysts. The promoting metal, very
often a main group metal, can either be introduced by coimpregnation or by subsequent impregnation of the support. The intermetallic compounds observed in

12

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 4. XRD of 10 wt.-% Pd/ZnO during thermal treatment in reducing atmosphere. The
formation of the intermetallic compound ZnPd can clearly be seen and is not complete
until temperatures of more than 400 C are reached. At this temperature, significant sintering starts as revealed by the decreasing FWHM. Reprinted from Appl. Catal. A, 125, N.
Iwasa, S. Masuda, N. Ogawa, N. Takezawa, Steam Reforming of Methanol over Pd/ZnO:
Effects of the Formation of PdZn Alloys Upon the Reaction, 145-157, Copyright 1995, with
permission from Elsevier.

these catalysts result from the interaction of the catalytically active metal and
the promoting metal, usually during the reduction step of the synthesis or under reaction conditions. This effect was first reported by Swift and Bozik in 1968
when they tested tin-promoted Ni/SiO2 catalysts for the dehydrogenation of cyclohexanone, cyclohexanol, and cyclohexane and observed the formation of the
intermetallic compound Ni3 Sn2 after the reduction of the catalyst (105). Other
binary systems for de-/hydrogenation reactions of different substrates, which are
prone to the formation of intermetallic compounds, are listed in Table 1. Other

INTERMETALLIC COMPOUNDS IN CATALYSIS

13

Table 1. Intermetallic Compounds Observed in De-/Hydrogenation Reactions on


Promoted Catalysts
Catalyst

Intermetallic
Compounds

PtCr/C
PtV/C
PtZr/C
PtSn/Al2 O3
PtSn/ZnAl2 O4
PtSn/ZnAl2 O3
PtMg/ZnAl2 O3
PtSn/Al2 O3
NiSn/SiO2

PtCr3
PtV, PtV3
Pt11 Zr9
PtSn
PtSn, Pt3 Sn, PtSn2
PtZn
PtZn
PtSn, Pt3 Sn
NiSn, Ni3 Sn4 ,
Ni3 Sn2

PdSn/SiO2

Pd3 Sn2 , PdSn,


PdSn2 , PdSn3
Pd3 Pb
Pd3 Sn, Pd2 Sn

PdPb/SiO2
PdSn/Al2 O3

Reaction

References

Hydrogenation of phenol

(106)

Dehydrogenation of octane
Dehydrogenation of n-butane
Dehydrogenation of n-butane

(107)
(108)
(109)

Isomerisation of n-hexane
Dehydrogenation of cyclohexanone,
cyclohexylamine, cyclohexane,
2-propanol, and ethylene

(110)
(111,112)

Hydrogenation of acetylene
Hydrogenation of hexa-1,3-diene
and hexa-1,5-diene

(113)
(114,115)

examples, where intermetallic compounds are observed in hydrogenation reactions are the synthesis of ammonia over Fe-Co/MgCO3 Mg(OH)2 3 H2 O, where
the formation of Fe3 Co is increasing the activity (116) or the FischerTropsch reaction over Pt-Sn/Al2 O3 where PtSn and PtSn4 are formed during the reduction
of the catalyst (117).
A reaction in which promoting metals are used frequently is the hydrogenation of ,-unsaturated aldehydes. The reaction toward the unsaturated alcohols,
which are valuable intermediates in the pharmaceutical and fragrance industries, is thermodynamically not favored (118). Thus, the promoting concepts aim
at the preferential adsorption of the C O bond to preferentially hydrogenate the
C O bond. Intermetallic compounds are increasing the selectivity, most likely
by providing metal atoms with different Lewis-acidity on their surface due to
the electron transfer between the atoms mentioned earlier. Catalysts tested in
this reaction are PtFe/BN (formation of PtFe observed (119), PdSn/SiO2 and
PdSn/Al2 O3 (PdSn (120)), RhSn/SiO2 , and RuSn/Al2 O3 (RhSn2 , Rh3 Sn2 , and
Ru3 Sn7 (121,122)) as well as AgIn/SiO2 (Ag3 In (123,124)). Especially, the latter
provide selectivities and activities of commercial catalysts without a basic additive and can be used without solvent in the liquid phase.
Also under oxidizing conditions, like the partial oxidation of methanol to hydrogen, water, and CO2 , intermetallic compounds can be formed. As in the case
of methanol steam reforming over Pd/ZnO, the intermetallic compound ZnPd is
part of the catalytic system PdZn/Al2 O3 under these conditions (125). The observation of intermetallic compounds is not limited to oxidation reactions involving methanol. Also in the oxidation of aldoses such as lactose and glucose over

14

INTERMETALLIC COMPOUNDS IN CATALYSIS

promoted Pd-catalysts, intermetallic compounds are present (PdTl is formed on


PdTl/SiO2 (126); PdBi and PdBi2 are observed in the case of PdBi/SiO2 (127
129)). Intermetallic compounds can survive even stronger oxidizing reaction conditions as seen in the oxidation of hydrocarbons, CO and NO/CO mixtures. Here,
the intermetallic compounds CeRh and Pd0.48 In0.52 are detected after use of Ce
Rh/SiO2 (130) and PdIn/SiO2 (131), respectively, as catalysts in these reactions.
From all these observations, the question arises how the intermetallic compounds are formed by the reactive interaction between the supported metal and
the support. To answer this question, a series of model systems has been investigated by the groups of Hayek and Klotzer. The synthesis of the model systems
starts with a NaCl single crystal on which the transition metal is deposited in
ultrahigh vacuum (UHV) in the form of small particles. Subsequently, the supporting material of the catalytic system to study is deposited as a layer covering
the transition metal particles. An additional SiO2 layer stabilizes the arrangement. The NaCl is dissolved in water, leaving behind material, which represents
a model of a supported catalyst. The advantage of this thin-film sample is that
changes occurring to it under reducing conditions can be observed by TEM. Using this approach, the systems Pt/Al2 O3 (formation of Pt3 Al and PtAl (79), (132)),
Pt/CeO2 (Pt3 Ce (79,133)), Pt/SiO2 (Pt3 Si, PtSi, and Pt12 Si5 (79)), Pd/SnO2 (Pd2 Sn
and PdSn (134)), Pd/GeO2 (Pd2 Ge (134)), Pd/In2 O3 (PdIn (135)), Pd/ZnO (ZnPd
(136)), Pd/Al2 O3 (Pd4 Al3 and PdAl (137)), and Pd/Ga2 O3 (Pd5 Ga2 (138)) have
been investigated. While on easy-to-reduce supports like ZnO compound formation was observed at relatively mild conditions, hard-to-reduce supports (SiO2 ,
Al2 O3 , or Ga2 O3 ) needed very harsh conditions before intermetallic compounds
were formed. In many of these studies, the corresponding impregnated catalysts
were also investigated in comparison and this demonstrated impressively, that
with the exception of a few systems, it is very hard to obtain single-phase and
supported intermetallic compounds solely by reduction of impregnated catalysts.
For this reason two different approaches were developed for a targeted synthesis of supported intermetallic compounds. The first approach is CSR, which
was invented in the 1980s to synthesize supported Rh-alloys with Ge, Sn, and Pb
(75,76). Starting from a supported and reduced transition metal catalyst synthesized by conventional impregnation techniques, the transition metal is exposed to a highly reactive organometallic species at elevated temperatures either in a solvent or in the gas phase. The decomposition of the organometallic
species is mediated by the transition metal, leading to selective deposition on the
transition metal particles. Since the first report of this technique, a number of
supported intermetallic compounds were synthesized, based on platinum (PtZr,
Pt11 Zr9 , PtV, PtV3 , Cr3 Pt, Pt2 Ge, Pt3 Ge2 , PtGe, Pt2 Ge3 , PtGe2 (106,139)), nickel
(Ni3 Sn, Ni3 Sn2 , Ni3 Sn4 , Ni3 Ge (140142)) as well as RuTi (143). A disadvantage
of this approach is the use of very toxic and hard to handle chemicals. In addition, more than one intermetallic compound can be present in the final catalyst.
For intermetallic compounds consisting of one metal with a high vapor pressure
at moderate temperatures, for example, Zn, the above synthesis method can be
modified by using the metal vapor directly as a reactant, thus circumventing toxicity and handling problems. This method has only been developed very recently to
successfully transform the system Pt/C to PtZn/C (77), and future developments

INTERMETALLIC COMPOUNDS IN CATALYSIS

15

will show whether it is also possible to synthesize other intermetallic compounds


using this elegant technique.
Even with this progress, the challenge in the intended synthesis of supported and single-phase intermetallic compounds is to develop a synthesis which
is industrially feasible. This restricts not only the costs but also the toxicity and
the difficulty of handling the materials. To achieve single-phase compounds, the
use of a homogeneous single-phase precursor in which both of the metals are
distributed evenly is a good starting point. Starting from these requirements,
a hydrotalcite-based synthesis for the intermetallic compound Pd2 Ga was developed (8,9). The single-phase hydrotalcite precursor is synthesized by aqueous coprecipitation of Pd2+ , Mg2+ , and Ga3+ . Here, Mg2+ and Ga3+ form the
hydrotalcite-like lattice, whereas Pd2+ substitutes for a small fraction of Mg2+
in the layers. The precursor is then dried and optionally calcined. Subsequent
reduction leads to intermetallic Pd2 Ga particles, which are supported on a mixture of MgO and MgGa2 O4 . During the reduction, first elemental Pd particles are
formed, which then mediate the reduction of the Ga3+ in their vicinity by spillover
of the activated hydrogen. The different reduction potentials of Ga3+ and Mg2+
enable to reduce only the Ga to the elemental state and thus only to form the
GaPd intermetallic compound.
Two challenges arise, if one would like to conclude about the catalytic properties of the intermetallic compounds in the large variety of reactions mentioned
in this section: First of all, since in most of the systems not only one (or more)
intermetallic compounds are present but also the original transition metal, it is
hard to extract any information about the catalytic properties of the involved intermetallic compounds. Usually this is even more complicated, since the promoting metal is not only present in the intermetallic compound but also in different
oxidation states, thus modifying the support and/or the transition metal in one
way or the other. If we assume that the intermetallic compounds are catalytically
active in one way or the other, the problem is the mixture of different, potentially active, intermetallic compounds as well as the high number of different
interfaces, which are present on these catalysts. This hinders the assignment of
compound-specific catalytic properties and thus a knowledge-based development.
Unsupported Intermetallic Compounds in Catalysis. The use of unsupported and in situ stable intermetallic compounds as catalysts allows studying the intrinsic catalytic properties of the compounds rather than the properties
of a complex system. While the full characterization of unsupported intermetallic compounds, especially concerning the phase and chemical composition of the
material, is straightforward, this is a major task for the commonly used impregnated catalysts. Here, the synthesis normally does not result in the presence of
one intermetallic compound, but in a variety of different compounds. For example, the impregnation of a Pd/SiO2 catalyst with a Ga(acac)3 solution, followed
by reduction in hydrogen at high temperatures, results in a mixture of Pd and
the Pd-rich intermetallic compounds (144) in the GaPd binary phase diagram
(Fig. 5). Owing to the presence of a combination of different Pd-containing compounds, which are all potentially active in the hydrogenation, an assignment of
the observed catalytic properties to one compound is not possible.
The idea to use unsupported intermetallic compounds as catalysts is
not new, and the first to explore the catalytic properties of unsupported

16

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 5. The Ga-Pd phase diagram (212). Reprinted with permission of ASM
International . All rights reserved. www.asminternational.org.

intermetallic compounds was Rienacker


in 1934 (17). Cu-based intermetallic compounds with Ni, Pd, Au and Pt were used as catalysts in the decomposition of
formic acid to study the influence of temperature induced orderdisorder transitions, that is the transition of an intermetallic compound to the corresponding
alloy (18,19,145,146). As a result, the ordering resulted in a higher activity and
a lower activation energy of the reaction, viz an improvement in the catalytic
properties. Another group that studied the decomposition of formic acid over a
variety of alloys and intermetallic compounds was Schwab (147). They detected a
sharp maximum in the activation energy, whenever the intermetallic compound
belonged to the group of HumeRothery compound with -brass structure, that is
an electron precise compound (see the next section). These examples demonstrate
the significance of the definition given in the Introduction, because the differences
in catalytic properties are a direct result of the difference of the geometric and
electronic structure between the intermetallic compounds and the alloyseven if

they have the same composition as in the case of the studies of Rienacker.
Besides
formic acid, the decomposition of methanol and hydrogen peroxide also has been
studied over Ni3 Al (148) and GaSb or InSb (149), respectively.
A number of unsupported intermetallic compounds have been tested as hydrogenation catalysts. The hydrogenation of 1,3-butadiene can be catalyzed by
Pt3 Ge, Pt2 Ge, and PtGe (150). The low activity of the compounds compared to
elemental Pt was assigned to the low hydrogen dissociation activity, whereas

INTERMETALLIC COMPOUNDS IN CATALYSIS

17

the high selectivity to butenes could be attributed to the different adsorption


behavior. A strong influence on the catalytic properties by the compound formation was also observed in the hydrogenation of ethylene to ethane over Pt3 Sn,
PtSn, and PtSn2 (151). The selective hydrogenation of acetylene to ethylene has
been performed over CoGe, CoGe2 , and CoSn (152) as well as over Ni3 Sn, Ni3 Sn2 ,
and Ni3 Sn4 (153). In both cases, the high selectivity for the semihydrogenation
was assigned to the strongly changed adsorption behaviors compared to the corresponding elements. The reaction was also studied using GaPd intermetallic
compounds following the site-isolation concept. Since these studies include a full
characterization under reaction conditions, this case study is dealt with in the
next section. That the use of intermetallic compounds must not necessarily result in improved catalytic properties was shown by applying ThNi2 and UNi2 as
catalysts for the hydrogenation of isoprene (154). The aimed for unsaturated isopentanes were only formed with selectivities of 80% over ThNi2 and 50% over
UNi2 both significantly lower than elemental Ni, which shows a selectivity of
96%. These experimental findings are very valuable, since the catalytic properties can be assigned to the crystal and electronic structures of the intermetallic
compounds applied and, thus, enlarge our knowledge on structureproperty relationships in this reaction.
A recent development is the application of Zintl compounds to increase the
selectivity toward the unsaturated alcohol in the selective hydrogenation of citral,
an ,-unsaturated aldehyde. Guided by the idea that a polar surface should lead
to a preferential adsorption of the polar C O bond over the unpolar C C bond
and thus to a higher selectivity toward the unsaturated alcohol the intermetallic
compounds Mg2 Sn (155), MgCo4 Ge6 (156), MgCo6 Ge6 (157), CaNi4 Sn2 , SrNi4 Sn2 ,
Ca0.5 Sr0.5 Ni4 Sn2 , Ni3 Sn, and Ni3 Sn2 (158) have been tested. Zintl compounds are
intermetallic compounds combining covalent and ionic interactions and where
the anions or the anionic network can formally be considered to be valence satisfied (159). Since the charge transfer is very pronounced in these compounds,
the surface atoms should be strongly polarized. Starting from Ni3 Sn and Ni3 Sn2 ,
the addition of Ca or Sr leads to a higher polarization in CaNi4 Sn2 , SrNi4 Sn2 ,
and Ca0.5 Sr0.5 Ni4 Sn2 due to the higher electronegativity difference. Despite providing a higher polarity, the selectivity of the latter compounds toward geraniol
and nerol did not improve. From these findings, two possible scenarios can be derived. Either polarity is not the only factor that influences the selectivity, or the
intermetallic compounds were not stable under reaction conditions.
A series of Cu-based intermetallic compounds (Cu2 Mg, Cu9 Ga4 , Cu3 In,
Cu6 La, Cu6 Pr, Cu3 Si, Cu4 Ti, Cu3 Ti, Cu9 Zr, Cu3 Zr, Cu3 Ge, Cu31 Sn8 , Cu1.99 Se,
and Cu2 Te (160) were tested as catalysts for the oxidative dehydrogenation of
methanol to formaldehyde and water. While Cu3 Si and Cu31 Sn8 showed the highest yield of formaldehyde, Cu6 La and Cu6 Pr led to CO and CO2 and the other
compounds showed no activity. While the materials were characterized well before the reaction, no explanation was offered concerning the different catalytic
behavior of the compounds.
As mentioned above, several intermetallic compounds can be formed if
Pd/ZnO is applied as catalyst for the steam reforming of methanol and is reduced
at too low temperatures. ZnPd (161) and Zn6.1 Pd3.9 (162) have also been tested
in an unsupported state in this reaction. The tests resulted in rather poor CO2

18

INTERMETALLIC COMPOUNDS IN CATALYSIS

selectivities of less than 90%, which according to the XRD patterns can be attributed to traces of elemental Pd still present in the catalysts. Single-phase ZnPd
was also tested in an unsupported state by Tsai and co-workers, now reaching
97% CO2 selectivity (163). This comparison reveals how important the influence
of a second catalytically active phase can be. While ZnPd is a highly selective
catalyst toward CO2 in methanol steam reforming, elemental Pd catalyzes the
decomposition of methanol to CO and H2 even in the presence of water (164). Obviously, traces of elemental Pd in the catalytic system are sufficient to lower the
selectivity significantly.
Intermetallic compounds often show metallic conductivity enabling their
use in electrocatalysis. In the early 1960s, first studies on the water-splitting
properties of NiSi, NiAs, NiSb, and NiTe2 were performed and showed that the
obtained reaction rates for the hydrogen evolution of the intermetallic materials
were always larger than the ones of the constituting elements (165). The intermetallic compounds Ti2 Ni and TiNi were applied as cathodes in this reaction
with the aim to store the evolved hydrogen (166). While the catalytic activity
was lower than for a Raney cathode, they are able to store larger amounts of
hydrogen, enabling to combine the high specific capacity of an accumulator with
the high specific power of a fuel cell. The early work on this subject has been
reviewed and reevaluated by Brooman and Kuhn (167). With the observation
that the electrocatalytic activity parallels the changes in the electronic properties, they arrive at the conclusion that the changes in the electronic properties
result in modified adsorption behavior and thus in altered electrocatalytic properties. However, the subject was pursued by other groups, confirming the above
statements for transition-metal/transition-metal intermetallic compounds in the
hydrogen (HER) as well as the oxygen evolution reactions (168). In 1984, Jaksic
correlated the catalytic properties in the HER of intermetallic compounds consisting of two transition metals (169). From a large amount of data, he concluded that
the synergetic effects observed for intermetallic compounds consisting of transition metals from the right-hand side as well as from the left-hand side of the
d-block are due to the strong chemical bonding in the compounds as proposed
by Brewer and Engel. The electronic structure results in two kinds of electrons,
that is paired ones, responsible for the chemical bonding, as well as unpaired
ones, which dominate the electrocatalytic properties. In addition, studies were
performed on intermetallic compounds in the systems NiSn (170), CoSi (171),
and FeSi (172).
Besides for the water-splitting reaction, intermetallic compounds have also
been applied as catalysts for the oxidation of potential fuelslike methanol or
formic acidin direct fuel cells. The focus of these investigations was mainly
on Pt-based intermetallic compounds with main group elements (Bi (173,174),
In (174), Pb (175) (174), Sn (174,176,177), and Sb (174,178)), but some Pt compounds with Cd (179), Hg (179), Ru (180), or Ti (181) as well as Pd-based systems (174) have also been investigated. In the oxidation of such compounds like
methanol, CO is a by-product and also a strong poison for the fuel cell catalyst.
The advantage of most of the investigated compounds is their significantly higher
CO tolerance compared to Pt. Among the best of the compounds are PtBi and
PtPb, which have been investigated in detail concerning changes during surface

INTERMETALLIC COMPOUNDS IN CATALYSIS

19

treatments (182) and have also been synthesized in a nanoparticulate state to


increase the activity (183).
Despite the large number of publications, where intermetallic compounds in
a pure and unsupported form have been used as catalysts, the stability under reaction conditions was not investigated or established in any of these studies. How
the combination of unsupported intermetallic compounds as catalysts and thorough studies on their stability under reaction conditions enables a knowledgebased development is shown in the next section.

Chances for a Knowledge-Based Development


The use of unsupported intermetallic compounds allows simplifying catalytic
systems drasticallyif they are stable under reaction conditions. Having only
one catalytically active phase present enables the assignment of the observed
catalytic properties to the crystal and electronic structure of the intermetallic
compound. The absence of the support also means that the number of different
interfaceswhich could play a role in catalyzing the reactionis reduced. Since
this very clean approach comes along with the drawback of very small specific
surface areas (typically around 0.1 m2 /g), it is especially suited for basic studies.
Having developed an understanding of the underlying processes on such unsupported materials enables the identification of completely new catalysts that
outperform conventional catalysts in terms of selectivity, activity, and stability
by the gained knowledge. This can be achieved by testing isostructural compounds with different electronic structures to study the electronic influence as
well as testing compounds with similar electronic structures to reveal geometric influences. By this, a screening or trial-and-error approach can be circumvented, which also would be very challenging due to the synthesis hurdles posed
by the intermetallic compounds. In the next step, the best performing intermetallic compounds can be synthesized in a nanoparticulate way to study size effects.
With these materials, it is also possible to investigate the influence of different
supports on the catalytic performance by depositing the nanoparticles after synthesis on the support. Finally, the focus can be directed to the synthesis of highly
dispersed intermetallic systems in an industrially feasible way. The advantage
of this knowledge-based approach is that at any time of the development process
the catalytic properties of the unsupported intermetallic compounds can serve as
a benchmark, guiding the successful synthesis of the nanostructured catalysts.
In contrast to alloys, intermetallic compounds are well suited for such an approach. On the one hand, there is a pool of more than 10,000 binary compounds
(and much more ternary and higher ones) available with a broad range of crystal
and electronic structures. On the other hand, they very often comprise covalent
bonding, increasing their stability and, thus, reducing segregation. In addition,
the simplicity of the systems (concerning their phase composition, not necessarily their crystal structure) allows studying the structural integrity under reaction conditions which is a necessary requirement to set up structureproperty
relationships.
The described approach has been applied in the development of a new class
of highly selective semihydrogenation catalysts based on GaPd intermetallic

20

INTERMETALLIC COMPOUNDS IN CATALYSIS

compounds. These compounds represent the first example, where the knowledgebased approach was applied starting from the bulk materials (6,184,185), studying their stability under reaction conditions (186,187), developing an approach
to unsupported nanoparticulate materials (8) and finally inventing an industrially feasible route to highly active materials while preserving the selectivity and
stability of the bulk compounds (8,9).
Investigating the Stability In Situ . Heterogeneous catalysis takes place
on the surface of the intermetallic compounds. But the bulk is also influencing
the properties of the surface, which has been shown recently in the case of Pdcatalyzed hydrogenations (see below). So the stability under reaction conditions
has to be proven for the surface as well as for the bulk.
Generally, the use of finely powdered intermetallic compounds is advantageous for the bulk methods. On the one hand, the larger surface area (typically
0.1 m2 /g, 20 m particle diameter) results in sufficient activity. On the other
hand, the particles are usually small enough that changes, starting from the surface, can be detected. The good crystallinity of the intermetallic compounds and
their unsupported nature allows detecting changes in the bulk by temperatureprogrammed XRD in reactive atmosphere. Ideally, the reactive gas is sucked
through the powdered sample to ensure (1) the contact with the reaction atmosphere and (2) to be able to detect the products for example, by GC analysis. In
contrast to supported catalysts, this results in XRD patterns, which are not dominated by the support signals (Fig. 6). Instead, changes in the material can be
clearly detected by changes in the lattice parameters (eg, hydride formation), a
decrease in signal intensity (eg, formation of amorphous phases), or the appearance of additional compounds, for example, oxides in oxidizing environments.
To complement powder X-ray diffraction, which is only able to detect
changes in crystalline materials, in situ extended X-ray absorption fine structure (EXAFS) measurements are useful (188). Contrary to XRD, EXAFS detects
the local environment surrounding an element. Thus, changes in coordination of
the neighborhood, for example, hydride-formation, which often results in changes
in the distances between the elements, can be detected. But also the formation
of amorphous species is detectable. Since the method is element sensitive, any
compound of one element will contribute to the signal. As a consequence, a mixture of different compounds of one element, as they are present on conventionally
prepared bimetallic catalysts, will result in complicated spectra, which are circumventing the detection of small changes and are not easy to interpret. Here,
the use of a single-phase intermetallic compound becomes immediately obvious.
A widespread property of intermetallic compounds is the ability to form hydrides (7). The formation of a hydride involves changes in the structural and
electronic properties. And, in general, hydrides are very active hydrogenation
catalysts as was shown above. As a result, the observed catalytic properties are
not specific for the intermetallic compound, which was placed in the reactor, but
belong to the hydride formed in situ. The whole situation becomes even more
complex due to the instability of some hydrides under normal conditions. This
results in the decomposition of the hydride as soon as the reactive atmosphere
is removed, which makes the ex situ detection difficult or even impossible. Thus,
to be able to assign the observed catalytic properties to a specific intermetallic
compound, the possibility of hydride formation under reaction conditions has to

INTERMETALLIC COMPOUNDS IN CATALYSIS

21

Fig. 6. Typical powder XRD patterns of a) 5% Pd/Al2 O3 and b) unsupported PdGa


recorded with the same parameters (Cu K1 radiation).

be excluded. Especially if the hydride formation is not complete and different


phases are present, the detection of the small changes by XRD or EXAFS is not
easy. A method that has been proven very powerful, and has only recently been
extended to in situ investigations of reactions in the gas phase, is prompt gamma
activation analysis (189). The measurement has to be corrected for background
contributions (mainly water in the path of the neutron beam) and results in a
very accurate measurement of the hydrogen concentration in the sample. As result, hydride formation can be detected at very low concentration levels, allowing
to exclude or to proof the presence of a hydride under reaction conditions. In contrast to elemental Pd, where the formation of the -hydride phase PdH0.6 is detected, no significant hydrogen uptake was observed for the GaPd intermetallic
compounds (187).

22

INTERMETALLIC COMPOUNDS IN CATALYSIS

Heterogeneous catalysis takes place at the surface of the intermetallic compounds, so additionally surface sensitive methods have to be applied under reaction conditions to ensure the stability in situ. A method that lies in between
bulk and surface sensitive measurements is DTA/TG in different atmospheres.
Here, weight changes or thermal signals give hints to changes in the material.
Owing to the new generation of instruments, being extremely robust against all
kinds of gases, the atmosphere for the measurements can be varied widely. Besides gaseous compounds, even vapors of liquids (eg, water/methanol mixtures)
can be used to test the stability of intermetallic compounds. Before measuring
in reactive atmospheres, a measurement in inert atmosphere is necessary to account for changes in the material which are not due to the reactive atmosphere
applied (structural phase transitions, the loss of one component at elevated temperatures, and so on). The strength of the method lies not only in the simultaneous detection of thermal and weight effects, but also in the small amount
(50 mg) of the sample necessary.
Changes in the so-called near surface region can be detected by high pressure in situ X-ray photoelectron spectroscopy (XPS) (190). Complementary to the
bulk methods XRD or EXAFS, XPS does not give any structural information but
is restricted to the electronic state of the elements within the material. To detect changes under reaction conditions, the different XPS signals of the elements
within the material are first recorded under UHV conditions. In addition to the
spectra of the metallic elements, also the spectra of carbon and oxygen need to
be recorded to detect changes in these signals under reaction conditions. Using
synchrotron radiation allows recording depth profiles and the detection of the
subsurface chemistry, which can strongly influence the catalytic properties. A
recent example is the detection of a subsurface palladiumcarbon phase, when
elemental palladium is used as a catalyst for the hydrogenation of alkynes (191
193). Subsurface phases strongly alter the electronic properties of the surface in
a feedback loop with the chemical potential of the reactants controlling the reaction properties of the catalyst. As soon as palladium is exposed to a hydrogencontaining atmosphere, -PdH0.6 is formed. The highly active hydride leads to
total hydrogenation of the respective alkynes. Carbon, resulting from the decomposition of the alkynes, diffuses into the palladium and forms a subsurface Pd/C
phase. The Pd/C acts as a diffusion barrier for the hydridic hydrogen, which now
cannot diffuse to the surface anymore. This results in a complete change from
the unwanted total hydrogenation of the alkyne to a selective semihydrogenation
to the corresponding alkene. After recording the UHV spectra and collecting the
information for different depths, in situ measurements can be performed. The reactive atmosphere is introduced into the UHV chamber at pressures of around 1
mbar, and then the sample is heated to the desired temperature. Mass spectrometry is applied to detect catalytic activity, and spectra are recorded after equilibration of the sample. Differences in the UHV spectra indicate alterations in the
compounds, like decomposition or the involvement of subsurface species. Besides
additional signals for an element, which result from the presence of an additional
compound, a shift of the signals is also possible. Owing to the formation of the intermetallic compound, the signals of the involved elements are already altered
compared to the elements and shifts of more than 1 eV are possible depending on
the kind and content of the second metal (Fig. 7).

INTERMETALLIC COMPOUNDS IN CATALYSIS

23

Fig. 7. Pd3d5/2 XPS signals of Pd in comparison to the intermetallic compounds Pd2 Ga,
PdGa and Pd3 Ga7 at a photon energy of 720 eV.

Having excluded changes in the near surface region or any subsurface chemistry, the surface itself also needs to be closely investigated to exclude restructuring, segregation, or the formation of agglomerates. A method that collects
structural as well as electronic information from the outmost atomic layer of the
intermetallic compound is infrared spectroscopy on adsorbed probe molecules like
carbon monoxide. CO is an excellent probe for conventional catalysts, especially
Pd supported on metal oxides. A vast collection of data exists, which allows attributing changes in the CO infrared spectra to either electronic or structural
features of the catalyst. In Figure 8, typical IR-spectra of Pd/Al2 O3 and PdGa
are opposed. Here, bands at high wave numbers are due to the adsorption of CO
on different Al2 O3 adsorption sites. Other bands can be assigned to CO on top a
Pd atom, or bridged between two and three neighboring Pd-atoms, respectively.
Changing the partial pressure of CO leads to a frequency shift of the CO bands
because of dipoledipole effects of neighboring CO molecules. This is a typical
behavior of extended active sites present on Pd/Al2 O3 , where the CO molecules
are adsorbed close enough to each other to get altered by dipole interactions
(194,195). The spectra of CO on the unsupported intermetallic compound PdGa
are shown on the right side of Figure 8. The much simpler spectra is in part the
result of the absence of any support. According to the literature, the only observed
band can be assigned to CO on top of the Pd atoms and the strong shift compared
to Pd indicates a higher electron density on the Pd in agreement with quantum
chemical calculations (6,187). The absence of bands, which can be assigned to
bridged CO molecules in the spectra, is a direct proof for only isolated atoms on
the surface of PdGa. Furthermore, no dipoledipole interactions are present, indicating that the CO molecules are adsorbed on Pd atoms which are further away
from each other than in elemental Pd. Both observations are in full accordance
with the isolated Pd atoms in the PdGa bulk structure (see the next section).

24

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 8. FT-IR spectra of a) 5% Pd/Al2 O3 and b) unsupported PdGa. Pd/Al2 O3 was subjected to 1 mbar of CO at 77 K and then evacuated (a-f) with subsequent heating to 124 K
(g-i). For PdGa a pressure of 50 mbar was necessarry at 293 K to observe a signal (a-d: 0
to 60 min).

The different in situ methods mentioned above are not comprehensive, and
other methods might also be suitable to determine the structural stability of
the intermetallic compounds. However, the set of measurements presented here
is sufficient to ensure the stability of the bulk and the surface of the intermetallic compounds under reaction conditions, forming the starting point for the
knowledge-based development of the catalysts.
Geometric and Electronic Concepts. Increasing selectivity is very often observed, when intermetallic compounds are formed on a conventionally supported catalysteither by reaction with promoting metals or the partially reduced support. Despite this interesting fact, the underlying reasons for the higher
selectivity are usually not addressed and are usually ascribed to the partial intermetallic compound formation or the promoting effect of the second metal instead
of considering explicitly the structural and electronic changes connected with the
formation of the corresponding intermetallic compound. Furthermore, the presence of more than one active species as well as the support hinders a full characterization of these complex systems. To exclude these complications, studies on
unsupported and single-phase intermetallic compounds are suited.
Considering only the structural situation within the used intermetallic compounds is connected with one big challenge. Structural changes always involve
changes in the electronic structure since crystal and electronic structure are not
independent properties of the corresponding intermetallic compound. Ideally, the
catalytic properties of compounds with different geometric motifs but alike electronic structures are compared to identify structural influences. Since it is not
known a priori, which part of the electronic structure of the intermetallic compounds interacts with the reactants this is not an easy task. On the other hand,
testing for electronic influences is straightforward, since here structural differences can be mostly excluded by employing isostructural compounds. But here

INTERMETALLIC COMPOUNDS IN CATALYSIS

25

also, small variations in the atomic distances, which might influence the adsorption properties, still persist.
In the following approaches for a knowledge-based development guided by
geometric or electronic concepts are summarized. Intermetallic compounds can
realize a wide variety of different surroundings for the transition metal involved,
thus allowing testing specific structural motifs for their catalytic properties. In
contrast to the substitutional alloys, the different crystallographic positions in
the structure of intermetallic compounds are not occupied statistically by the different metals, but in an ordered way. Segregation, commonly a problem in alloys,
is often drastically reduced due to the covalent bonding present in the intermetallic compounds. The resulting structural stability can preserve the geometric arrangement under reaction conditions.
The aforementioned GaPd intermetallic compounds represent an unconventional approach to increase the selectivity in the semihydrogenation of
alkynes by isolating the active sites from one another, the so-called active-site
isolation concept, introduced by Sinfelt, Sachtler, and Ponec in the 1970s (14). For
the semihydrogenation of acetylene small active sites are beneficial (196199). In
contrast, large active sites lead to ethane as well as carbonaceous deposits, which
poison the commonly used supported AgPd or AuPd alloy catalysts (200). Applying this site-isolation concept as a guide, the intermetallic GaPd compounds
could be identified as potential catalysts. In contrast to substitutional alloys, the
active-site isolation has not to be generated but is an intrinsic property of the
crystal structures of the intermetallic compounds. While in elemental Pd, each
Pd atom is surrounded by 12 Pd neighbors, this coordination is reduced statistically by substituting part of the Pd by Ga. According to the phase diagram in
Figure 5, Pd can take up approximately 5 at% of Ga without changing the crystal structure (Pd95Ga5). In contrast, the intermetallic compounds Pd2 Ga (201)
(Co2 Si type), PdGa (202) (FeSi type), and Pd3 Ga7 (203) (Ir3 Ge7 ) show drastically
reduced PdPd coordination numbers of 8, 1, and 0, respectively. The stability of
the atomic arrangement in the compounds under reaction conditions leads to the
excellent selectivities compared to Pd/Al2 O3 or the Pd95Ga5 substitutional alloy
(Fig. 9).
The first who studied the electronic influences of the intermetallic compounds was Schwab, who studied the decomposition of formic acid to CO2 and H2 .
Having studied alloys and intermetallic compounds in the binary systems AgAl
(204), AgPb (147), AgSb (147,205), AuCd (206), and CuSn (147,205), it could
be shown that the activation energy within the homogeneity range of an alloy
increased with increasing electron concentration. Comparing alloys and intermetallic compounds within these systems, the activation energy always showed a
maximum for the HumeRothery -phase, whereas the activation energies of the
alloys with maximum electron concentration did not vary significantly (Fig. 10).
The maximum in activation energy of the gamma compound was explained by the
nearly perfect filling of the Fermi sphere with electronsa phenomenon which is
connected to the optimized covalent bonding situation in the gamma compounds.
As a result, the catalytic properties are correlated to the electric conductivity as
well as to the mechanical properties of the compounds (207).
Another example is the aforementioned methanol steam reforming, which is
usually carried out over Cu/ZnO/Al2 O3 catalysts (208). Drawbacks of this system

26

INTERMETALLIC COMPOUNDS IN CATALYSIS

Fig. 9. Catalytic properties of the unsupported intermetallic compounds Pd2 Ga, PdGa
and Pd3 Ga7 in comparison to a commercial 5% Pd/Al2 O3 catalyst and the substitutional
alloy Pd95Ga5. Conversion (a) and selectivity to ethylene (b) are given for catalytic data
obtained in a plug-flow reactor at isothermal conditions (200 C) with a total flow of
30 mL/min and a feed composition of 0.5% C2 H2 , 5% H2 , 50% C2 H4 in helium. The pictograms in (b) show the first coordinaton sphere in the corresponding intermetallic compounds (Pd: red, Ga: blue), the alloy Pd95Ga5 as well as elemental Pd.

comprise is the low stability against sintering of the Cu particles, which leads to
deactivation or the pyrophoric nature of the materials. Iwasa and co-workers introduced ZnPd/ZnO, which behaves very similar to Cu-based catalysts, but shows
higher stability (93). By comparison of the electronic structures of ZnPd and elemental Cu, which is one of the most selective catalysts for the steam reforming of
methanol, Tsai and co-workers revealed a high similarity (163). They stated that
the electronic structure is responsible for the observed high selectivity of ZnPd.
To prove their statement, they tested the intermetallic compound PdCd having
a very similar electronic structure as ZnPd. PdCd showed a high CO2 selectivity
of 93%. NiZn and PtZn, on the other hand, possess different electronic structures
to ZnPd, and only showed limited selectivities of 8% and 40%, respectively. This
was the first time that the electronic structure of intermetallic compounds could
directly be correlated to their catalytic properties, enabling the prediction of the
catalytic performance of PdCd in this reaction.
As shown above, intermetallic compounds have also been used as catalysts
for the selective hydrogenation of ,-unsaturated aldehydes to the unsaturated
alcohols. Here, a large amount of work in conventional catalyst development

INTERMETALLIC COMPOUNDS IN CATALYSIS

27

Fig. 10. Activation energy of the decomposition of formic acid on Cu-Sn substitutional
alloys as well as Cu-Sn intermetallic compounds. For different systems a strong increase
in the activation energy wass observed for the electron-precise Hume-Rothery compounds
(138). Reproduced by permission of The Royal Society of Chemistry http://www.rsc.org.

is dedicated to increase the polarity of the surface to increase the adsorption


strength, and thus the hydrogenation probability, of the C O function (209). Up
to know, the beneficial influence of a highly polarized surface could not be proven,
because the generated data are often obscured by support influences and the presence of different catalytically active species. Nevertheless, the concept of polarization has attracted researchers to identify the reason for increased selectivity to

the unsaturated alcohol by using well-defined intermetallic compounds. Fassler,


Claus and co-workers (155158) tested a number of Zintl compounds as catalysts
in the reaction as mentioned above. The compounds differed in polarity, but no
correlation was found between polarity and the selectivity. That polarity somehow plays a role in the reaction that could be shown by Galloway and co-workers
by investigating the intermetallic compound PtZn as catalyst for the gas-phase
hydrogenation of crotonaldehyde (210). While unsupported PtZn shows only a selectivity of 37% toward crotyl alcohol, the selectivity could be increased to 88%
by treating the intermetallic compound with bromoethane. As shown by XPS, the
treatment led to the formation of an additional, positively charged Zn species increasing the polarity of the surface. As a result of these studies, polarity seems to
play a role, but cannot be the only important factor.

Conclusions
As shown in this article, intermetallic compounds are not rare species in catalysis. They are quite often formed on conventionally supported catalysts by subjecting them to reducing conditionseither during the preparation or under reaction conditionsthus being present in a broad range of reactions and catalytic
systems. Complementarily, a number of methods are known for the targeted

28

INTERMETALLIC COMPOUNDS IN CATALYSIS

synthesis of intermetallic compounds. They cover the full range of materials, that
is unsupported, supported, as well as nanostructured systems.
Intermetallic compounds realize a huge number of different crystal and electronic structures, the latter often providing stability by covalent bonding. Correlating the catalytic properties with the crystal and electronic structures of in situ
stable intermetallic compounds enables the knowledge-based development of new
catalytic systems, circumventing the wide spread trial-and-error approach. Setting up generic case studies as described above is essential to develop strategies
toward a broader application and exploitation of the potential intermetallic compounds present in catalysis.

BIBLIOGRAPHY
1. G. Centi, in B. Cornils, W. A. Herrmann, R. Schlogl, and C.-H. Wong, eds., Catalysis from A to Z, Wiley-VCH Verlag GmbH & Co. KgaA, Weinheim, Germany, 2003,
pp 490491.
2. J. H. Sinfelt, in I. T. Horvath, ed., Encyclopedia of Catalysis. Wiley-VCH Verlag
GmbH & Co. KgaA, Weinheim, 2003, pp 620665.
3. V. Ponec and G. C. Bond, Stud. Surf. Sci. Catal. 95, 299391 (1995).
4. N. F. Mott, Adv. Phys. 13, 325422 (1964).
5. J. H. Sinfelt, Adv. Chem. Eng. 5, 3774 (1964).

6. K. Kovnir, M. Armbruster,
D. Teschner, T. V. Venkov, F. C. Jentoft, Knop-A. Gericke,
Y. Grin, and R. Schlogl, Sci. Technol. Adv. Mater. 8, 420427 (2007).
7. H. Kohlmann, in Encyclopedia of Physical Science and Technology, Academic Press,
San Diego, Calif., 2002, pp 441458.

8. M. Armbruster,
K. Kovnir, M. Behrens, D. Teschner, Y. Grin, and R. Schlogl, J. Am.
Chem. Soc. 132, 1474514747 (2010).

9. A. Ota, M. Armbruster,
M. Behrens, D. Rosenthal, M. Friedrich, I. Kasatkin, F. Girgsdies, W. Zhang, R. Wagner, and R. Schlogl, J. Phys. Chem. C 115, 13681374 (2011).
10. M. El-Boragy, R. Szepan, and K. Schubert, J. Less-Common Met. 29, 133140
(1972).
11. E. H. Kisi and J. D. Browne, Acta Crystallogr., Sect. B: Struct. Sci. 47, 835843 (1991).

12. Y. Grin, F. R. Wagner, M. Armbruster,


M. Kohout, A. Leithe-Jasper, U. Schwarz, U.
Wedig, and H. G. von Schnering, J. Sol. State Chem. 179, 17071719 (2006).
13. C. Macchioni, J. A. Rayne, S. Sen, and C. L. Bauer, Thin Solid Films 81, 7178 (1981).
14. W. M. H. Sachtler, Catal. Rev. Sci. Eng. 14, 193210 (1976).
15. W. Steurer, Z. Kristallogr. 219, 391446 (2004).
16. A. F. Al Alam, S. F. Matar, M. Nakhl, and N. Quaini, Solid State Sci. 11, 10981106
(2009).

17. G. Rienacker,
Z. Elektrochem. 40, 487488 (1934).

18. G. Rienacker,
Z. Anorg. Allg. Chem. 227, 353375 (1936).

G. Wessing, and G. Trautmann, Z. Anorg. Allg. Chem. 236, 252262


19. G. Rienacker,
(1938).
20. U.S. Pat. 1563587 (1925), M. Raney.
21. US Pat. 1628190 (1927), M. Raney.
22. R. Sassoulas and Y. Trambouze, Bull. Soc. Chim. France 985988 (1964).
23. B. Kammermeier in Ullmanns Encyclopedia of Industrial Chemistry, Wiley-VCH
Verlag GmbH & Co. KgaA, Weinheim, Germany, 2008.
24. Y. Xu, S. Kameoka, K. Kishida, M. Demura, A. P. Tsai, and T. Hirano, Intermetallics
13, 151155 (2005).

INTERMETALLIC COMPOUNDS IN CATALYSIS

29

25. C. P. Nash, F. M. Boyden, and L. D. Whittig, J. Am. Chem. Soc. 82, 62036204 (1960).
26. H. Imamura, Y. Kato, and S. Tsuchiya, Z. Phys. Chem., Neue Folge 141, 129132
(1984).
27. J. B. Friedrich, M. S. Wainwright, and D. J. Young, J. Catal. 80, 113 (1983).
28. M. Haruta, N. Yamada, T. Kobayashi, and S. Iijima, J. Catal. 115, 301309 (1989).
29. S. Kameoka and A. P. Tsai, Catal. Lett. 121, 337341 (2008).
30. M. Turner, V. Golovko, O. P. H. Vaughan, P. Abdulkin, A. Berenguer-Murcia, M.
Tikhov, B. F. G. Johnson, and R. M. Lambert, Nature 454, 981983 (2008).
31. R. Richards, G. Geibel, W. Hofstadt, and H. Bonnemann, Appl. Organomet. Chem. 16,
377383 (2002).
32. O. K. Davtyan and R. I. Makordei, Soviet Electrochem. 7, 18161816 (1971).
33. U.S. Pat. 4374712 (1983), T. J. Gray (Olin Corporation, New Haven, Conn.).
34. S. Kameoka, T. Tanabe, and A. P. Tsai, Catal. Today 9395, 2326 (2004).
35. A. P. Tsai and T. Yoshimura, Appl. Catal. A 214, 237241 (2001).
36. T. Tanabe, S. Kameoka, and A. P. Tsai, Catal. Today 111, 153157 (2006).
37. M. Yoshimura and A. P. Tsai, J. Alloys Compd. 342, 451454 (2002).

38. M. Behrens, M. Armbruster


in L. Guczi, ed., Catalysis for Alternative Energy Generation, Springer, Berlin, Heidelberg, New York.
39. A. Borodzinski and A. Janko, React. Kinet. Catal. Lett. 7, 163170 (1977).
40. V. T. Coon, T. Takeshita, W. E. Wallace, and R. S. Craig, J. Phys. Chem. 80, 18781879
(1976).
41. W. E. Wallace, A. Elattar, T. Takeshita, V. T. Coon, C. A. Bechman, and R. S. Craig, in
J. Mulak, W. Suski, and R. Troc, eds., 2nd International Conference on the Electronic
Structure of the Actinides, Zaklad Narodowy Imienia Ossolinskick Wydawnictwo Polskiej Akademii NAUK, Wroclaw, Poland, 1977, pp 357365.
42. T. Takeshita, W. E. Wallace, and R. S. Craig, J. Catal. 44, 236243 (1976).
43. J. E. France and W. E. Wallace, Lanthanide Actinide Res. 2, 165180 (1988).
44. W. E. Wallace, CHEMTECH 12, 752754 (1982).
45. G. Owen, C. M. Hawkes, D. Lloyd, J. R. Jennings, R. M. Lambert, and R. M. Nix,
Appl. Catal. 33, 405430 (1987).
46. E. A. Shaw, T. Rayment, A. P. Walker, R. M. Lambert, T. Gauntlett, R. J. Oldman,
and A. Dent, Catal. Today 9, 197202 (1991).
47. K. Takeishi and K.-I. Aika, J. Catal. 136, 252257 (1992).
48. A. P. Walker, T. Rayment, and R. M. Lambert, J. Catal. 117, 102120 (1989).
49. J. B. Branco, C. J. Dias, and A. P. Goncalves, J. Alloys Compd. 478, 687693
(2009).

50. P. Rotterskamp, A. Kuklya, M.-A. Wustkamp,


K. Kerpen, C. Weidenthaler, and M.
Demuth, Angew. Chem., Int. Ed. 46, 77707774 (2007).
51. Q. Li and G. Lu, Catal. Lett. 125, 376379 (2008).
52. V. V. Lunin, Y. M. Bondarev, L. N. Padurec, S. I. Kondratev, and A. A. Chertkov,
Dokl. Akad. Nauk SSSR 220, 383385 (1975).
53. V. V. Lunin, V. I. Deineka, S. I. Abasov, and A. F. Plate, Neftekhimiya 15, 832835
(1975).
54. K. Soga, H. Imamura, and S. Ikeda, J. Phys. Chem. 81, 17621766 (1977).
55. K. Soga, H. Imamura, and S. Ikeda, J. Catal. 56, 119126 (1979).
56. K. Soga, H. Imamura, and S. Ikeda, Nippon Kagaku Kaishi 923929 (1978).
57. K. N. Semenenko and L. A. Petrova, Neftekhimiya 19, 2631 (1979).
58. P. A. Chernavskii and V. V. Lunin, Kinet. Catal. 24, 769773 (1983).
59. R. M. Frak, J. Less-Common Met. 109, 279286 (1985).
60. A. Ashraf, O. V. Chetina, L. A. Erivanskaya, E. S. Shpiro, G. V. Antoshin, and V. V.
Lunin, Neftekhimiya 23, 6165 (1983).
61. V. V. Lunin, Inorg. Mater. 14, 12451248 (1978).

30

INTERMETALLIC COMPOUNDS IN CATALYSIS

62. T. Imamoto, T. Mita, and M. Yokoyama, J. Chem. Soc., Chem. Commun. 163164
(1984).
63. R. K. Ibrasheva, T. A. Solomina, B. B. Bekkulov, and K. A. Zhubanov, Kinet. Katal.
27, 12781278 (1986).
64. R. K. Ibrasheva, T. A. Solomina, G. I. Leonova, V. V. Kiselev, P. A. Zhdan, G. I. Kaplan,
V. P. Mordovin, and R. G. Baisheva, Kinet. Catal. 31, 10571063 (1990).
65. V. V. Molchanov, V. V. Goidin, and R. A. Buyanov, Kinet. Catal. 47, 744746 (2006).
66. R. Valarivan, C. N. Pillai, and C. S. Swamy, React. Kinet. Catal. Lett. 53, 419428
(1994).
67. H. Oesterreicher and F. Spada, Mater. Res. Bull. 15, 477481 (1980).
68. D. E. Hall and V. R. Shepard, Int. J. Hydrogen Energy 9, 10051009 (1984).
69. S. Y. Vasina, S. A. Stuken, O. A. Petrii, I. L. Gogichadze, and V. A. Mukhin, Soviet
Electrochem. 23, 10661069 (1987).
70. V. Paul-Boncour, L. Hilaire, and A. Percheron-Guegan, in K. A. Gschneidner and
L. Eyring, eds., Handbook on the Physics and Chemistry of Rare Earths, Elsevier,
Amsterdam, the Netherlands, 2000, pp 544.
71. N. M. Parfenova, I. R. Konenko, A. L. Shilov, A. A. Tolstopyatova, E. I. Klabunovskii,
and M. E. Kost, Inorg. Mater. 14, 13331336 (1978).
72. H. Imamura, H. Yamada, K. Nukui, and S. Tsuchiya, J. Chem. Soc., Chem. Commun.
367368 (1986).
73. H. Imamura, S. Kasahara, T. Takada, and S. Tsuchiya, J. Chem. Soc., Faraday Trans.
1 84, 765772 (1988).
74. K. Shashikala, N. M. Gupta, P. Suryanarayana, A. Sathyamoorthy, V. S. Kamble, and
P. Raj, J. Molec. Catal. 91, 223235 (1994).
75. J. P. Candy, O. A. Ferretti, G. Mabilon, J.-P. Bournonville, El A. Mansur, J. M. Basset,
and G. Martino, J. Catal. 112, 210220 (1988).
76. U.S. Patent 4456775 (1984), C. Travers, T. D. Chan, R. Snappe, and J.-P.
Bournonville. Institut Francais du Petrole, Rueil-Malmaison, France.
and F. J. DiSalvo, Chem. Mater. 21,
77. A. Miura, H. Wang, B. M. Leonard, H. D. Abruna,
26612667 (2009).
78. X. Zhong, J. Zhu, and J. Liu, J. Catal. 236, 913 (2005).
79. S. Penner, D. Wang, D. S. Su, G. Rupprechter, R. Podloucky, R. Schlogl, and K. Hayek,
Surf. Sci. 532535, 276280 (2003).
80. S. Bernal, J. J. Calvino, M. A. Cauqui, J. M. Gatica, C. L. Cartes, J. A. P. Omil, and J.
M. Pintado, Catal. Today 77, 385406 (2003).
81. S. J. Tauster, S. C. Fung, and R. L. Garten, J. Am. Chem. Soc. 100, 170175 (1978).
82. S. J. Tauster, Acc. Chem. Res. 20, 389394 (1987).
83. N. Iwasa, T. Akazawa, S. Ohyama, K. Fujikawa, and N. Takezawa, React. Kinet.
Catal. Lett. 55, 245250 (1995).
84. G. Chen, S. Li, and Q. Yuan, Catal. Today 120, 6370 (2007).
85. M. Lenarda, L. Storaro, R. Frattini, M. Casagrande, M. Marchiori, G. Capannelli, C.
Uliana, F. Ferrari, and R. Ganzerla, Catal. Commun. 8, 467470 (2007).
86. S. Liu, K. Takahashi, K. Uematsu, and M. Ayabe, Appl. Catal. A 277, 265270 (2004).
87. S. Liu, K. Takahashi, and M. Ayabe, Catal. Today 87, 247253 (2003).
88. M. L. Cubeiro and J. L. G. Fierro, Appl. Catal. A 168, 307322 (1998).
89. M. L. Cubeiro and J. L. G. Fierro, J. Catal. 179, 150162 (1998).
90. Y. Wang, J. Zhang, H. Xu, and X. Bai, Chin. J. Catal. 28, 234238 (2007).
91. Y. Wang, J. Zhang, and H. Xu, Chin. J. Catal. 27, 217222 (2006).
92. S. Liu, K. Takahashi, K. Uematsu, and M. Ayabe, Appl. Catal. A 283, 125135
(2005).
93. N. Iwasa, S. Masuda, N. Ogawa, and N. Takezawa, Appl. Catal. A 125, 145157
(1995).

INTERMETALLIC COMPOUNDS IN CATALYSIS

31

94. Y. Kobayashi, Y. Tadaki, K. Takahashi, and M. Konno, J. Ceram. Soc. Japan 114,
654656 (2006).
95. G. D. Zakumbaeva, V. A. Naidin, T. S. Dagirov, and E. N. Litvyakova, Russ. J. Phys.
Chem. 61, 801808 (1987).
96. A. Sarkany, Z. Zsoldos, B. Furlong, J. W. Hightower, and L. Guczi, J. Catal. 141,
566582 (1993).
97. D. V. Sokolskii, L. M. Anisimova, and L. N. Edygenova, Russ. J. Phys. Chem. 60,
16391641 (1986).
98. F. Boccuzzi, A. Chiorino, G. Ghiotti, F. Pinna, G. Strukul, and R. Tessari, J. Catal.
126, 381387 (1990).
99. W. Li, Y. Chen, C. Yu, X. Wang, Z. Hong, and Z. Wei, in Proceedings of the 8th International Congress on Catalysis, Berlin, 1984, pp 205216.
100. M. Abid, G. Ehret, and R. Touroude, Appl. Catal. A 217, 219229 (2001).
101. E. Gebauer-Henke, R. Touroude, and J. Rynkowski, Kinet. Catal. 48, 562566 (2007).
102. S. Chettibi, R. Wojcieszak, E. H. Boudjennad, J. Belloni, M. M. Bettahar, and N.
Keghouche, Catal. Today 113, 157165 (2006).
103. W.-H. Cheng, Appl. Catal. A 130, 1330 (1995).
104. T. Okanishi, T. Matsui, T. Takeguchi, R. Kikuchi, and K. Eguchi, Appl. Catal. A 298,
181187 (2006).
105. H. E. Swift and J. E. Bozik, J. Catal. 12, 514 (1968).
106. S. T. Srinivas and P. K. Rao, J. Catal. 179, 117 (1998).
107. Z. Huang, J. R. Fryer, C. Park, D. Stirling, and G. Webb, J. Catal. 159, 340352
(1996).
108. N. A. Pakhomov, R. A. Buyanov, E. N. Yurchenko, A. P. Chernyshev, G. R. Kotelnikov,
E. M. Moroz, N. A. Zaitseva, and V. A. Patanov, Kinet. Catal. 22, 374380 (1981).
109. S. A. Bocanegra, A. Guerrero-Ruiz, S. R. de Miguel, and O. A. Scelza, Appl. Catal. A
277, 1122 (2004).
ar,
K. Matusek, and Z. Paal,

110. C. Kappenstein, M. Guerin, K. Laz


J. Chem. Soc., Faraday Trans. 94, 24632473 (1998).
111. M. Masai, K. Honda, A. Kubota, S. Ohnaka, Y. Nishikawa, K. Nakahara, K. Kishi,
and S. Ikeda, J. Catal. 50, 419428 (1977).
112. M. Masai, K. Mori, H. Muramoto, T. Fujiwara, and S. Ohnaka, J. Catal. 38, 128134
(1975).
113. J. Stachurski and J. M. Thomas, Catal. Lett. 1, 6772 (1988).
114. E. A. Sales, M. de Jesus Mendes, and F. Bozon-Verduraz, J. Catal. 195, 96105 (2000).
115. E. A. Sales, J. Jove, M. de Jesus Mendes, and F. Bozon-Verduraz, J. Catal. 195, 8895
(2000).
116. R. J. Kalenczuk, J. Chem. Technol. Biotechnol. 54, 349357 (1992).
117. R. Bacaud and P. Bussi`ere, J. Catal. 69, 399409 (1981).
118. M. A. Vannice and B. Sen, J. Catal. 115, 6578 (1989).
119. J. C. S. Wu, T.-S. Cheng, and C.-L. Lai, Appl. Catal. A 314, 233239 (2006).

120. G. Cardenas,
R. Oliva, P. Reyes, and B. L. Rivas, J. Molec. Catal. A 191, 7586 (2003).

121. M. D. C. Aguirre, P. Reyes, M. Oportus, I. Melian-Cabrera,


and J.-L. G. Fierro, Appl.
Catal. A 233, 183196 (2002).
122. K. Ishii, F. Mizukami, S.-I. Niwa, M. Toba, and T. Sato, Catal. Lett. 30, 297304
(1995).
123. M. Lucas and P. Claus, Chem. Eng. Technol. 28, 867870 (2005).
124. M. Steffan, M. Lucas, A. Brandner, M. Wollny, N. Oldenburg, and P. Claus, Chem.
Eng. Technol. 30, 481486 (2007).
125. E. Moretti, M. Lenarda, L. Storaro, R. Frattini, P. Patrono, and L. Pinzari, J. Colloid
Interface Sci. 306, 8995 (2007).
126. S. Karski, I. Witonska, and J. Goluchowska, J. Molec. Catal. A 245, 225230 (2006).

32
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.

INTERMETALLIC COMPOUNDS IN CATALYSIS


S. Karski, Przem. Chem. 85, 201204 (2006).
S. Karski, J. Molec. Catal. A 253, 147154 (2006).
S. Karski and I. Witonska, J. Molec. Catal. A 191, 8792 (2003).
T. Chojnacki, K. Krause, and L. D. Schmidt, J. Catal. 128, 161185 (1991).
S. Naito, T. Hasebe, and T. Miyao, Chem. Lett. 11191120 (1998).
K. Hayek, H. Goller, S. Penner, G. Rupprechter, and C. Zimmermann, Catal. Lett. 92,
19 (2004).
S. Penner, G. Rupprechter, H. Sauer, D. S. Su, R. Tessadri, R. Podloucky, R. Schlogl,
and K. Hayek, Vacuum 71, 7176 (2003).
H. Lorenz, Q. Zhao, S. Turner, B. L. Lebedev, Van G. Tendeloo, B. Klotzer, C. Rameshan, K. Pfaller, J. Konzett, and S. Penner, Appl. Catal. A 381, 242252 (2010).
H. Lorenz, S. Turner, O. I. Lebedev, Van G. Tendeloo, B. Klotzer, C. Rameshan, K.
Pfaller, and S. Penner, Appl. Catal. A 374, 180188 (2010).
S. Penner, B. Jenewein, H. Gabasch, B. Klotzer, D. Wang, A. Knop-Gericke, R. Schlogl,
and K. Hayek, J. Catal. 241, 1419 (2006).
S. Penner, B. Jenewein, and K. Hayek, Catal. Lett. 113, 6572 (2007).
S. Penner, H. Lorenz, W. Jochum, M. Stoger-Pollach, D. Wang, C. Rameshan, and B.
Klotzer, Appl. Catal. A 358, 193202 (2009).
T. Komatsu, M. Mesuda, and T. Yashima, Appl. Catal. A 194195, 333339 (2000).
A. Onda, T. Komatsu, and T. Yashima, J. Catal. 201, 1321 (2001).
A. Onda, T. Komatsu, and T. Yashima, J. Catal. 221, 378385 (2003).
T. Komatsu, T. Kishi, and T. Gorai, J. Catal. 259, 174182 (2008).
T. Komatsu and M. Fukui, Appl. Catal. A 279, 173180 (2005).
T. Komatsu, K. Inaba, T. Uezono, A. Onda, and T. Yashima, Appl. Catal. A 251, 315
326 (2003).

G. Rienacker,
Z. Elektrochem. 47, 805809 (1941).

G. Rienacker
and H. Hildebrandt, Z. Anorg. Allg. Chem. 248, 5264 (1941).
G.-M. Schwab, Trans. Faraday Soc. 42, 689697 (1946).
D.-H. Chun, Y. Xu, M. Demura, K. Kishida, M.-C. Kim, M.-H. Oh, T. Hirano, and
D.-M. Wee, J. Korean Inst. Met. Mater. 43, 801809 (2005).
P. P. Clopp and G. Parravano, J. Phys. Chem. 62, 10551059 (1958).
T. Komatsu, S.-I. Hyodo, and T. Yashima, J. Phys. Chem. B 101, 55655572 (1997).
H. Verbeek and W. M. H. Sachtler, J. Catal. 42, 257267 (1976).
T. Komatsu, M. Fukui, and T. Yashima, Stud. Surf. Sci. Catal. 101, 10951104 (1996).
A. Onda, T. Komatsu, and T. Yashima, Phys. Chem. Chem. Phys. 2, 29993005 (2000).
J. B. Branco, A. P. Goncalves, and A. P. de Mato, J. Alloys Compd. 465, 361366
(2008).

P. Claus, F. Raif, S. Cavet, S. Demirel-Gulen,


J. Radnik, M. Schreyer, and T. Fassler,
Catal. Comm. 7, 618622 (2006).

C. Gieck, M. Schreyer, T. F. Fassler,


F. Raif, and P. Claus, Eur. J. Inorg. Chem. 3482
3488 (2006).

C. Gieck, M. Schreyer, T. F. Fassler,


S. Cavet, and P. Claus, Chem. Eur. J. 12, 1924
1930 (2006).

V. Hlukhyy, F. Raif, P. Claus, and T. F. Fassler,


Chem. Eur. J. 14, 37373744 (2008).
S. M. Kauzlarich, in R. B. King, ed., Encyclopedia of Inorganic Chemistry, Wiley,
Hoboken, N.J., 2006.
B. I. Popov, N. G. Skomorokhova, V. V. Karonik, and V. E. Kolesnichenko, React. Kinet.
Catal. Lett. 27, 419423 (1985).

M. Friedrich, A. Ormeci, Y. Grin, and M. Armbruster,


Z. Anorg. Allg. Chem. 636,
17351739 (2010).
N. Iwasa, T. Mayanagi, S. Masuda, and N. Takezawa, React. Kinet. Catal. Lett. 69,
355360 (2000).

INTERMETALLIC COMPOUNDS IN CATALYSIS


163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.
180.
181.
182.

183.
184.
185.
186.
187.
188.
189.
190.

191.

192.

33

A. P. Tsai, S. Kameoka, and Y. Ishii, J. Phys. Soc. Japan 73, 32703273 (2004).
N. Takezawa and N. Iwasa, Catal. Today 36, 4556 (1997).
A. K. M. S. Huq and A. J. Rosenberg, J. Electrochem. Soc. 111, 270278 (1964).
E. W. Justi, H. H. Ewe, A. W. Kalberlah, N. M. Saridakis, and M. H. Schaefer, Energy
Conv. 10, 183187 (1970).
E. W. Brooman and A. T. Kuhn, Electroanal. Chem. Interfacial Electrochem. 49, 325
353 (1974).
M. H. Miles, Electroanal. Chem. Interfacial Electrochem. 60, 8996 (1975).
M. M. Jaksic, Electrochim. Acta 29, 15391550 (1984).
A. Belanger, and A. K. Vijh, Int. J. Hydrogen Energy 12, 227233 (1987).
A. B. Shein, Russ. Electrochem. 24, 12331236 (1988).
A. K. Vijh, G. Belanger, and R. Jaques, Mater. Chem. Phys. 21, 529538 (1989).
A. C. D. Angelo, C. Lind, F. J. DiSalvo, and H. D. Abruna,

E. Casado-Rivera, Z. Gal,
Chem Phys Chem 4, 193199 (2003).

E. Casado-Rivera, D. J. Volpe, L. Alden, C. Lind, C. Downie, T. Vazquez-Alvarez,


A. C.
J. Am. Chem. Soc. 126, 40434049 (2004).
D. Angelo, F. J. DiSalvo, and H. D. Abruna,
J. Electrochem.
F. Matsumoto, C. Roychowdhury, F. J. DiSalvo, and H. D. Abruna,
Soc. 155, B148-B154 (2008).
M. M. P. Janssen and J. Moolhuysen, Electrochim. Acta 21, 861868 (1976).
Z. Liu, D. Reed, G. Kwon, M. Shamsuzzoha, and D. E. Nikles, J. Phys. Chem. C 111,
1422314229 (2007).
L. Zhang and D. Xia, Appl. Surf. Sci. 252, 21912195 (2006).
T. Ghosh, Q. Zhou, J. M. Gregoire, R. B. van Dover, and F. J. DiSalvo, J. Phys. Chem.
C 114, 1254512553 (2010).
Langmuir 25, 77257735
H. Wang, L. R. Alden, F. J. DiSalvo, and H. D. Abruna,
(2009).
and F. J. DiSalvo, J.
H. Abe, F. Matsumoto, L. R. Alden, S. C. Warren, H. D. Abruna,
Am. Chem. Soc. 130, 54525458 (2008).
D. J. Volpe, E. Casado-Rivera, L. Alden, C. Lind, K. Hagerdon, C. Downie, C. Ko J. Electrochem. Soc. 151, A971A977
rzeniewski, F. J. DiSalvo, and H. D. Abruna,
(2004).
and F. J. DiSalvo, Chem.
C. Roychowdhury, F. Matsutomo, P. F. Mutolo, H. D. Abruna,
Mater. 17, 58715876 (2005).

J. Osswald, K. Kovnir, M. Armbruster,


R. Giedigkeit, R. E. Jentoft, U. Wild, Y. Grin,
and R. Schlogl, J. Catal. 258, 219227 (2008).

K. Kovnir, J. Osswald, M. Armbruster,


R. Giedigkeit, T. Ressler, Y. Grin, and R.
Schlogl, Stud. Surf. Sci. Catal. 162, 481488 (2006).

J. Osswald, R. Giedigkeit, R. E. Jentoft, M. Armbruster,


F. Girgsdies, K. Kovnir, T.
Ressler, Y. Grin, and R. Schlogl, J. Catal. 258, 210218 (2008).

K. Kovnir, M. Armbruster,
D. Teschner, T. V. Venkov, L. Szentmiklosi, F. C. Jentoft,
A. Knop-Gericke, Y. Grin, and R. Schlogl, Surf. Sci. 603, 17841792 (2009).
J. Singh, C. Lamberti, J. A. van Bokhoven, Chem. Soc. Rev. 39, 47544766 (2010).
Z. Revay, T. Belgya, L. Szentmiklosi, Z. Kis, A. Wootsch, D. Teschner, M. Swoboda, R.
Schlogl, J. Borsodi, and R. Zepernick, Anal. Chem. 60666071 (2008).

H. Bluhm, M. Havecker,
A. Knop-Gericke, E. Kleimenov, R. Schlogl, D. Teschner, V.
I. Bukhtiyarov, D. F. Ogletree, and M. Salmeron, J. Phys. Chem. B 108, 1434014347
(2004).

D. Teschner, E. Vass, M. Havacker,


S. Zafeiratos, P. Schnorch, H. Sauer, Knop-A.
Gericke, R. Schlogl, M. Chamam, A. Wootsch, A. S. Canning, J. J. Gamman, S. D.
Jackson, J. McGregor, and L. F. Gladden, J. Catal. 242, 2637 (2006).

D. Teschner, J. Borsodi, A. Wootsch, Z. Revay, M. Havecker,


A. Knop-Gericke, S. D.
Jackson, and R. Schlogl, Science 320, 8689 (2008).

34

INTERMETALLIC COMPOUNDS IN CATALYSIS

193. D. Teschner, Z. Revay, J. Borsodi, M. Havecker,


A. Knop-Gericke, R. Schlogl, D. Milroy, S. D. Jackson, D. Torres, and P. Sautet, Angew. Chem., Int. Ed. 47, 92749278
(2008).

194. K. Wolter, O. Seiferth, J. Libuda, H. Kuhlenbeck, M. Baumer,


and H.-J. Freund, Surf.
Sci. 402404, 428 (1998).

195. K. Wolter, O. Seiferth, H. Kuhlenbeck, M. Baumer,


and H.-J. Freund, Surf. Sci. 399,
190 (1998).
196. Y. Jin, A. K. Datye, E. Rightor, R. Gulotty, W. Waterman, M. Smith, M. Holbrook, J.
Maj, and J. Blackson, J. Catal. 203, 292306 (2001).
197. N. A. Khan, S. Shaikhutdinov, and H.-J. Freund, Catal. Lett. 108, 159164 (2006).
198. A. A. Lamberov, S. R. Egorova, I. R. Ilyasov, Kh. Kh. Gilmanov, S. V. Trifonov, V. M.
Shatilov, and A. Sh. Ziyatdinov, Kinet. Catal. 48, 136142 (2007).
199. D. Mei, M. Neurock, and C. M. Smith, J. Catal. 268, 181195 (2009).
200. H. Zea, K. Lester, A. K. Datye, E. Rightor, R. Gulotty, W. Waterman, and M. Smith,
Appl. Catal. A 282, 237245 (2005).

201. K. Kovnir, M. Schmidt, C. Waurisch, M. Armbruster,


Y. Prots, and Y. Grin, Z.
Kristallogr.NCS 223, 78 (2008).

202. M. Armbruster,
H. Borrmann, M. Wedel, Y. Prots, R. Giedigkeit, and P. Gille, Z.
Kristallogr.NCS 225, 617618 (2010).
203. K. Khalaff and K. Schubert, J. Less-Common Met. 37, 129140 (1974).
204. G.-M. Schwab and E. Schwab-Agallidis, Ber. Dtsch. Chem. Gesellsch. 76, 12281256
(1943).
205. G.-M. Schwab and A. Karatzas, Z. Elektrochem. 50, 242249 (1944).
206. G.-M. Schwab and S. Pesmatjoglou, J. Phys. Chem. 52, 10461053 (1948).
207. G.-M. Schwab, Experientia 2, 103105 (1946).

208. B. Frank, F. C. Jentoft, H. Soerijanto, J. Krohnert, R. Schlogl, and R. Schomacker,


J.
Catal. 246, 177192 (2007).
209. V. Ponec, Appl. Catal. A 149, 2748 (1997).

210. E. Galloway, M. Armbruster,


K. Kovnir, M. Tikhov, and R. M. Lambert, J. Catal. 261,
6065 (2009).
211. J. L. Murray, Int. Met. Rev. 30, 211233 (1985).
212. T. B. Massalski, in T. B. Massalski, ed., Binary Alloy Phase Diagrams, ASM International, Materials Park, Ohio, 1990, pp 18361837.

M. ARMBRUSTER
Chemische Physik
Max-Planck-Institut fur
fester Stoffe
Dresden, Germany

También podría gustarte