Anda di halaman 1dari 10

Topics in Catalysis 8 (1999) 115124

115

Catalyst potential measurement:


a valuable tool for understanding and controlling liquid phase
redox reactions
T. Mallat and A. Baiker
Laboratory of Technical Chemistry, Swiss Federal Institute of Technology, ETH-Zentrum, CH-8092 Zurich, Switzerland

The experimental technique and possible applications of the steady state catalyst potential measurement during Pt metal catalyzed
hydrogenation and oxidation reactions in slurry reactors has been reviewed. The value of this in situ method is illustrated by many
examples, including the separation of consecutive reaction steps, determination of the oxidation state of platinum metal catalyst and
promoter during reaction, elucidation of the role of surface impurities, and controlling the rate of alcohol oxidation by optimizing the
surface concentration of the oxidizing species.
Keywords: catalyst potential, hydrogenation, alcohol oxidation, supported metal catalysts, oxidation state

1. Introduction
The possibilities of using in situ methods for studying
the solid/liquid interface in catalysis are rather limited, compared to the numerous spectroscopic, thermoanalytical and
other techniques available for investigating solid/gas type
reactions. Perhaps the most successful method for examining solid catalysed reactions in the liquid phase is electrocatalysis, an interdisciplinary field between heterogeneous
catalysis and electrochemistry. Electrocatalysis has provided valuable information on the behaviour of metal and
oxide surfaces in polar (mainly aqueous) medium. Great
advantages of this technique are that the (oxidation) state of
the metal surface can easily be controlled via the electrode
potential, and the influence of oxidation state and hydrogen
or oxygen coverage on the catalytic performance can be
studied. The characteristics of heterogeneous catalytic and
electrocatalytic systems are in many respect rather similar
[1,2]. Still, the transfer of this knowledge to heterogeneous
catalysis is not always straightforward. For example, when
comparing to a supported metal catalyst, even a high surface area metal electrode possesses extremely low dispersion (i.e., the number of surface atoms related to the total
number of atoms). An engineering problem is that the rate
of mass transfer is frequently a limiting factor in the electrochemical cell.
Muller and Schwabe [36] were the first who tried to
separate the two major functions of the electrode-catalyst:
they decomposed formic acid by a suspended noble metal
catalyst powder and measured the catalyst potential during reaction by an inactive (very low surface area) Pt
wire as a collector electrode. After their pioneer work,
mainly Russian and German scientists [1,716] developed
the method further for the in situ investigation of metal
hydrogenation catalysts in slurry reactors. The general
electrochemical properties of these slurry electrodes com J.C. Baltzer AG, Science Publishers

pared to those of conventional compact electrodes have also


been thoroughly investigated [1722].
The application range of catalyst potential measurement is not very broad due to the obvious limitations in
conductivity. However, in several cases this is the only
method which can provide real, on-time information on the
working catalyst. Catalyst potential measurements demonstrated that during alcohol oxidation on Pt with molecular oxygen the metal surface can be in a reduced state
and partially covered by hydrogen [3,23]. On the other
hand, the metal surface may be (partially) oxidized during the hydrogenation of strongly oxidizing reactants and
only the appropriate choice of reaction conditions can help
in avoiding this situation. For example, during hydrogenation of nitro-compounds in a slurry reactor, continuous addition of this reactant is necessary to decrease
the redox potential of the reactant/(intermediates)/product
system and avoid the over-oxidation and deactivation of
Raney Ni.
In the following chapters various applications of the catalyst potential measurement in slurry reactors will be discussed. The measurement of catalyst potential at high temperature in a gas/solid type reactor and the electrochemical modification of catalytic activity using solid electrolytes
will not be addressed here; for these application the reader
should consult with former reviews [2427].

2. Experimental technique
When measuring the potential of a catalyst slurry, the
crucial question is whether the potential indicated by the
collector electrode (also termed as probe, sensor or measuring electrode) is identical with the mean potential of the
catalyst particles. In general, four different types of charge
transfer are possible: (i) between the catalyst particles and

116

T. Mallat, A. Baiker / Catalyst potential measurement

the electrode, (ii) between two catalyst particles, (iii) between the catalyst particles and the reactants, and (iv) between the electrode and the reactants [18,19,28,29]. The
rate of charge transfer is proportional with the frequency
of collision between catalyst particles and electrode, the
duration of a single collision and the potential difference
between the particles and the electrode, and inversely proportional with the sum of the resistance of the electronic
contact and the electrolyte [30,31]. Accordingly, a bright
(low real surface area), small capacity metal wire or net
with relatively high collision surface should be used as a
collector electrode.
It is clear from the above discussion that the choice of
electrode material depends on the nature of the reaction
studied. Pt, Au and Ag were suggested most frequently to
minimize the electrochemical reaction at the surface of the
collector electrode, compared to the reactions at the catalyst surface [29,3235]. However, in acidic medium Ag
can easily corrode [35]). Cu, Ni and glassy carbon, or an Fe
screen collector electrode as suggested recently [36], are
usually not suitable [18,37]. During applications in oxidation reactions, surface contamination can seriously distort
the measured value [38]. Regular replacement and purification of the probe electrode, or the simultaneous application
of two different electrodes (e.g., Pt and Au) can reveal this
complication.
It is important that the method can be extended to noble metals on non-conductive supports (e.g., alumina and
silica [1,14,37]), on conditions that the metal particles are
mainly located on the outer surface of the support, and the
metal loading and the catalyst concentration in the slurry
are sufficiently high. Carbon particles may adhere to the
electrode and the efficiency of mass transfer is significantly
lower in this porous system compared to the well mixed
slurry [39]. The poor exchange between adhered and suspended particles can distort the observed potential when
using carbon-supported metals.
Two further points have to be considered for catalytic
applications. The low conductivity of the reaction medium
may limit the application. Conducting aqueous acidic or basic solutions may be unacceptable for many purposes. The
presence of a conducting electrolyte which adsorbs weakly
on the metal catalyst, e.g., quaternary ammonium salts for
organic medium and Na2 SO4 for aqueous and aqueousalcoholic solutions, can be a good compromise.
Several different types of reactors have been suggested
for the reliable measurement of the potential of suspended
catalysts during a redox reaction [31,3941]. It is important to note that catalyst contamination by mercury is almost inevitable when using the Hg/HgCl2 /Cl reference electrode. Control experiments are necessary even in case of the
Ag/AgCl/Cl electrode. The glass filter or ground, which
usually separates the reaction medium from the reference
electrode, can slow down but not hinder the migration of
metal and halogen ions to the catalyst surface.

Scheme 1.

3. Applications in hydrogenation reactions


3.1. Hydrogenation of alkenes and alkynes; radical
mechanism
When a metal catalyst is used for the hydrogenation of
an organic reactant in a polar medium, the metalhydrogen
reactant system can be considered as a catalytic and as an
electrochemical system (scheme 1). The hydrogenation of
some organic reactants obeys a simple radical mechanism
without forming any ionic intermediate or product [42,43].
At constant pH the catalyst potential shift during reaction
is governed by the change in surface hydrogen activity,
according to the Nernst equation:
E=

RT aH,1
ln
,
F
aH,2

(1)

where R is the gas constant, T the absolute temperature


and F the Faraday constant. At ambient conditions a
change of surface hydrogen activity by an order of magnitude shifts the catalyst potential by ca. 60 mV. Under
appropriate conditions the catalyst potential can be measured within a range of 5 mV (510 mV during alcohol
oxidation, see later). These values illustrate the sensitivity
and accuracy of the method.
Figure 1 presents the hydrogen consumption and catalyst potential during the hydrogenation of maleic acid over
Pt [44]. After addition of the reactant to the slurry the potential increased by almost 80 mV (related to the hydrogen
electrode in the same solution, RHE), and then decreased
slowly and reached the original hydrogen equilibrium potential at complete conversion. Several other examples
can be found in the literature, including the potential measurement during the hydrogenation of maleic acid sodium
salt [40], crotonic acid [22], allyl alcohol [45], cinnamyl alcohol [45,46], methyl cinnamate [45], 2-methyl-3-butyne2-ol [12,14,40,47,48], and 2-butyne-1,4-diol [1,49]. When
the rate of hydrogen supply is sufficiently high, the shift
of catalyst potential in alkene and alkyne hydrogenation
is moderate, usually below 100 mV. A potential shift by
several hundred mV may be a strong indication for mass
transfer limitations in the reactor.
3.2. Kinetic analysis of a zero-order reaction
In the hydrogenation of some functionalized alkenes and
alkynes the rate of hydrogen consumption was found to be
constant in a broad range of reactant concentration, while

T. Mallat, A. Baiker / Catalyst potential measurement

117

Figure 3. Schematic representation of the steady state catalyst potential


considered as a mixed potential of two redox couples.
Figure 1. Hydrogen consumption and catalyst potential during the hydrogenation of maleic acid over a Pt powder catalyst. Conditions: 1 M
aqueous sulfuric acid, SPt = 12 m2 g1 , potential related to RHE [44].

chemical potential of hydrogen, offering a new explanation


for the observed zero reaction order [44].
3.3. Hydrogenation by ionic (electrochemical) mechanism
The hydrogenation and hydrogenolysis of nitro and
many carbonyl and halogenated compounds obeys an ionic
or electrochemical mechanism [43,49], according to the following simplified scheme:
Had
H+ + e
R + e + H+
RH

Figure 2. Rate of hydrogen consumption and catalyst potential as a function of the logarithm of maleic acid concentration. Conditions: 1 M
sulfuric acid, SPt = 12 m2 g1 , potential related to RHE [44].

the catalyst potential changed significantly [1,47]. A thoroughly investigated example is the hydrogenation of maleic
acid, as shown in figure 1. The kinetic analysis revealed
that the surface coverage by the reactant decreased with the
logarithm of its bulk concentration (figure 2). This effect
on the reaction rate was compensated by the change of the

(2)
(3)

The equation (3) can be dissected into two processes, the


reduction of the reactant (R) via the transfer of an electron
and subsequently a proton, or via the first addition of a proton followed by an electron. A more complex scheme has
been proposed for the reduction of quinone and nitrobenzene [49].
As soon as the overall catalytic process is composed
of two or more electrode reactions, the measured steady
state potential is a mixed potential. The development of
the mixed potential by applying the additivity principle of
Wagner and Traud [50] is illustrated in figure 3. The two
redox couples are shown in a general and simplified form
and can be applied for both catalytic hydrogenation and
alcohol oxidation reactions (see later):
red1
ox1 + n1 e
ox2 + n2 e
red2

(4)
(5)

Each Pt particle functions as a short-circuited electrochemical cell, where both the anodic and cathodic halfreactions take place at the same rate and the same potential
(Emix ). By Faradays law the reaction rate is proportional
to the mixture current (imix ). At the mixed potential the

118

T. Mallat, A. Baiker / Catalyst potential measurement

anodic and cathodic currents are equal, and no net current


flows through the catalyst. E1 and E2 can be expressed by
the Nernst equation:
E = E +

RT
[ox]
ln
.
F
[red]

(6)

A qualitative analysis of figure 3 [51] shows that imix and


thus the reaction rate will increase, if:
(i) the difference between E2 and E1 is bigger (which
corresponds to the hydrogenation of a strongly oxidizing reactant),
(ii) the concentrations of red1 and ox2 are enhanced (which
will change the equilibrium potentials and the exchange currents),
(iii) each of the two curves is more angular (larger exchange current density; in catalysis rather the turn-over
frequency is used for characterising the active sites).
More detailed discussions of the mixed potential theory and
the limited validity of the additivity principle in heterogeneous catalysis can be found in former papers and reviews
[2,42,43,5258].
There are several examples in the literature on the potential measurement during the hydrogenation of nitro and
carbonyl compounds [1,12,40,45,47], and the hydrogenation hydrogenolysis of halonitro compounds [59]. The
parameters varied are the solvent, the catalyst composition
and the addition of some catalyst poisons and modifiers.
In the hydrogenation of strongly oxidizing compounds the
steady state catalyst potential before and during reaction
can deviate by up to a few hundred mV. Note that during
the reduction of nitro compounds the parallel measurement
of pH is necessary for the interpretation of the catalyst

Figure 4. Catalyst potential at ca. 50% conversion of nitrobenzene as


a function of pH in buffered aqueous methanolic solutions. Conditions:
50% aqueous methanol, 5 wt% Pd/C, 1 bar, room temperature, potential
related to RHE [49].

potential due to the formation of basic intermediates and


products.
Figure 4 illustrates that the catalyst potential at ca. 50%
conversion of nitrobenzene (half wave reduction potential) varied systematically with the pH of the medium [49].
The likely explanation for the observed correlation was that
the rate determining step changed with the variation of the
pH. On the basis of the slope of the redox potential related to a Ag/AgCl/Cl reference electrode and plotted as
a function of pH (not shown here) it was suggested that at
pH < 3 a radical cation intermediate, at 3 < pH < 7 a
neutral free radical, and at pH > 7 an anionic intermediate
were formed [49].
3.4. Selective hydrogenation by the separation of
consecutive reaction steps
It has been suggested [1,12,40] that following the steady
state catalyst potential during a partial hydrogenation reaction can help in detecting the end-point of one reaction
step in a consecutive reaction series. The most frequently
used example is the selective hydrogenation of alkynes to
alkenes. The hydrogenation of the C=C bond is faster than
that of the triple bond, but the alkyne adsorbs stronger on
the metal (Pd) surface and the hydrogenation of the alkene
intermediate is minor until the alkyne is consumed.
The change of the dominant surface reaction is usually
clearly indicated by the change of the catalyst potential
at the end of the first step. Still, the separation is not
always easy as the catalyst potential shifts during the first
and second steps vary with the reaction conditions and the
catalyst composition. The effect of lead poisoning of a
Pd/C catalyst in the hydrogenation of 2-butyne-1,4-diol is
illustrated in table 1, as an example [49]. Lead poisoning
retarded the hydrogenation of the C=C double bond to a
larger extent than that of the triple bond, as expected. The
decrease of the rates in the first and second reaction steps
was mirrored by the catalyst potential shifts. At medium
Pb/Pd ratio the potential shifts were rather similar in alkyne
and alkene hydrogenation which renders the separation of
the two reaction steps ambiguous.
Another, practically important partial hydrogenation reaction is the transformation of chloro-nitrobenzenes to the
corresponding chloroanilines [59]. The development of
catalyst potential during the hydrogenation of 2-chloronitrobenzene is shown in figure 5. The reaction ran in two
reasonably separated steps: the hydrogenation of the nitro
group and the subsequent hydrogenolysis of the CCl bond.
The catalyst potential reached a local minimum at around
100% hydrogen uptake related to the hydrogenation of the
nitro group. At this point the selectivity to the intermediate chloroaniline was the highest (70%). Interrupting the
hydrogenation at the potential minimum seems to be relatively simple, at least on laboratory scale. For industrial
application it was more fruitful to use some soluble catalyst
modifiers which hindered the hydrogenolysis side reaction.
Addition of amidine derivatives, such as formamidine ac-

T. Mallat, A. Baiker / Catalyst potential measurement

119

Table 1
Change of catalyst potential (half wave potential) and reaction time due to selective poisoning of Pd by lead
nitrate. Conditions: 0.4 g 5 wt% Pd/C, 0.2 g 2-butyne-1,4-diol, 1 M aqueous acetic acid, pH = 2.3, room
temperature, 1 bar [49].

Pb/Pd
atom/atom

0
0.06
0.12
0.16
0.18

Hydrogenation time
(min)
A

14
17
20
25
34

11
16
24
70
>180

Figure 5. Formation of chloroaniline () and aniline (N), and the change


of catalyst potential (dashed line) during the hydrogenation of 2-chloronitrobenzene. Conditions: Raney Ni, methanol, 30 C, 1.1 bar, potential
related to a Ag/AgCl reference electrode [59].

etate or dicyandiamide, to the reaction mixture afforded 91


99.7% selectivities in the partial hydrogenation of various
aromatic chloronitro compounds [59]. In the presence of
these modifiers the hydrogenolysis side reaction was minor
and the catalyst potential remained at its minimum value
after the nitro amino transformation.
3.5. Oxidation state of promoter
Metal ion promoters or poisons are widely used in catalytic hydrogenations to control the product distribution.
The interpretation of the selectivity improvement and localization of the real active species is usually rather spec-

Catalyst potential (RHE)


(mV)
A
B
145
140
125
90
60

195
165
110
30
20

ulative. Ex situ measurement of the oxidation state of the


bi- or multi-component catalyst cannot provide unambiguous results. The following example illustrates the value of
catalyst potential measurement for this purpose.
The hydrogenation of nitrate ions to hydroxylamine
in a strongly acidic aqueous solution is an industrial
process [60]. The reaction is carried out in a slurry reactor
with Pd/C promoted by GeO2 . It was assumed that the promoter is reduced to metal by hydrogen on the Pd surface,
and the outstanding catalytic performance was attributed to
the presence of bimetallic active site ensembles. Later, electrocatalytic measurements confirmed this assumption [61]
and demonstrated that (partially) discharged Ge adatoms
deposited onto Pd at ambient conditions in aqueous acidic
medium (underpotential deposition of adsorbed Ge [62]).
Ge adatoms do not adsorb hydrogen [61] and excluding some special reactions such as nitrate ion and allyl alcohol reduction they act as catalyst poisons in hydrogenation reactions [45,46]. Astonishingly, in some reactions the
poisoned catalyst progressively regenerated during reaction
and the rate of hydrogen consumption increased dramatically with increasing conversion. The hydrogenation of
nitrobenzene is shown in figure 6 as an example. The measurement of catalyst potential provided a reasonable explanation to the unexpected rate enhancement [63]. During the hydrogenation of nitrobenzene the catalyst potential
was in the range of 150190 mV (figure 6). At this potential a fraction of Ge adatoms were oxidized [61] and
dissolved during reaction. The corrosion of Ge promoter
could be confirmed by atomic emission spectroscopy (figure 6). As a result, more and more Pd atoms were accessible for the hydrogenation reaction and the rate increased
remarkably.
In many other reactions the poisoning of Pd by Ge was
irreversible due to the significantly lower catalyst potential
shift during reaction: e.g., 1015 mV in the hydrogenation
of methyl cinnamate and 1520 mV during acetophenone
reduction [45]. At these low potentials the Ge adatoms
were rather stable and the poor activity of Ge-covered Pd
was preserved during reaction.

120

T. Mallat, A. Baiker / Catalyst potential measurement

Figure 7. Catalyst potential during catalyst prereduction and during


the enantioselective hydrogenation of ethyl pyruvate over cinchonidinemodified Pt/alumina.
Conditions:
5 wt% Pt/alumina, solvent:
1-propanol + 5 ml 0.1 M aqueous Na2 SO4 , cinchonidine, ethyl pyruvate,
1 bar, room temperature, potential related to RHE [64]. The characteristics
of the purified () and impure () systems are described in the text.

In the second case (impure system) the reaction mixture was stirred for 2 h in a nitrogen atmosphere before
hydrogenation. The decreasing catalyst potential manifests
the reduction of Pt by the solvent 1-propanol,
CH3 CH2 CH2 OHad CH3 CH2 CHO + 2Had
PtOx + Had Pt + H2 O
Figure 6. Corrosion of a GePd/C catalyst during the hydrogenation of
nitrobenzene. Conditions: Ge/Pds = 2.6 atom/atom, 60 vol% aqueous
EtOH, 25 C, 1 bar [63].

3.6. Direct evidence to the role of surface impurities


In many surface science techniques a large effort is required to purify the system and minimize the surface contamination, which can distort the investigated phenomenon
on the low surface area metal substrate. Practical catalytic
hydrogenation is relatively insensitive to impurities due to
the much higher specific surface area of supported metal
catalysts. A direct evidence for the disturbance by surface
impurities has been reported recently in the enantioselective
hydrogenation of ethyl pyruvate [64]. Figure 7 illustrates
the catalyst potential up to 8090% conversion. In one case
(purified system) ethyl pyruvate was carefully distilled
before use and the Pt/alumina catalyst pretreated at 400 C
in flowing hydrogen. During transferring the catalyst to
the reactor, the metal surface was reoxidized by air, but
the catalyst potential dropped immediately to 145150 mV
after introduction of hydrogen to the reaction mixture. The
potential increased slowly with time to 220230 mV and
decreased again only above 90% conversion (>500 min).

(7)
(8)

but the reduction was incomplete due to the destructive


adsorption of 1-propanol and the resulting poisoning of Pt
[6567]:
CH3 CH2 CH2 OHad COad + Cx Hy Oz

(9)

Introduction of hydrogen at the beginning of pyruvate reduction decreased the catalyst potential to only 220 mV
and the potential remained between 220 and 235 mV up to
90% conversion. Figure 7 demonstrates that at low conversion the catalyst potentials in the two reactions differed
by 70 mV, which corresponds to a significant variation
in surface concentrations including hydrogen and organic
species (mixed potential). The different surface concentrations were clearly mirrored in the enantiomeric excesses
obtained at 80% conversion: 34% in the purified system
but only 20% in the impure system.
Above 50% conversion the difference between the time
resolved potential curves was minor. It was proposed
[64,68] that hydrogen can efficiently compete with CO
and reduce the hydrocarbon fragments to weakly adsorbing species (C2 H6 , C3 H8 ). This is the likely reason why
relatively good enantiomeric excess can be achieved even
in impure systems at high hydrogen pressure (70100 bar)

T. Mallat, A. Baiker / Catalyst potential measurement

121

at which condition the surface hydrogen concentration is


sufficiently high for the rapid removal of contaminants and
to minimize their disturbing effect on the enantiodifferentiation.

4. Applications in the oxidation of alcohols


The liquid phase aerobic oxidation of alcohols is an ideal
field for the application of catalyst potential measurements
as the reactions are usually carried out in weakly basic or
acidic aqueous solutions over Pt metal catalysts [69,70].
A further advantage is the rather broad range of catalyst
potential during reaction, between ca. 100 and 850 mV
[71,72]. On the other hand, oxidation reactions are less selective than catalytic hydrogenations and the strong adsorption of (high molecular weight) by-products on the collector
electrode can distort the measured potential [37].
4.1. Oxidation state of mono- and bimetallic catalysts
Catalyst potential measurements revealed very early that
during the oxidation of some simple primary and secondary
alcohols the catalyst potential is rather low, only 0.10.2 V
related to RHE [3,23]. In aqueous acidic or basic solutions the Pt and Pd surface is partially covered by hydrogen below 0.30.4 V and by oxygen (or OHad ) above 0.6
0.7 V [73]. The interesting observation that there is adsorbed hydrogen on the metal surface during the oxidation
with molecular oxygen was explained by a dehydrogenation mechanism; in a subsequent step the adsorbed hydrogen is oxidized to water [69,70,74]. The low potential of
a Bi-promoted Pt/alumina catalyst during the oxidation of
diphenyl carbinol with air is shown in figure 8, as an illustration.
It was demonstrated later that the reduced state of Pt
and Pd during reaction is not a general characteristics of
alcohol oxidation [37,38,7578]. In many less facile reactions, such as the partial oxidation of cinnamyl alcohol,
1-methoxy-2-propanol and -tetralol, the noble metal was
in a partially oxidized state even at moderate conversions.
The likely explanation was some catalyst poisoning due to
the destructive adsorption of alcohol on Pt or Pd, as discussed above [6567]. On the basis of the mixed potential
theory (see figure 3) it is easy to understand that poisoning
the dehydrogenation reaction increases the catalyst potential. The oxidation of L-sorbose to 2-keto-L-gulonic acid
represents an extreme example for this case (figure 8): there
was barely any reaction below ca. 0.6 V. Above this potential the CO and Cx Hy Oz type poisoning species were
(partially) oxidized and some Pt sites were available for
the oxidation of L-sorbose [79]. The presence of poisoning
species and the possibility of their oxidative removal was
demonstrated by cyclic voltammetry of model Pt catalysts
in the same aqueous alkaline solution.
The electrochemical polarization of Bi-promoted Pt in
weakly basic aqueous solutions revealed that the Bi adatoms

Figure 8. Potential of a 0.75 wt% Bi-5 wt% Pt/alumina catalyst during


the oxidation of L-sorbose () and diphenyl carbinol () with air. The
potentials are related to RHE. Conditions for L-sorbose: aqueous Na2 CO3 ,
pH = 7.3, 50 C, 1 bar [79]. Conditions for diphenyl carbinol: 1 wt%
aqueous Li2 CO3 , dodecylbenzenesulfonic acid Na salt detergent, 75 C,
1 bar [37].

on Pt were in a reduced state below 0.60.7 V, but above


this potential they were successively oxidized. The stability of oxidized surface metal atoms and their resistance
towards dissolution depend mainly on kinetic factors. During L-sorbose oxidation the catalyst potential was in the
range of 0.70.85 V (figure 8) and most of the oxidized Bi
adatoms dissolved already at the beginning of the reaction,
as confirmed by XPS and ICP-AES methods [79]. The
rapid leaching of the promoter was favored by the product
possessing excellent chelating properties. On the contrary,
in the oxidation of diphenyl carbinol the bimetallic catalyst
was in an oxidized state (>0.7 V) only at the end of the
reaction when the alcohol was consumed (figure 8). In this
reaction and in many other facile oxidations of secondary
alcohols to ketones the catalyst was stable and no leaching
could be detected.
4.2. Control of the rate of oxidation
A generalized correlation between the rate of alcohol
oxidation and the potential of Pt or Pd during reaction is
shown in figure 9. This bell-shaped curve was found to be
valid for the oxidation of several primary and secondary
alcohols. The calculated standard potentials for simple alcohol carbonyl compound transformations are around
0.1 V. Accordingly, the catalyst potential at moderate conversion of these alcohols and at low oxygen concentration
(mixed potential, figure 3) can be as low as 0.10.2 V. In

122

T. Mallat, A. Baiker / Catalyst potential measurement


Table 2
Controlled oxidation of diphenyl carbinol in a nitrogenoxygen mixture.
Conditions: 0.75 wt% Bi-5 wt% Pt/alumina, 1 wt% aqueous Li2 CO3 ,
pH = 11, dodecylbenzenesulfonic acid Na salt detergent, 75 C, 1 bar
total pressure [37].

Poxygen
(bar)
0.05
0.21
1.0
1 0.05
a
b

Ecat a
(mV)

Time
(h)

Conversionb
(%)

80180
100320
630910
630800

5.5
2.5
8.0
1.0

99
99
98
99

Catalyst potential between 5 and 90% conversion.


Selectivity close to 100%.

Figure 9. General correlation between the relative rate of alcohol oxidation


and the catalyst potential.

some other reactions (e.g., sorbose oxidation) the lower potential limit is determined by the oxidative removal of surface impurities formed in the initial destructive adsorption
of the reactant, and the rate of alcohol oxidation enhances
only above 0.6 V [80,81].
The upper limit (ca. 0.80.9 V) is determined by the
availability of free Pt0 active sites. High catalyst potentials
are measured at high surface oxygen concentration and low
alcohol/carbonyl compound ratio (high conversion). As the
oxygen concentration increases, more active sites are covered by the oxidizing species and less free sites are available for alcohol adsorption. Accordingly, the observed reaction rate decreases rapidly above the optimum (figure 9).
The phenomenon that high oxygen concentration can deactivate the catalyst is termed over-oxidation in the literature [70,71]. The in situ measurement of catalyst potential
provided the first direct evidence for this early observation
[75,76].
From an engineering point of view it is important to
carry out the reaction in the potential range were the reaction rate is the highest. This can be achieved by the proper
tuning of the rate of oxygen supply. An example is shown
in table 2. When the oxygen concentration was set low
(0.05 bar oxygen partial pressure), the catalyst potential
was low indicating the reduced state of metallic sites, and
the reaction time was long. High oxygen pressure (1 bar)
resulted in a relatively positive catalyst potential. Under
these conditions the rate at low and medium conversions
was very high (not shown here). However, with increasing
conversion the rate of oxygen supply from the gas phase
to the surface sites became higher than the rate of alcohol oxidation. The catalyst was successively oxidized and
deactivated, resulting in the longest reaction time necessary for complete conversion. The shortest reaction time
was achieved when the rate of oxygen supply was properly
tuned to the actual rate of alcohol oxidation on the basis of
the catalyst potential. High oxygen concentration was applied at low conversion. The oxygen partial pressure was

Figure 10. Catalyst potential during the partial oxidation of cinnamyl


alcohol to cinnamaldehyde with air. Conditions: 5 wt% Pt/alumina or
0.75 wt% Bi-5 wt% Pt/alumina catalysts, 4 wt% aqueous Li2 CO3 , pH =
11, dodecylbenzenesulfonic acid Na salt detergent, 40 C, 1 bar [78].

decreased with increasing conversion so that the potential


remained in the range of 0.60.8 V, maintaining by this way
the balance between the oxidation and reduction processes
on the metal surface.
4.3. Control of selectivity based on catalyst potential
measurement
In the partial oxidation of cinnamyl alcohol to cinnamaldehyde it was found that promotion of a Pt/alumina
catalyst by selective deposition of Bi onto the surface of Pt
particles enhanced dramatically the reaction rate and aldehyde selectivity [38,78]. The rate enhancement was explained by an ensemble effect: the decreasing size of Pt site
ensembles suppressed the destructive adsorption of the reactant to a larger extent than the alcohol oxidation reaction.
Catalyst potential measurements provided a feasible explanation to the improved selectivity, as shown in figure 10.
Both promoted and unpromoted catalysts were in a partially oxidized state during reaction, but at comparable con-

T. Mallat, A. Baiker / Catalyst potential measurement

versions the potential of the Bi-promoted Pt/alumina was


lower by about 0.4 V. (The conversions after 5 h were 10
and 96% for the unpromoted and Bi-promoted Pt/alumina,
respectively.) It was found in an independent experiment
that the oxidation of the intermediate aldehyde to the corresponding acid required at least 0.7 V, under otherwise
identical conditions. The potential of BiPt/alumina was
below this limit up to ca. 90% conversion [38,78], which
explains the excellent performance of this catalyst: 99%
selectivity to cinnamaldehyde at 10% conversion, which
decreased below 97% only at 99% conversion. On the
contrary, the potential of Pt/alumina was above the 0.7 V
limit after a few percent conversion. After 5 h reaction
time the selectivity was 88.5% at only 10% conversion
and the selectivity dropped further with increasing conversion.

5. Conclusions
The steady state potential of (supported) metal catalysts
during redox reactions reflects the changes in the relative surface concentrations of reactant, intermediates and/or
products. Accordingly, variation of any reaction parameter
and the catalyst properties can shift the observed potential,
making the interpretation of the data compelling. Provided
that the reaction conditions are appropriately selected, catalyst potential measurements offer the opportunity to gain
valuable, on-time information on the catalyst working in
the liquid phase, which is hardly available with any other
analytical method.

References
[1] F. Beck, Chem. Ing. Tech. 48 (1976) 1096.
[2] G. Horanyi, Catal. Today 19 (1994) 285.
[3] E. Muller and K. Schwabe, Z. Elektrochem. Angew. Phys. Chem.
34 (1928) 179.
[4] E. Muller and K. Schwabe, Z. Elektrochem. Angew. Phys. Chem.
35 (1929) 165.
[5] E. Muller and K. Schwabe, Kolloid Z. 52 (1930) 163.
[6] K. Schwabe, Z. Elektrochem., Ber. Bunsenges. Phys. Chem. 61
(1957) 744.
[7] D.V. Sokolskii and V.A. Druz, Dokl. Akad. Nauk SSSR 73 (1950)
949.
[8] D.V. Sokolskii and V.A. Druz, Zh. Fiz. Khim. 26 (1952) 364.
[9] D.V. Sokolskii, ed., Catalytic Hydrogenation and Oxidation (Akad.
Nauk Kaz. SSR, Alma Ata, 1955) (in Russian).
[10] D.V. Sokolskii, G.D. Zakumbaeva, N.A. Zakarina, E.I. Ten and
B.Z. Aldamzharova, Akad. Nauk. Kaz. SSR, Tr. Inst. Khim. Nauk
14 (1966) 226.
[11] D.V. Sokolskii and A.M. Sokolskaya, Hydrogenation Metal Catalysts (Nauka, Kaz. SSR, Alma Ata, 1970) (in Russian).
[12] G.D. Zakumbaeva, N.A. Zakarina, L.A. Beketaeva and V.A. Naidin,
Metal Catalysts (Nauka, Kaz. SSR, Alma Ata, 1982) (in Russian).
[13] F. Beck, German Patent No. 1205539 (1963).
[14] F. Beck, Ber. Bunsenges. 69 (1965) 199.
[15] H. Kinza, Z. Phys. Chem. Leipzig 255 (1974) 180.

123

[16] H. Kinza, Z. Phys. Chem. Leipzig 255 (1974) 517.


[17] P. Boutry, O. Bloch and J.C. Balaceanu, Compt. Rend. 254 (1962)
2583.
[18] J. Held and H. Gerischer, Ber. Bunsenges. 67 (1963) 921.
[19] H. Gerischer, Ber. Bunsenges. 67 (1963) 164.
[20] K. Schwabe and A. Satsco, J. Electroanal. Chem. 11 (1966) 308.
[21] A.B. Fasman, D.V. Sokolskii and K.A. Shurov, Dokl. Akad. Nauk
SSSR 153 (1966) 653.
[22] D.N. Baria and H.M. Hulburt, J. Electrochem. Soc. 120 (1973) 1333.
[23] K. Heyns and H. Paulsen, Angew. Chem. 69 (1957) 600.
Polyanszky and T. Mathe, in: Hydrogen Effects
[24] J. Petro, T. Mallat, E.
in Catalysis, eds. Z. Paal and P.G. Menon (Dekker, New York, 1988)
p. 225.
[25] C.G. Vayenas and S.G. Neophytides, Catalysis 12 (1996) 199.
[26] I.S. Mietcalfe, Catalysis 13 (1997) 1.
[27] C.G. Vayenas and I.V. Yentekakis, in: Handbook of Heterogeneous
Catalysis, Vol. 3, eds. G. Ertl, H. Knozinger and J. Weitkamp (VCH,
Poitiers, 1997) p. 1310.
[28] J.F. van der Plas, E. Barendrecht and H. Zeilmaker, Electrochim.
Acta 25 (1980) 1471.
[29] B. Kastening, W. Schiel and M. Henschel, J. Electroanal. Chem. 191
(1985) 311.
[30] E. Keren and A. Soffer, J. Electroanal. Chem. 44 (1973) 53.
[31] B. Kastening, Ber. Bunsenges. Phys. Chem. 92 (1988) 1399.
[32] J. Held and H. Gerischer, Ber. Bundesges. Phys. Chem. 67 (1963)
164.
[33] G. Svehla, Automatic Potentiometric Titrations (Pergamon, Oxford,
1978).
[34] Y. Podvyazkin and A.I. Shligin, Zh. Fiz. Khim. (Russ.) 31 (1957)
1305.
[35] J.F. van der Plas and E. Barendrecht, Electrochim. Acta 25 (1980)
1477.
[36] S.A. Roberto and R.L. Augustine, in: Catalysis of Organic Reactions, ed. R.E. Malz (Dekker, New York, 1996) p. 465.
[37] T. Mallat, Z. Bodnar and A. Baiker, Stud. Surf. Sci. Catal. 78 (1993)
377.
[38] T. Mallat, Z. Bodnar, P. Hug and A. Baiker, J. Catal. 153 (1995)
131.
[39] G. Heydecke and F. Beck, Ber. Bunsenges. Phys. Chem. 92 (1988)
1418.
[40] D.V. Sokolskii, Hydrogenation in Solutions (Akad. Nauk Kaz. SSR,
Alma Ata, 1962) (in Russian).
[41] B. Kastening and S. Spinzig, J. Electroanal. Chem. 214 (1986) 295.
[42] I. Telcs, S. Szabo and F. Nagy, Magy. Kem. Foly. 71 (1965) 468.
[43] S. Szabo, I. Telcs and F. Nagy, Acta Chim. Hung. 73 (1972) 213.
[44] I. Paseka, J. Catal. 121 (1990) 349.
[45] Z. Bodnar, T. Mallat, I. Bakos, S. Szabo, Z. Zsoldos and Z. Schay,
Appl. Catal. 102 (1993) 105.
[46] Z. Bodnar, T. Mallat and A. Baiker, in: Catalysis of Organic Reactions, eds. M.G. Scaros and M.L. Prunier (Dekker, New York, 1995)
p. 393.
[47] G.D. Zakumbaeva and D.V. Sokolskii, Akad. Nauk Kaz. SSR, Tr.
Inst. Khim. Nauk 14 (1966) 32.
[48] N.A. Zakarina and G.D. Zakumbaeva, Highly Dispersed Metal Catalysts (Nauka, Kaz. SSR, Alma Ata, 1987) (in Russian).
[49] J.W. Jenkins, Plat. Met. Rev. 28 (1984) 98.
[50] C. Wagner and W. Traud, Z. Electrochem. Angew. Phys. Chem. 44
(1938) 52.
[51] M. Spiro, J. Chem. Soc. Faraday Trans. I 75 (1979) 1507.
[52] M. Paunovic, Plating (1968) 1161.
[53] M. Spiro, Chem. Soc. Rev. 15 (1986) 141.
[54] M. Paunovic, in: Proceedings of the Symposium on Electroless Deposition of Metals and Alloys, Vol. 88-12, eds. Paunovic and I. Ohno
(Electrochem. Soc., 1988) p. 3.
[55] M. Spiro, Catal. Today 17 (1993) 517.
[56] A.M. Creeth and M. Spiro, J. Electroanal. Chem. 312 (1991) 165.
[57] R.O. Farchmin, U. Nickel and M. Spiro, J. Chem. Soc. Faraday
Trans. I 89 (1993) 229.

124

T. Mallat, A. Baiker / Catalyst potential measurement

[58] L. Jelemensky, B.F.M. Kuster and G.B. Marin, Ind. Eng. Chem. Res.
36 (1997) 3065.
[59] P. Baumeister, H.U. Blaser and W. Scherrer, Stud. Surf. Sci. Catal.
59 (1991) 321.
[60] C.G.M. van de Moesdijk, in: Catalysis of Organic Reactions, ed.
J.R. Kosak (Dekker, New York, 1984) p. 379.
[61] I. Bakos, S. Szabo, F. Nagy, T. Mallat and Z. Bodnar, J. Electroanal.
Chem. 309 (1991) 293.
[62] S. Szabo, Int. Rev. Phys. Chem. 10 (1991) 207.
[63] Z. Bodnar, T. Mallat and A. Baiker, Catal. Lett. 26 (1994) 61.
[64] T. Mallat, Z. Bodnar, B. Minder, K. Borszeky and A. Baiker, J. Catal.
168 (1997) 183.
[65] R. Parsons and T. VanderNoot, J. Electroanal. Chem. 257 (1988) 9.
[66] B. Beden, J.M. Leger and L. C., in: Modern Aspects of Electrochemistry, Vol. 22, eds. O.M. Bockris, B.E. Conway and R.E. White
(Plenum, New York, 1992) p. 97.
[67] T. Iwasita and F.C. Nart, Adv. Electrochem. Sci. Eng. 4 (1995) 123.
[68] E. Pastor, S. Wasmus, T. Iwasita, M.C. Arevalo, S. Gonz`alez and
A.J. Arvia, J. Electroanal. Chem. 350 (1993) 97.
[69] K. Heyns and H. Paulsen, Adv. Carbohydr. Chem. 17 (1962) 169.
[70] H. van Bekkum, in: Carbohydrates as Organic Raw Materials, ed.
F.W. Lichtenthaler (VCH, Weinheim, 1990) p. 289.

[71] T. Mallat and A. Baiker, Catal. Today 19 (1994) 247.


[72] T. Mallat and A. Baiker, Catal. Today 24 (1995) 143.
[73] R. Woods, in: Electroanalytical Chemistry, Vol. 9, ed. A.J. Bard
(Dekker, New York, 1976) p. 1.
[74] Y. Schuurman, B.F.M. Kuster, K. van der Wiele and G.B. Marin,
Appl. Catal. 89 (1992) 31.
[75] T. Mallat, Z. Bodnar and A. Baiker, in: Proc. DGNK Conf. Selective
Oxidations in Petrochemistry, DGMK Tagungsbericht 9204 (Goslar,
1992) p. 237.
[76] T. Mallat, Z. Bodnar and A. Baiker, in: Catalytic Selective Oxidation,
Vol. 523, eds. S.T. Oyama and J.W. Hightower (ACS, Washington,
1993) p. 308.
[77] T. Mallat, Z. Bodnar, A. Baiker, O. Greis, H. Strubig and A. Reller,
J. Catal. 142 (1993) 237.
[78] T. Mallat, Z. Bodnar, M. Maciejewski and A. Baiker, Stud. Surf.
Sci. Catal. 82 (1994) 561.
[79] C. Bronnimann, Z. Bodnar, P. Hug, T. Mallat and A. Baiker, J. Catal.
150 (1994) 199.
[80] C. Bronnimann, Z. Bodnar, R. Aeschimann, T. Mallat and A. Baiker,
J. Catal. 161 (1996) 720.
[81] T. Mallat, C. Bronnimann and A. Baiker, Appl. Catal. A 149 (1997)
103.

Anda mungkin juga menyukai