You are on page 1of 16

feature

Supercritical fluids
opportunities in heterogeneous catalysis
The use of supercritical fluids (SCFs) in technology and research has been

abbreviations

triggered by several unique features of these media. SCFs allow improvement


D diffusion coefficient

of technical processes such as extraction, separation, particle formation and

DEA N,N diethylamine

chemical reactions. However, it has been just in the last decades that SCFs

DEF N,N diethylformamide

have gained their importance in science and industrial applications, due to the

DMA N,N dimethylamine


DMF N,N dimethylformamide

Hvap enthalpy of vaporization


k x mole fractionbased
rate constant
K x mole fractionbased
equilibrium constant
ln natural logarithm
MF methylformate

Opportunities provided by the use of supercritical media in chemical reaction


design have been the focus of several reviews[117]. Here, we concentrate on a
discussion of opportunities that SCFs provide in heterogeneous catalysis. First,
we will consider the intrinsic properties of SCFs, then discuss the effect of these
reactions. Finally, we illustrate some of the opportunities of SCF applications

pc critical pressure

in heterogeneously catalyzed reactions, using pertinent case studies.

R gas constant,
8.3145 J K mol

the rigorous demands on the equipment used.

properties on chemical reactions, especially focusing on heterogeneously catalyzed

p pressure

relatively high pressures usually required when working with these media and

RESS rapid expansion of

Roland Wandeler and Alfons Baiker

(corresponding author)

supercritical solution

Laboratory of Technical Chemistry, Swiss Federal Institute of Technology, ETH Zentrum

(technique)

CH-8092 Zurich, Switzerland

SAS supercritical antisolvent

E-mail: baiker@tech.chem.ethz.ch

(technique)
sc supercritical

Figure 1 Isochoric transition of CO2

(with compound)

from the twophase region to the super-

SCF supercritical fluid

critical state. All pictures have been taken

T temperature

with the computercontrolled autoclave

Tc critical temperature

system for the investigation of hetero-

TOF turnover frequency

geneously catalyzed reactions in SCFs

V volume

Vr
V

depicted in Fig. 8. The snapshots show

reaction volume

the reactor content, and the shaft of the

activation volume

(suspended) mechanical stirrer. Since the

dynamic viscosity

covered window area of the two CCD

isothermal compressibility

cameras used to record the images over-

kinematic viscosity

lap to a certain degree, the phase border

density

is shown twice, i.e., from the side by the

critical density

lower camera and slightly from the top by


the upper camera. As the temperature of
the system is increased, the density of the

gas phase increases while the liquid becomes less dense. This can be directly observed by
the phase border becoming more and more blurred and turbulent as the system approaches
its critical point. At the critical point, the densities of gaseous and liquid CO2 become equal

T = 19.5 C
p = 5.7 MPa

and the phase border disappears. The critical parameters of the system are slightly higher
than that of pure CO2 due to traces of air in the system.

128
Volume 4, no. 2, 2000

T = 24.9 C
p = 6.4 MPa

T = 32.1 C
p = 7.5 MPa

T = 33.0 C
p = 7.6 MPa

Properties of SCFs
The properties of SCFs are often said to lie between those of
a liquid and a gas. Although this is generally true it is important to note that some of the properties of an SCF are more
liquidlike, whereas others are more gaslike. Furthermore,
most properties of SCFs can significantly change with relatively small changes in pressure and temperature, rendering
SCFs easily tunable with respect to their qualities as a solvent and/or reactant. It is this mixture of more liquid and
more gaslike properties together with the easy tunability of
those properties that render SCFs such promising media for
chemical reactions and extractions.

Supercritical fluids

What is an SCF?
In 1822, Baron Charles Cagniard de la Tour conducted
experiments that showed there is a critical temperature above
which a substance can only be in one fluid state instead of
either gaseous or liquid [18]. As a fluid in gas liquid equilibrium is heated and approaches its critical temperature, the
density of the gas phase increases while the liquid becomes
less dense. If the two densities become equal the liquid and
gaseous phase converge into a single one with no change in
local symmetry. This transition from a liquid gas equilibrium
to the supercritical state is shown in Figure 1 using the example of CO2. At the condition where a liquid and gas phase in
equilibrium converge into a single phase the coexistence line
of the liquid and gas has a critical point, defined by e.g. a critical temperature and pressure as shown in Figure 2.
As follows from the expression supercritical fluid, this
term refers to a fluid at conditions beyond its critical point.
Although this seems to clearly assign the supercritical state, a
problem arises in laying down the term beyond in that definition; while some definitions demand for the temperature
and pressure of a fluid to exceed their critical values in order
to be supercritical, others omit all restrictions on pressure
and only require that the temperature be greater than Tc[19].
In this article, we will use the latter definition of the supercritical state for reasons of convenience: a fluid is called supercritical if the temperature exceeds its critical value (and the
pressure is below the one required to condense it into a solid);
or, in other words, if it cant be transformed into two nonsolid
phases by isothermally expanding or compressing it. With this
definition, a supercritical fluid reveals the unique property of
being tunable in density from gaslike values to liquidlike
ones without a phase transition.

SCF
solid

liquid

CP

TP

two phase region

gas

Tc
TTP
V

Figure 2 (above) Phase diagram of a pure compound in p,V,Tspace showing solid, liquid, gaseous (vapor), and supercritical domains as well as corresponding phase borders and phase coexistence areas. TP and CP indicate the triple
point and critical point of the compound, respectively; corresponding isotherms are
traced with thick lines.

T = 33.6 C
p = 7.7 MPa

T = 34.6 C
p = 7.7 MPa
129

10000
1000

density
/ kg m3

Supercritical fluids

gas a)

100
10
1
0.1
100

diffusion coefficient
D / 106 m2 s1

10
1
0.1
0.01
0.001

dynamic viscosity
/ mPa s

0.0001
10
1
0.1
0.01

kinematic viscosity b)
/ 106 m2 s1

1000
100
10
1

liquid a)

Density and related properties


the liquidlike qualities of an SCF

As stated above, the densities of a liquid and a gas


in equilibrium become equal at their critical point. Consequently, the density of an SCF at its critical point lies
between values typical for the liquid and gaseous state. However, regarding the order of magnitude, an SCF reveals densities typical of a liquid (between 100 and 1150 kg m3, see
Figure 3) and hence its characteristic dissolving power.
The density of a fluid in the supercritical region can be
continuously adjusted from liquidlike to gaslike values, as
shown in Figure 4. In the vicinity of the critical point a small
change in pressure can result in a sharp change in the fluids
density. With increasing distance from the critical point, the
effect of pressure on density becomes less dramatic. In regard
to tunability of densityrelated properties of an SCF the zone
immediately above the critical point is thus most effective for
changes with minimal variations in pressure and/or temperature. The density of an SCF increases with increasing pressure and decreases with increasing temperature.
The dissolving power of a given fluid depends on its
density[20]; while a liquid reveals great dissolving power and
a gas is a poor solvent, the dissolving power of an SCF
can be altered over a wide range by changing the pressure
and/or temperature. Generally, the solubility of a given solute
increases with both increasing density of the solvent and
increasing vapor pressure of the solute. Consequently, the
solubility of a solute in an SCF increases with increasing
pressure, whereas it decreases, remains constant or increases
with increasing temperature, depending on the predominant
factor, i.e. density of the solvent or vapor pressure of the
solute. Note that solubilities in liquids usually surpass those
in SCFs and that the dissolving power of an SCF only
approaches that of a liquid solvent at high enough densities.
However, it is the unique combination of relatively high
solubilities and other, gaslike features that render SCFs a
medium occasionally referred to as supersolvent.
The dielectric constant also shows a behavior parallel to
that of density to some extent. It increases sharply with pressure in the region near the critical point. Note that the magnitude of this increase depends on the nature of the SCF.
Viscosity, diffusivity and surface tension
the gaslike qualities of an SCF

The hydrodynamic properties of an SCF also lie between


those of a liquid and a gas (see Fig. 3), imparting favorable
qualities for reactions and extractions to the SCF in that
respect.
Figure 3 Comparison of physical properties of liquids, supercritical fluids
(SCFs) and gases in the near critical region. a) at ambient conditions; b) kinematic

0.1
0.01

130

SCF

viscosity was estimated from dynamic viscosity and density, = / . (Data taken
from [57] , with the exception of the density range of SCFs which includes c (Xe) =
1110 kg m-3[81] in this figure.)

Supercritical fluids

T3 >> Tc
T2 > Tc
T1 < Tc
p

p
p
p

CP
T

TP

Figure 4 Effect of pressure on density at subcritical (T1 < Tc) and supercritical (T2 > Tc, T3 >> Tc) conditions. Isotherm T1 illustrates the discontinuity in the density vs. pressure function at subcritical conditions due to the phase change. Isotherms T2 and T3 typify the continuous change from gaslike to liquidlike densities with increasing pressure.
Note that the effect of pressure on density for an SCF, in terms of change in density with a small variation in pressure, is more pronounced near the critical point.

The dynamic viscosity of an SCF is rather comparable


to that of a gas at the same temperature less so than a liquid.
However, the density of an SCF near its critical point is at
least two orders of magnitude higher than that of a gas,
resulting in a very low kinematic viscosity = / . This
is very advantageous for mass transfer, since natural convection effects are inversely proportional to the square of the
kinematic viscosity [21].
Diffusivities of SCFs can also be considered as being gas
like with respect to their order of magnitude. Consequently,
diffusion coefficients of solutes in SCFs are gaslike too,
adding to the attractiveness of SCFs with regard to processes
where diffusion of a component constitutes a limiting step.
Viscosity and diffusivity of an SCF are dependent on
temperature and pressure, as is the case for density. In an
SCF, viscosity increases and diffusivity decreases as pressure
is increased, approaching values of a liquid. An increase
in temperature leads to a decrease in viscosity of the SCF
(note that the opposite is true in the case of a gas), whereas
diffusivity will increase with an increase in temperature.
The surface tension of a liquid that is in equilibrium
with its vapor decreases with increasing temperature and
becomes zero at its critical point. Therefore an SCF acts like
a gas in uniformly filling up the space available, allowing
SCFs to penetrate readily porous solids and packed beds.
Tunability of properties the unique quality of SCFs

In addition to the very special combination of gaslike and


liquidlike properties, the easy tunability of these properties along with relatively small changes in pressure and/or

temperature render SCFs very promising media for chemical


purposes.
The hydrodynamic properties of an SCF can be tuned
over a wide range by changing temperature and/or pressure
of the fluid. In addition to that, the density of an SCF can
be continuously adjusted from liquidlike to gaslike and so
can all the density related properties of the SCF as for example solvent power. Particularly in the region about the critical point large changes in fluid density and related properties
are observed with small changes in pressure. Accordingly,
in regard to tunability of the SCFs properties the region
around the critical point, the socalled near critical region
of the fluid, holds the greatest interest for research and process engineering.
Properties at or near the critical point
the special quality of SCFs

The behavior of SCFs discussed so far relates to the


supercritical region in general. However, systems near their
critical point exhibit some very special properties, due to
the nature of the critical point being the endpoint in the
coexistence line of gas and liquid, and corresponding phase
transitions being of second order.
The region where a liquid and gas phase can coexist
becomes smaller with increasing temperature and is reduced
to a single point at the critical temperature: the critical point
of the fluid. With the unique nature of the critical point
come a number of peculiarities:
The difference between liquid and vapor suddenly disappears at the critical point.
131

Supercritical fluids

A number of physical properties become zero, (e.g.


enthalpy of vaporization Hvap), whereas others diverge
theoretically (e.g. isothermal compressibility T). Some
of these anomalies can be directly derived from the
p,V,Tdiagram (see Fig. 2) where the Tcisotherm in
the p,Vplane reveals a point of inflection with a horizontal tangent at the critical point. Consequently, the
first and second partial derivatives of temperature at this
point are both zero, unequivocally defining the critical
point. However, since the isothermal compressibility T
becomes infinite in a secondorder phase transition, as it
is the case at the critical point, density fluctuations and
resulting effects have to be considered as for example in
the fluctuation dissipation theorem in statistical mechanics. It is supposed that the behavior of various measurable quantities near the critical point can be decomposed
into a regular part remaining finite and a singular
part that may be divergent; the divergent part is found
to be proportional to |(T Tc) Tc1|i where i, one of the
so called critical exponents (, , , , , ), defines the
nature of the singularity (note that in experiments even
widely different systems share approximately the same
critical exponents) [22,23]. This superposition of a regular
and a singular term describes the real behavior of SCFs
quite well. It explains the deviation of properties in the
nearcritical region (including the liquid region near the
critical point) whereby properties may deviate strongly
or weakly depending on the corresponding path on the
p,V,Tsurface [23].
The fluid shows the phenomenon of critical opalescence,
a characteristic signature of the phase transition at the
critical point. As stated above, the large isothermal compressibility allows for density fluctuations at low cost
of free energy. Since the range of those density fluctuations can become comparable to the wavelength of
visible light, light passing through an SCF near the
critical point will be scattered and may be partly or
fully absorbed, resulting in observation of the socalled
cloud under these conditions.
These peculiarities just give a hint that the region around
the critical point possesses some very special properties, in
comparison to the supercritical region in general, which
might be of particular interest in researching chemical processes in SCFs.
Other properties of SCFs

In addition to their promising physicochemical properties,


SCFs may also provide very favorable qualities with regard to
ecology and economy: the ideal SCF for industrial applications
is nontoxic, environmentally benign, nonflammable and
available in high purity at low cost; it is gaseous under
ambient conditions and has moderate critical conditions to
facilitate process design. It is very convenient that there are
SCFs that meet all or most of these criteria in addition to
their very interesting supercritical features.
132

SCFs most commonly used in chemistry

The majority of studies involving SCFs have focused on


CO2, ethene, ethane and water[24]. They all exhibit very
interesting qualities with regard to their physicochemical
properties as well as ecology and economy.
Carbon dioxide (Tc = 31.0 C, pc = 73.8 bar) is by far the
most widely used SCF in chemistry and in extraction processes. It is used in high purity or with organic modifiers to
increase fluid polarity. CO2 based fluids are nonflammable,
nontoxic and environmentally benign; they can be vented
to the atmosphere or recycled without harm. Furthermore,
CO2 is available in a high state of purity at comparably low
cost. It is a moderately good solvent for lipophilic organics
of lowtomedium molecular weight and not very reactive,
which is an advantage when using CO2 as a solvent but
demanding on the catalyst when using it as a reactant.
Supercritical ethane (Tc = 32.3 C, pc = 48.8 bar) and
ethene (Tc = 9.2 C, pc = 50.5 bar) show marginally higher solvation power than CO2. In contrast to CO2 both are flammable, and ethene can even polymerize, asking for corresponding precautions when working with these hydrocarbons.
Water, too, is cheap, nontoxic, nonflammable, and ecologically benign. However, while CO2, ethane and ethene
all have moderate critical parameters, supercritical water
requires relatively high temperatures and pressures in the
nearcritical region (Tc = 374.0 C, pc = 220.6 bar). Furthermore, scH 2O is highly corrosive and demands for special alloys for all contacting parts. Nonetheless, the very special properties of water in the supercritical region make it an
ideal solvent for the oxidation of organic wastes (supercritical water oxidation, SCWO) [12]. Supercritical water behaves
like a moderately polar organic liquid in the vicinity of the
critical point. The high miscibility of scH 2O with organics
and oxygen leads to the formation of simple and nontoxic
compounds such as CO2 and H 2O. This renders scH 2O a
valuable medium for the decomposition of substances that
are rather difficult to process otherwise, despite the high
cost of reactors.
Effects in chemical reactions under
supercritical conditions
The unique properties of SCFs and the high pressures
usually applied in chemical reactions under supercritical
conditions can lead to a change in the key parameters
of chemical reactions in SCFs as compared to processes
conducted at ambient pressure. A number of effects not
playing a significant role in classical chemistry have to
be considered in SCF processes, such as effect of pressure
on reaction thermodynamics and kinetics, clustering due to
local density enhancement, phase behavior of the system,
and enhanced mass and heat transfer.

Chemical reactions under supercritical conditions usually


require relatively high pressures due to the nature of the
supercritical state and the kind of fluids commonly used.
Consequently, pressure effects on chemical equilibria and
chemical reaction rates [25] have to be accounted for.
The effect of pressure on the mole fractionbased equilibrium constant K x of a chemical reaction depends on the
reaction volume Vr of a reaction, i.e. the difference between
the partial molar volumes of the product(s) and those of the
reactant(s):

lnK x
p

T,x

Vr
RT

(1)

Due to solvation effects and the high isothermal compressibility in SCFs, partial molar volumes in SCFs may be of
dramatic values near the critical point [26] and the resulting
reaction volume Vr may strongly alter the effect of pressure
on equilibrium with regard to the reaction carried out in a
liquid or gaseous phase.
The effect of pressure on kinetics is mostly described in
the context of transition state theory [27] and a supposed reaction pathway. According to this theory, the mole fraction
based rate constant k x of an elementary reaction depends on
the activation volume V , i.e. the difference between the
partial molar volume of the activated complex and the sum
of those of the reactant(s):

lnk x
p

T,x

V
RT

(2)

Supercritical fluids

Effect of pressure on thermodynamics and kinetics

Again, solvation effects and the high isothermal compressibility in SCFs may lead to dramatic values of partial
molar volumes in SCFs, and consequently to activation volumes V in the order of liters per mol. Consequently, the
effect of pressure on the rate constant in SCFs can be dramatic as compared to reactions in a liquid or gaseous phase,
where activation volumes are typically in the range of 0.03
to 0.03 liters per mol, or are even restricted to the intrinsic
activation volume, respectively [25]. However, this effect is
due in a large part to the high compressibility of SCFs and
therefore mainly reflects the change from a reaction rate in
a gaslike fluid to one in a liquidlike fluid. Consequently,
activation volumes V in SCFs are generally not useful to
extract mechanistic information about the reaction, in contrast to gas or liquid reactions, where the activation volume
more or less reflects the size change from the reactants to the
activated complex.
Local enhancement of density clustering
The high compressibility in SCFs and the gaslike behavior
with regard to surface tension allows attractive forces to
move molecules into energetically favorable locations. The
resulting nonuniform spatial distribution of solvent and
cosolvent molecules about solute molecules, as schematically
illustrated in Figure 5, gives rise to interesting solvent effects
not ordinarily found in liquid mixtures. This phenomenon,
which has been termed local density enhancement [28] , clustering [29] , or molecular charisma [30] , can affect the rates and
selectivities of chemical reactions through both physical and
chemical mechanisms [31].
Local anisotropy phenomena also may occur at supercritical fluidsolid interfaces as a consequence of different
interaction strength (adsorption enthalpy) of solute (reactant), solvent, and cosolvent, with the solid surface and clustering phenomena in the bulk phase near its critical point.
Fundamental knowledge of these interactions is important
to understand the mechanism of solid catalyzed reactions.
Unfortunately, there appears to be no fundamental work
dealing with this aspect crucial for understanding catalytic
surface reactions in solutesolvent (reactantsolvent) systems.

Figure 5 Schematic illustration of spatial distribution of molecules for a liquid,


supercritical and gaseous system. The upper left part of the circles show a pure
fluid; the lower right part also involves a volatile solute. Whereas liquid
p

and gaseous systems show a uniform spatial distribution of molecules,


the large isothermal compressibility of SCFs allows attractive forces
to move molecules into energetically favorable locations at low
cost of free energy and thus leads to density fluctuations and

CP
T

clustering phenomena. The range of those density fluctuations is


often comparable to the wavelength of visible light. Note that the

TP

extent of clusters for the SCF is shown in reduced scale for reasons of
V

clearness.
133

Supercritical fluids

Phase behavior and solubility

An interesting opportunity of using supercritical conditions


for catalytic reactions is that all reactants can exist in a
single homogeneous fluid phase. However, catalytic reactions being either homogeneous or heterogeneous will generally involve multicomponent systems for which the location of the phase border curves in pressuretemperature
composition p,T,xspace [32] may not be known. Furthermore, the composition of this multicomponent system will
change with reaction time (batch reactor, autoclave) or with
location (continuous fixed bed reactor) and so may the phase
behavior of the system at given reaction conditions. Knowledge of the phase behavior of the reaction system is a necessary prerequisite to make beneficial use of the supercritical state and for interpreting its effect on the rate and selectivity of a catalytic reaction. Since at this point there is no
general equation satisfactorily describing multicomponent
phase diagrams, experimental determination of the phase
behavior is indispensable.
In addition to the reaction conditions, solubilities
of reactant(s) and/or product(s) can be influenced by a
cosolvent which undergoes a strong specific interaction
(e.g. hydrogen bonding) with a solute. Although rendering
the reaction system even more complex with regard to
the number of molecules present in the reaction system,
employing one or more cosolvents provides an elegant tool
for tuning the phase behavior. Unfortunately only very limited knowledge is presently available to make proper use
of this potential tool. However, progress in understanding
the molecular interactions involved will undoubtedly lead
from empirical trial and error methods to rational design of
solutesolventcosolvent systems.
Mass and heat transfer

Transport properties in SCFs such as diffusivity, viscosity,


and thermal conductivity are very favorable for chemical
reactions as compared to liquid and gaseous media. This,
together with the elimination of possible interphases in
multicomponent systems and the high density of SCFs
with ensuing very low values of the kinematic viscosity
= / and strong natural convection, leads to a significantly enhanced mass and heat transfer under supercritical conditions. The available data on diffusivity [3335] is very
limited, but recent research is expanding the data base for
binary diffusion coefficients as well as exploring the influence on solute diffusivities.
Opportunities for application
of SCFs in heterogeneous catalysis
Conducting heterogeneously catalyzed reactions in supercritical media looks very promising due to the properties of
SCFs and their effect on various reaction parameters. Indeed,
proper use of SCFs in heterogeneous catalysis can eliminate
gas liquid phase transfer resistance, afford an enhancement
of the reaction rate, enhance mass and heat transfer, pro134

vide a further means for control of selectivity, increased catalyst lifetime and regeneration, or facilitate easier product separation and process intensification. Moreover, supercritical
fluids offer several unique opportunities in catalyst preparation. Some of the opportunities discussed here are valid
for chemical processes in general, whereas those concerning
the solid catalyst are unique for heterogeneously catalyzed
chemical reactions.
Elimination of gas liquid phase transfer resistances

The most obvious opportunity for the use of SCFs in heterogeneous catalysis is the elimination of gas liquid phase
transfer resistances in reactions involving three phases under
subcritical conditions (see Fig. 6). This together with lowered external fluid film diffusion resistances resulting from
lower viscosity of SCFs may significantly accelerate the reaction, since diffusion of the gaseous reactant to the catalyst
surface often represents the rate limiting step in threephase
reactions.
Enhancement of reaction rate

The effect of pressure on reaction rates in the supercritical


region can be assigned to the kinetic pressure effect,
enhanced mass and heat transfer as well as occasional higher
reactant solubilities. Since all these influences on the reaction rate strongly depend on pressure, temperature and the
fluid itself, reaction rates can be tuned by corresponding
adjustments in the reaction conditions.
Control of selectivity

The rate and equilibrium of a given reaction can be tuned in


SCFs by altering pressure and temperature and/or adding a
corresponding cosolvent. For a network of parallel or competing reactions, the different reactions may be influenced in a
different way and degree by these alterations in the reaction
conditions. Consequently such tuning of the reaction conditions may favor one of the reactions over the others, offering
some potential to enhance the selectivity to the desired product. However, this method of controlling selectivity is rather
demanding, because the activation and reaction volumes of
the various reactions as well as the influence of the components on the phase behavior of the system need to be known.
Further possibilities for the control of selectivity are
linked with tuning of solutesolvent interactions (change of
local density, clustering), and the use of cosolvents which
through specific interactions (usually hydrogen bonds) with
a transition state or a product can alter both rates [36] and
product distributions [37].
Enhanced mass and heat transfer

Fluids exhibit high diffusivities and very low kinematic viscosities in the supercritical region, resulting in high mass
and heat transfer rates. Consequently, working in the supercritical region may accelerate mass transfercontrolled liquid
reactions and lead to better heat removal in highly exother-

diffusion of gaseous reactant A from


bulk gas phase to the gas/liquid
interface

absorption of A at the gas/liquid


interface and ensuing diffusion to the
liquid bulk phase

diffusion of reactants from bulk liquid


phase through stagnant fluid film
surrounding solid catalyst particle

diffusion of reactants into porous


network of the catalyst

adsorption of reactant(s)

surface reaction

desorption of product(s)

diffusion of product(s) through porous


network to outer surface of the
catalyst

diffusion of product(s) through


boundary layer into bulk fluid

Supercritical fluids

Physical and chemical steps relevant


in a tree-phase reaction

Mass transfer resistances at subcritical conditions


c(A)
gas

= catalyst

liquid

catalyst

= gas bubble (reactant A)

= liquid (reactant B)

Magnified cross section of solid catayst

Mass transfer resistances at supercritical conditions

c(A)

SCF
5

catalyst

catalyst

bulk
phase

boundary
layer

= catalyst
= SCF (reactants A and B)

4
5
6

Figure 6 Sequence of physical and chemical steps occurring in a heterogeneously catalyzed gas liquid reaction (e.g. hydrogenation of a liquid compound) and comparison
of such a reaction at subcritical and supercritical conditions. A representative section of the reactor content, corresponding masstransfer boundaries and reactant concentration profile are shown for both sub and supercritical conditions. Note that gas liquid transfer resistance (steps 1 and 2) is eliminated and external fluid film diffusion resistance
(step 3) is lowered under supercritical conditions due to lower viscosity of SCFs.
135

Supercritical fluids

mic gas phase reactions, where careful temperature control is


essential for selectivity and product stability. In either case,
transport properties of SCFs are very favorable for conducting chemical reactions.
Catalyst lifetime and regeneration

A problem frequently encountered with heterogeneously catalyzed gas phase reactions is the formation of heavy organics
which may act as catalyst blocking agents and thereby deactivate catalysts and promote coking. Since supercritical fluids
exhibit considerably higher solubilities than corresponding
gases for such heavy organics, catalyst deactivation may be
suppressed by changing working conditions from gas phase
to dense supercritical medium [38,39]. Furthermore enhanced
diffusivity can accelerate the transfer of poisons from the
internal and external catalyst surface. The lifetime of a catalyst can therefore be significantly prolonged by working in
the dense supercritical region.
Regeneration of catalysts deactivated by coking can be
accomplished by extracting the carbonaceous deposits from
the catalyst surface. Due to the high solubilities and favorable
transport properties in the supercritical region, SCFs provide
a valuable instrument for such extraction processes.
Facilitated separation

The easy tunability of a fluids solvent properties in the


critical region is an efficient means for the separation of
one or more reactant(s) and/or product(s) in a reaction
system. Furthermore, the release of pressure after the reactor
will condense or precipitate remaining (unused) reactants
and/or products, allowing their reintroduction into the feed
(reactant recycling) and separation as well as a recycling of
the now gaseous reaction medium. Since most SCFs used in
practice are small and high volatile compounds, the precipitated product(s) are almost free of solvent traces and mostly
do not need further cleaning in that respect.
The tunability of a solutes solubility in an SCF is
adventageously used in the highpressure polymerization of
ethene: The reaction is conducted in a medium and under
conditions that polymers will fall out of the supercritical
solution when they reach a certain molecular weight, corresponding to the solubility limit. In an equilibriumlimited
reaction, this continuous removal of the product would
enhance conversion. Again, a slight release in pressure after
the reactor will precipitate unused reactant(s), allowing their
reintroduction into the feed and the recycling of the SCF.
In the case of a solid product, the tunable solvation properties of SCFs could be used not only for separation of
the product from the solvent, but also to tailor the size of
the product particles, a prerequisite feature of pharmaceuticals used for inhalation or ready oral resorption. Particle
forming techniques involving the application of supercritical fluids entail the rapid expansion of supercritical solutions
(RESS) [32,40] and the supercritical antisolvent (SAS) [4143]
technique. RESS is based on the rapid depressurization of
136

a supercritical fluid phase in which the solute(s) of interest


are dissolved, whereas in SAS the solute(s) are dissolved in a
liquid organic phase and precipitation is achieved by bringing this solution into contact with a supercritical fluid having
a low affinity for the solutes and a large mutual solubility
with the organic phase.
Process intensification

The use of SCFs in heterogeneously catalyzed reactions


can lead to significantly intensified chemical processes. The
high reaction rates, mass transfer and heat transfer in SCFs
allow the construction of continuous reactors, considerably
smaller than required for conventionally operated continuous
reactors of equal performance [44,45]. This opportunity provides
interesting advantages concerning process safety and space
requirement of chemical plants [46,47]. Furthermore, the facile
product separation and ecologically beneficial recycling of
solvent and/or reactant(s) may simplify the design of the
reaction process.
Catalyst preparation

Supercritical fluids provide unique opportunities in the


preparation of catalytic materials and supports. Proper use
of SCFs allows to tailor and optimize the morphology of a
catalyst.
In the preparation of aerogels [48] via the solution solgel
route, supercritical drying is imperative. If the liquid (solvent) entrapped in the tenuous solgel network is directly
evaporated, the structure of the gel is severely damaged due
to the acting capillary pressure when the liquid recedes into
the solgel body. This capillary stress can be circumvented
either by transferring the entrapped solvent from the liquid
to the supercritical state or by replacing the solvent typically
with supercritical CO2, thus eliminating any liquidvapor
interface inside the solgel product during solvent extraction. The solgel method combined with ensuing supercritical drying provides unique opportunities for the preparation
of mixed oxides and metal/metal oxide catalysts as extensively described in recent reviews [4850].
Another interesting opportunity for the use of SCFs in
catalyst preparation is the possible control of the particle
size and morphology of catalytic materials due to the highly
adjustable properties in the supercritical region with small
changes in temperature or pressure. Catalysts could also be
prepared using other particle forming techniques such as the
rapid expansion of supercritical solutions (RESS) and the
supercritical antisolvent (SAS) technique discussed earlier.
Although both techniques have not yet found significant
application in the preparation of catalytic materials, they
promise interesting opportunities for the control of particle
size and morphology.
Experimental laboratory techniques
Typically, reactions using SCFs are performed at elevated
pressure (up to 400 bar) and temperatures (up to 600C in

batch reactors

phases (e.g. spectroscopy or online sampling). In contrast


to homogeneous reactions, monitoring of heterogeneously
catalyzed reaction systems is impaired by the presence of
the suspended solid catalyst particles, necessitating a more
complex design of the analytical system. A highpressure
system specifically designed for in situ investigation of heterogeneously catalyzed reactions in supercritical fluids has
recently been reported [64]. This computercontrolled system
allows the investigation of changes in phase behavior and
composition by use of online video imaging as well as
online sampling and analysis by gas chromatography. A
flow chart of the system is given in Figure 8.
Spectroscopic studies have played a major role in
unraveling information on solvation and effects of local
molecular phenomena by probing molecular scale interactions. Together with theoretical and computational investigations they led to the ideas of enhanced local densities and
local compositions around solutes (reactants) in supercritical solutions [65]. Several spectroscopic techniques are amenable to highpressure investigations such as ultraviolet absorbance, fluorescence emission, infrared, electron spin resonance and nuclear magnetic resonance spectroscopy [6669].
However, in situ monitoring of surface reactions under
highpressure conditions is still in its infancy. In the progress of the fundamental understanding of supercritical fluid
solid (catalyst) interactions and surface reactions, in situ
monitoring of surface processes is probably the most difficult and most crucial hurdle to overcome.

continuous reactors

a)

Supercritical fluids

the case of SCWO). The use of such conditions is demanding on the experimental equipment used [51]. Furthermore,
the potential danger of these conditions should never be
ignored and full safety precautions should be made for all
experiments.
For heterogeneous catalytic reactions two principle reactor types are suitable, batch reactors (autoclaves) or continuous flow reactors. The advantages and disadvantages of the
two reactor types are well known in heterogeneous catalysis
and also apply to the use of SCFs. For industrial applications, however, the extremely good mass and heat transfer
properties of SCFs render tubular fixedbed reactors ideal
for heterogeneously catalyzed reactions, both on grounds of
safety (low reactor volume) and costs. Figure 7 schematically
shows the reactor types most frequently used for heterogeneous catalytic reactions at supercritical conditions.
A necessary requirement for designing heterogeneous catalytic reactions at near or supercritical conditions is knowledge of the phase behavior[28,5257]. Due to the limited predictability of theoretical calculation of phase behavior[58] experimental measurements are indispensable. Several approaches
for monitoring the phase behavior of homogeneous catalytic reactions have been reported and reviewed in the
literature [13,51,5963]. These systems are mainly based on a
highpressure cell equipped with a means to detect phase
boundaries (optically through a glass window, by conductometric analysis, or with ultrasound) and preferably an
additional instrumental channel for further analysis of the

d)
PM

c)
PM

b)
H

Figure 7 Reactor types suitable for the study of heterogeneously catalyzed reactions at supercritical conditions. a) stirred autoclave, b) stirred autoclave with internal
recycle, c) differential (gradientless) reactor, and d) continuous flow reactor. PM: premixing chamber; H: heat exchanger.
137

Supercritical fluids

Compressed Air

MV 2

MV 1

MV P
Rinsing

PI

NIC

RV A

PC

IS

MI
FI
TI

MV A
PI

RV B

PI

TIC
PV P

PC

NV 1

MV B

RM 1

PT

PV 1

TT
TI

FI
SV o

TI

TIC

el.
NV 2

RM 2

PV 2

kryo
TIC
KS

CCD o

el.

monochrome AV

kryo
monochrome AV

CCD u
AO

ER-8

TS

AI/RTD

TI

SC-2042

SC-2050

SV u

el.

SFA Software
SFA SubVIs
BV/IMAQ VIs

PCI

AT

Driver SubVIs

RS 232
DVC Recorder

BV Engine
Data Logging
Alarm Man.

MIO

BV Tag Server

AV

NI IMAQ

Human Machine Interface


SFA Applications (NI BridgeVIEW 2.0)

SFA Computer
IC

TC

HP GC 6890

Figure 8 Computercontrolled highpressure system for in situ monitoring of heterogeneously catalyzed reactions under supercritical conditions using online video
imaging as well as online sampling and analysis by gas chromatography. The figure shows a flow chart including the organization of process control and acquisition of data
and images. The hatched bar represents a wall separating the high pressure cabinet (upper part of the flow sheet) from the operating area (lower part of the flow sheet). RVx:
highpressure reduction valves; MVx: magnetic valves; NVx: needle valves; RMx: massflow meter; PVx: pneumatic valves; IS: stirrer; KS: kryostat; TS: thermostat; SVx:
heated sampling valve system; CCDx: 8bit monochrome CCDcameras. (Reproduced with permission from [64] .)

Case studies
Some of the opportunities discussed in the preceding sections are illustrated in the following case studies: the synthesis of formic acid derivatives from scCO2, and the amination
of diols in scNH3. A comprehensive review containing several other studies demonstrating the opportunities of SCFs
in heterogeneous catalysis has recently been published [1].
Synthesis of formic acid derivatives
from supercritical carbon dioxide

Beneficial application of scCO2 in catalyst preparation as


well as in reaction design has recently been demonstrated

a)

CO2

H2

H
N

for the syntheses of the formic acid derivatives such as


N,Ndimethylformamide (DMF), N,Ndiethylformamide
(DEF) and methyl formate (MF) (see Fig. 9). Industrially
these products are synthesized from carbon monoxide as
C1building unit, however stimulated by environmental considerations there is a growing interest to apply nontoxic,
highly abundant CO2 for these syntheses. Noyoris group [7072]
demonstrated that the application of scCO2 affords a far
more efficient process for the syntheses of DMF and MF
than the corresponding liquid phase syntheses using organic
solvents. In this reaction scCO2 serves as both a reactant and
a solvent. The higher efficiency of the scCO2 based reaction

Ru hybrid aerogel

R = CH 3 , CH2CH3
b)

CO2 +

H2

+ CH3OH

Ru hybrid aerogel

O
N
R

HCOOCH3

H 2O

+ H 2O

Figure 9 Synthesis of (a) formic acid derivatives such as N,N dimethylformamide (DMF), N,N diethylformamide (DEF) and (b) methyl formate (MF) from CO2, H2 and the
corresponding alkylamine and alkanol, respectively, using a ruthenium hybrid aerogel as catalyst.
138

a)

Ph2 Cl
P
Ru
P
Ph2 Cl

Ph2
P
P
Ph2

(EtO) 3 Si

(EtO) 3 Si

Ph Cl
P
Ru
P
Ph Cl

Supercritical fluids

b)

Si(OEt)3

Ph
P
P
Ph

Si(OEt)3

c)

Figure 10 Preparation of the ruthenium hybrid solgel precursor. A ruthenium phosphine complex (a) showing high activity and selectivity in the synthesis
of formic acid derivatives, is functionalyzed by silyl ether groups (b) and immobilized within a silica matrix affording aerogel or xerogel hybrid catalysts (c) by corresponding drying procedures.

design was attributed to favorable mass transfer effects, high


solubility of H 2 and weak solvation of the homogeneous catalyst RuCl 2 [P(CH3) 3] 4 in scCO2. Although this solventfree
reaction design was far more efficient than the corresponding liquid phase syntheses of DMF and MF, the low global
reaction rate and problems concerned with airstability and
separation of the ruthenium catalyst restricted its technical
utilization.
Intrigued by these results, we explored the opportunity
to develop corresponding heterogeneous catalytic processes
based on scCO2 [7375] , which avoid the disadvantages, connected with catalyst stability and separation inherent to the
homogeneous processes.
This aim was pursued in two steps. First a thorough
screening of the catalytic behavior of monodentate group
(VIII) transition metal complexes [76] confirmed that Ru
complexes are most favorable showing selectivities up to
100% to the desired products. A further striking improvement could be reached by applying bidentate type Ru complexes as the one shown in Figure 10a, which combine highest activity with good stability against air[77]. In a second
step the bidentate ruthenium complexes were functionalized
with silylether groups (Fig. 10b) which served to anchor the
highly active complexes by means of a solgel process in a
porous silica matrix (Fig. 10c) [74]. The pore size distribution
of the hybrid gel catalysts was found to be crucial for their
catalytic performance. Conventional evaporative drying of
the gels afforded microporous xerogels [75] with significantly
lower catalytic activity compared to the corresponding mesoporous aerogels which were prepared (dried) by extraction
with scCO2 at low temperature. In the latter procedure
the tenuous porous gel network remains largely preserved.
Figure 11 illustrates how the different drying processes
affect the pore size distribution of the final hybrid gels.
Both xero and aerogels afforded 100% selectivity to N,N
diethylformamide, and showed no sensitivity to air or water,
and no leaching under reaction conditions. The differences

HO
OH
O
Si
O O
HO
O
O OH O
O
Si
Si
O
O
O Si O Si O
HO Si
Si O O
Ph Cl Ph
OO
O
P
P
O Si O
O
Ru
O Si O
HO
P
P
OH
O Si
Si
Ph Cl Ph
O
Si O O
Si
O Si O
HO
Si O
O Si
O
O
OO O
Si O
Si
O
O
O O
O O
O Si
OO O

in the pore size distributions of the xero and aerogels have


a dramatic effect on their catalytic activity. The turnover
frequency (TOF) measured for the xerogel was 2210 h 1,
which was far exceeded by the aerogel with a TOF of 18400
h 1. Most striking, the activity of the aerogel also greatly
exceeded that of the corresponding bidentate complex in the
homogeneous reaction (TOF = 3130 h 1). This clearly shows
the importance of the textural properties of these hybrid catalysts which can be achieved by applying extraction of the
liquid entrapped in the pores using scCO2 at low temperature. In this drying process capillary stress detrimental to
the tenuous gelnetwork is suppressed.
In the synthesis of N,Ndiethylformamide carbon dioxide not only acts as a reactant but also strongly influences the
phase behavior of the system. As shown in Figure 12 for DEF
synthesis two fluid phases are present under the conditions
applied: a liquid phase at the bottom of the reactor, containing mainly DEA/DEF, dissolved CO2 and H 2 as well as the
suspended catalyst; and a supercritical phase at the top of the
reactor, containing mainly H 2 and CO2 with traces of DEA/
DEF. Working in the high density supercritical region of the
upper phase is favorable for the phase transport properties
and increases the concentration of CO2 and H 2 in the DEA
based phase containing the catalyst.

139

3.0

2.0

dV/dlog[d / nm] / cm3 g-1

Supercritical fluids

Aerogel

4.0

1.0

0.0
1.4

Xerogel

1.2
1.0
0.8
0.6
0.4
0.2
0.0

10

100

1000

Pore Diameter / nm

Figure 11 Comparison of differential pore size distribution of the ruthenium


hybrid solgel precursor after (a) drying with supercritical CO2 and (b) conventional evaporative drying yielding the corresponding aerogel and xerogel catalysts,
respectively. The curves were derived from nitrogen physisorption at 77 K. (Taken
from [74] .)
Amination of diols in supercritical ammonia

The heterogeneously catalyzed amination of alcohols is the


most important industrial process for the manufacture of a
variety of aliphatic and aromatic amines [78]. However, in the
amination of aliphatic diols to primary diamines yields and
selectivities are usually rather low, due to the higher reactivity

of the monoaminated intermediate of the reaction as compared to ammonia and the ensuing side reactions (see Fig. 13).
Application of highdense scNH3 as a solvent and reactant
significantly improves the selectivities to the corresponding
primary diamine in the amination of e.g. 1,3propanediol [79].
The corresponding experiments were conducted using a
CoFe catalyst in a tubular flow reactor at a temperature of
195 C with a molar ratio of 1,3propanediol : H 2 : NH3 =
1 : 2 : 40. As expected for a multicomponent system, the
reaction system exhibits a nontrivial phase behavior: experiments in a view cell showed that at the temperature applied
the reaction system was onephase and presumably supercritical at pressures > 90 bar, whereas at lower pressures
a liquid alcoholbased phase and a seemingly supercritical
ammoniaH 2 phase with traces of alcohol were present. As
shown in Figure 14, the pressure of the reaction system has
a striking effect on the selectivity to 1,3diaminopropane:
the production of byproducts can be significantly reduced
by working at pressures above 90 bar. The change in selectivity has been attributed to an increase in the concentration of
ammonia on the catalyst surface. This favors the fast ammonia addition to the aldehydetype intermediates, and suppresses the degradation (hydrogenolysis) type side reactions
as well as the reaction of the alcoholic group of the reactant
with the intermediate or product amines leading to dimerization products. The increase in ammonia surface concentration is traced to the transformation of the twophase system to
a homogeneous supercritical fluid, and to the ensuing elimination of the phase transfer resistance in the twophase regime,
increase in mass transfer and eventually influences of the
supercritical state on reaction thermodynamics and kinetics.
Most recently it has been shown that amination of
1,4cyclohexanediol in supercritical ammonia affords an efficient process for the synthesis of 1,4diaminocyclohexane [80].
An unsupported cobalt catalyst stabilized by 5 wt% iron
afforded the main reaction products 1,4diaminocyclohex-

T =
p =

89.0 C
3.0 MPa

T =
p =

91.3 C
13.2 MPa

T =
p =

108.0 C
14.1 MPa

T =
p =

9.4 C
9.0 MPa

DEA
H2
CO2

0.5 mol
0.6 mol

DEA
H2
CO2

0.5 mol
1.1 mol
3.0 mol

DEA
H2
CO2

0.5 mol
1.1 mol
3.0 mol

DEA
H2
CO2

0.5 mol
1.1 mol
3.0 mol

Figure 12 Phase behavior of the reactants used for synthesis of N,N diethylformamide (without hybrid aerogel catalyst) as recorded with the computercontrolled autoclave system for the investigation of heterogeneously catalyzed reactions in SCFs depicted in Fig. 8. The snapshots show the reactor content, and the shaft of the (suspended)
mechanical stirrer. Since the level of the liquid is quite low, only the images taken by the lower CCDcamera are shown. The bottom of the reactor (up to the lower end of the
glass window) is covered with a liquid phase consisting mainly of N,N diethylamine (DEA). Note that the level of this liquid phase does not significantly increase or decrease
with addition of CO2, isochoric heating to reaction conditions and subsequent isochoric cooling of the system; the reactor content obviously comprises two phases during the
synthesis of N,N diethylformamide from CO2, H2 and DEA, a liquid DEAbased phase and a supercritical CO2based phase.
140

OH OH
H2

R R
CO

R
OH

OH O
+ NH 3 / H2O
+ H2

R R

H2

HCHO

R R
NH3
+ H2

OH

R
OH

+ NH 3 / H2O
+ H2

H2

R R

NH 2

+ NH 3 / H2O
+ H2

H2

R R

R
NH2

NH 2

NH2

+ NH 3 / H2O
+ H2

R R
R = H, CH3

NH 2 NH 2

Figure 13

Important reactions occurring in the amination of alkanediols

with ammonia. Detected main and byproducts are indicated by orange and blue
shades, respectively. (Taken from [79].)

40

selectivity / %

30

20

10

40

60

80

100

120

Supercritical fluids

ane and 4aminocyclohexane with a cumulative selectivity


of 97% in a continuous fixed bed reactor. The unreacted diol
and amino alcohol intermediate can be recycled providing a
highly efficient process. This direct amination had not been
reported before. Currently, diaminocyclohexanes are industrially produced by the catalytic hydrogenation of aromatic
amines, such as pphenylenediamin. Considering the availability, oxidation stability and toxicology of the reactant, the
amination of cyclohexandiol in scNH3 is an attractive alternative.

R R

140

pressure / bar
Figure 14 Amination of 3amino1propanol with ammonia over a CoFe
(95/5) catalyst. Influence of pressure on the selectivity to 1,3diaminopropane.
Note the 10fold increase of selectivity around a pressure of 105 bar. Conditions:
continuous tubular reactor, T = 195 C, molar feed ratio of reactants ROH : NH3 =
1 : 40. The dashed line indicates the critical pressure of NH3. (Taken from [79].)

Outlook
The unique and easily tunable properties of fluids near their
critical point render SCFs very promising media for chemical
reactions. In addition, fluids used in chemistry at supercritical conditions are often very attractive from both economical and ecological points of view. Combining these media
with the industrially advantageous features of heterogeneous
catalysis thus holds great promise for optimization of chemical processes.
Limitations of SCF application arise from the higher
demands on the construction of suitable reaction apparatus
and the expenses of highpressure equipment. Furthermore,
the fundamentals of supercritical fluids in heterogeneous
catalysis have not yet been fully investigated and predictions
made on the basis of theoretical models and calculations are
of limited reliability so far.
Although the possibilities of SCF application in heterogeneous catalysis seem very impressive, the benefits of
SCFs have to be weighed against the higher costs involved
in designing, building and maintaining corresponding processes. To date, SCFs are mainly used where process intensification or favorable reaction behavior lead to economically
more attractive processes or where SCFs allow to alleviate
environmental constraints. However, if societys demand for
safe and environmentally benign chemical processes continues to increase, SCFs may become the ideal media for many
more chemical processes. For this purpose it will be necessary to further improve the knowledge in this area on both
fundamental and applied level.
Although SCFs are currently used mainly as solvents or
reactants in chemical reactions, there is an attractive opportunity where SCFs act simultaneously as both reactant and
solvent. This solventfree reaction design is a remarkable
step towards green processes. The easy tunability of an
SCFs unique properties enables the scientist and engineer to
explore the wide field of heterogeneous catalysis with markedly less restrictions than faced under classical conditions.
Consequently, great efforts are expended for exploring
this relatively new field and hopefully this will lead to further development of the theoretical basis of catalytic reactions under supercritical conditions. A deeper understanding should allow more directed optimization procedures for
chemical processes which would be beneficial for both application and research.
141

Supercritical fluids

questions and answers

What are the main applications of SCFs in industry?


SCFs are well established in extraction processes that
are used, e.g., in the food industry where scCO2 has
been employed for more than two decades as solvent
for the industrial production of decaffeinated coffee and
the extraction of hops aroma. Furthermore SCFs find
industrial application in chromatography, materials processing and the production of some bulk chemicals (see
below). However, SCFs are becoming increasingly popular as media for environmentally benign chemical processes.
Are there any industrial chemical processes currently
using SCFs?
Several well established processes for the production of
commodities are conducted at high pressures and temperatures beyond the critical value of the reaction mixture. The heterogeneously catalyzed synthesis of ammonia using the HaberBosch process (1913) and the homogeneously catalyzed high pressure polymerization of
ethene (1936) yielding lowdensity polyethene (LDPE)
are probably the most outstanding examples. Although
these processes are of great importance to the chemical
industry, the fact that they are conducted at the supercritical state of the reaction mixture is seldom acknowledged.
What are the limitations of SCFs in heterogeneous
catalysis?
As pointed out in the article, limitations arise mainly by
the higher costs of apparatus which withstand the necessary pressures.

acknowledgements

Financial support by the Schweizerisches Bundesamt fr Energie (BFE) and


Schweizerische Kommission fr Technologie und Innovation (KTI) as well as
F. HoffmannLa Roche AG and Degussa AG is gratefully acknowledged.

142

curricula vitae

Roland Wandeler was born in


Basel, Switzerland, in 1972. He
studied Chemical Engineering
at the Swiss Federal Institute
of Technology (ETH) in Zurich,
where he got his diploma in 1996.
In the same year he was awarded
the Willi Studer Prize of the Swiss Federal Institute of Technology (ETH) and started his work as a Ph.D. student in
the research group of Alfons Baiker. His project focusses on
the effect of phase behavior and composition on heterogeneously catalyzed reactions in supercritical fluids.

Alfons Baiker was born in Zurich,


Switzerland, in 1945. He studied Chemical Engineering at the
Swiss Federal Institute of Technology (ETH) in Zurich, where
he earned his Ph.D. in 1974. He
subsequently spent some years at
foreign universities as a postdoctoral fellow. During a
research stay with Michel Boudart at Stanford University
(19781980) he completed his habilitation thesis. After
returning to ETH he built up his research group at the Laboratory of Technical Chemistry, becoming Associate Professor in 1989 and full Professor in 1990. From 1993 to 1995,
he was Head of the Laboratory of Technical Chemistry. His
research lies in the areas of heterogeneous catalysis, chemical reaction engineering, and solid state chemistry. His
general goal is the development of environmentally benign
chemical processes which make optimal use of raw materials and energy. Present research interests embrace catalyst design and novel catalytic materials, enantioselective
catalysis, environmental and fine chemical catalysis, catalytic conversion of carbon dioxide, and the application of
supercritical fluids in heterogeneous catalysis. Baiker was
a guest professor at the University of California, Berkeley,
and at the Max Planck Institut fr Kohleforschung, Mlheim a.d. Ruhr, Germany. He has published more than 450
research papers largely in the field of catalysis and holds
numerous patents.

Supercritical fluids

references

[1]

Baiker, A. Chem. Rev. 1999, 99, 453473.

[2]

Brennecke, J. F.; Chateauneuf, J. E. Chem. Rev. 1999, 99, 433452.

[3]

Jessop, P. G.; Ikariya, T.; Noyori, R. Chem. Rev. 1999, 99, 475493.

[4]

Kendall, J. L.; Canelas, D. A.; Young, J. L.; DeSimone, J. M. Chem. Rev. 1999,

[43]

Fourth Italian Conference on Supercritical Fluids and their Applications, Capri


(Napoli), Italy, 1997; p 385391.
[44]

99, 603621.
[5]

Savage, P. E. Chem. Rev. 1999, 99, 603621.

[6]

Chemical Synthesis Using Supercritical Fluids ; Jessop, P. G.; Leitner, W.,

Reverchon, E.; Barba, A. A.; Della Porta, G.; Ciambelli, P.; Sannino, D.

Pickel, K. H.; Steiner, K. 3rd International Symposium on Supercritical Fluids,


Strasbourg, France, 1994; p 712.

[45]

Poliakoff, M. The 4th International Symposium on Supercritical Fluids, Sendai,


Japan, 1997; p 757762.

Eds.; WileyVCH: Weinheim, 1999.

[46]

Hoffmann La Roche Ltd. In Roche Magazin, 1992.

[7]

Leitner, W. Top. Curr. Chem. 1999, 206, 107132.

[47]

Stinson, S. C. C&EN 1997, 3755.

[8]

Dinjus, E.; Fornika, R.; Scholz, M. In Chemistry Under Extreme or Non Clas-

[48]

Schneider, M.; Baiker, A. Catal. Rev. Sci. Eng. 1995, 37, 515556.

sical Conditions ; van Eldik, R., Hubbard, C. D., Eds.; J. Wiley & Sons, Inc. and

[49]

Ward, D. A.; Ko, E. I. Ind. Eng. Chem. Res. 1995, 34, 421433.

Spektrum Akademischer Verlag copublication: New York, 1997.

[50]

Hsing, N.; Schubert, U. Angew. Chem. Int. Ed. Engl. 1998, 37, 2245.

Moller, P. Fourth Italian Conference on Supercritical Fluids and their Applica-

[51]

Fink, R.; Beckman, E. J. In Chemical Synthesis using Supercritical Fluids ;

[9]

tions, Capri (Napoli), Italy, 1997; p 315318.


[10]

Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim; New York, 1999.

Poliakoff, M.; George, M. W.; Howdle, S. M. In Chemistry Under Extreme or

[52]

Arai, K.; Adschiri, T. Fluid Phase Equil. 1999, 158160, 673684.

NonClassical Conditions ; van Eldik, R., Hubbard, C. D., Eds.; J. Wiley & Sons,

[53]

Peters, C. J. In Supercritical Fluids Fundamentals for Application ; Kiran, E.,

Inc. and Spektrum Akademischer Verlag co publication: New York, 1997.


[11]
[12]

Levelt Sengers, J. M. H., Eds.; Kluwer Academic Publishers: Dordrecht, 1994.

Savage, P. E. In Handbook of Heterogeneous Catalysis ; Ertl, G., Knzinger,

[54]

Brennecke, J. F.; Eckert, C. A. AIChE J. 1989, 35, 14091427.

H., Weitkamp, J., Eds.; WileyVCH: Weinheim, 1997; Vol. 3.

[55]

Schneider, G. M. In Supercritical Fluids Fundamentals for Application ; Kiran,

Ding, Z. Y.; Frisch, M. A.; Li, L.; Gloyna, E. F. Ind. Eng. Chem. Res. 1996, 35,

E., Levelt Sengers, J. M. H., Eds.; Kluwer Academic Publishers: Dordrecht,

32573279.
[13]

Eckert, C. A.; Knutson, B. L.; Debenedetti, P. G. Nature 1996, 383, 313318.

[14]

Savage, P. E.; Gopalan, S.; Mizan, T. I.; Martino, C. J.; Brock, E. E. AIChE J.

1994.
[56]

ence and Technology ; Hutchenson, K. W., Foster, N. R., Eds.; American

1995, 41, 17231778.


[15]

Wu, B. C.; Paspek, S. C.; Klein, M. T.; LaMarca, C. In Supercritical Fluid Tech-

Hutchenson, K. W.; Foster, N. R. In Innovations in Supercritical Fluids: SciChemical Society: Washington D.C., 1995.

[57]

nology: Reviews in Modern Theory and Applications ; Bruno, T. J., Ely, J. F.,

Taylor, L. T. Supercritical Fluid Extraction ; John Wiley & Sons, Inc.: New York,
1996.

Eds.; CRC Press: Boca Raton, 1991.

[58]

Johnston, K. P.; Peck, D. G.; Kim, S. Ind. Eng. Chem. Res. 1989, 28, 11151125.

[16]

Tiltscher, H.; Hofmann, H. Chem. Eng. Sci. 1987, 42, 959977.

[59]

Tsiklis, D. S. In Handbook of Techniques in HighPressure Research and

[17]

Subramaniam, B.; McHugh, M. A. Ind. Eng. Chem. Process Des. Dev. 1986,
25, 112.

[18]

Cagniard de LaTour, C. Ann. Chim. Phys. 1822, 21, 127132, 178182.

[19]

Jessop, P. G.; Leitner, W. In Chemical Synthesis Using Supercritical Fluids ;

Engineering ; Plenum Press: New York, 1968.


[60]

Eds.; Butterworths: London, 1975; Vol. 2.


[61]

Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim, 1999.


[20]

Chrastil, J. J. Phys. Chem. 1982, 86, 30163021.

[21]

Debenedetti, P. G.; Reid, R. C. AIChE J. 1986, 32, 20342046.

[22]

Huang, K. Statistical Mechanics ; 2nd ed.; John Wiley & Sons: New York, 1987.

[23]

Levelt Sengers, J. M. H. In Supercritical Fluids Fundamentals for Applica-

[24]

and Practice ; 2nd ed.; ButterworthHeinemann: Boston, 1994.


[63]

Dohrn, R.; Haverkamp, V. The 4th International Symposium on Supercritical


Fluids, Sendai, Japan, 1997; p 811818.

[64]

Wandeler, R.; Baiker, A. Chimia 1999, 53, 566569.

Dordrecht, 1994.

[65]

Brennecke, J. F. In Supercritical Fluid Engineering Science: Fundamentals

Lucien, F. P.; Foster, N. R. In Chemical Synthesis Using Supercritical Fluids ;


van Eldik, R.; Asano, T.; le Noble, W. J. Chem. Rev. 1989, 89, 549688.

[26]

Clifford, A. A. In Supercritical Fluids Fundamentals for Application ; Kiran, E.,

and Applications ; Kiran, E., Brennecke, J. F., Eds.; American Chemical Society: Washington DC, 1993; Vol. 11.
[66]

Moore, J. W.; Pearsons, R. G. Kinetics and Mechanism ; 3rd ed.; John Wiley

Spektrum Akademischer Verlag co publication: New York, 1997.


[67]

& Sons: New York, 1981.


Ekart, M. P.; Bennett, K. L.; Ekart, S. M.; Gurdial, G. S.; Liotta, C. L.; Eckert,

Howdle, S. M.; George, M. W.; Poliakoff, M. In Chemical Synthesis Using Supercritical Fluids ; Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim, 1999.

[68]

C. A. AIChE J. 1993, 39, 235248.

Rathke, J. W.; Klingler, R. J.; Gerald II, R. E.; Fremgen, D. E.; Woelk, K.;
Gaemers, S.; Elsevier, C. J. In Chemical Synthesis Using Supercritical Fluids ;

Brennecke, J. F.; Debenedetti, P. G.; Eckert, C. A.; Johnston, K. P. AIChE J.


1990, 36, 19271928.

Hubbard, C. D.; van Eldick, R. In Chemistry Under Extreme or NonClassical Conditions ; van Eldik, R., Hubbard, C. D., Eds.; J. Wiley & Sons, Inc. and

Levelt Sengers, J. M. H., Eds.; Kluwer Academic Publishers: Dordrecht, 1994.

[29]

McHugh, M. A.; Krukonis, V. J. In Supercritical Fluid Extraction Principles

tion ; Kiran, E., Levelt Sengers, J. M. H., Eds.; Kluwer Academic Publishers:

[25]

[28]

Sherman, W. F.; Stadtmuller, A. A. Experimental Techniques in HighPressure Research ; John Wiley & Sons, Inc.: London, 1987.

[62]

Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim, 1999.

[27]

Schneider, G. M. In Experimental Thermodynamics ; Le Neindre, B., Vodar, B.,

Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim, 1999.


[69]

Yonker, C. R.; Linehan, J. C.; Fulton, J. In Chemical Synthesis Using Super-

[30]

Eckert, C. A.; Knutson, B. L. Fluid Phase Equil. 1993, 83, 93100.

[31]

Eckert, C. A.; Chandler, K. The 4th International Symposium on Supercritical

[70]

Jessop, P. G.; Ikariya, T.; Noyori, R. Chem. Rev. 1995, 95, 259272.

Fluids, Sendai, Japan, 1997; p 799806.

[71]

Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1994, 116,

[32]

McHugh, M. A.; Krukonis, V. J. Supercritical Fluid Extraction Principles and


Practice ; 2nd ed.; ButterworthHeinemann: Boston, 1994.

[33]

critical Fluids ; Jessop, P. G., Leitner, W., Eds.; WileyVCH: Weinheim, 1999.

88518852.
[72]

Vesovic, V.; Wakeham, W. A. In Supercritical Fluid Technology: Reviews in

Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Chem. Soc. Chem.
Commun. 1995, 707708.

Modern Theory and Applications ; Bruno, T. J., Ely, J. F., Eds.; CRC Press:

[73]

Krcher, O.; Kppel, R. A.; Frba, M.; Baiker, A. J. Catal. 1998, 178, 284298.

Boca Raton, 1991.

[74]

Schmid, L.; Rohr, M.; Baiker, A. Chem. Commun. 1999, 23032304.

[34]

Vesovic, V.; Wakeham, W. A. J. Phys. Chem. Ref. Data 1990, 19, 763808.

[75]

Schmid, L.; Krcher, O.; Kppel, R. A.; Baiker, A. Micropor. Mesopor. Mat.

[35]

Reid, R. C.; Prausnitz, J. M.; Poling, B. E. The Properties of Gases and Liq-

[36]

2000, 3536, 181193.

uids ; 4th ed.; McGraw Hill: New York, 1988.

[76]

Krcher, O.; Kppel, R. A.; Baiker, A. Chimia 1997, 51, 4851.

Dillow, A. K.; Brown, J. S.; Liotta, C. L.; Eckert, C. A. J. Phys. Chem. A 1998,

[77]

Krcher, O.; Kppel, R. A.; Baiker, A. J. Chem. Soc. Chem. Commun. 1997,

102, 76097617.
[37]

Dillow, A. K.; Hafner, K. P.; Yun, S. L. J.; Deng, F.; Kazarian, S. G.; Liotta, C.

453454.
[78]

L.; Eckert, C. A. AIChE J. 1997, 43, 515524.


[38]

Tiltscher, H.; Wolf, H.; Schelchshorn, J.; Dialer, K., United States Patent

Mallat, T.; Baiker, A. In Handbook of Heterogeneous Catalysis ; Ertl, G., Knzinger, H., Weitkamp, J., Eds.; VCH: Weinheim, 1997; Vol. 5.

[79]

4605811, 1986.

Fischer, A.; Mallat, T.; Baiker, A. Angew. Chem. Int. Ed. Engl. 1999, 38,
351354.

[39]

Subramaniam, B.; McCoy, B. J. Ind. Eng. Chem. Res. 1994, 33, 504508.

[80]

Fischer, A.; Mallat, T.; Baiker, A. J. Catal. 1999, 182, 289291.

[40]

Tom, J. W.; Debenedetti, P. G. J. Aerosol Sci. 1991, 22, 555.

[81]

Handbook of Chemistry and Physics ; 72nd ed.; Lide, D. R., Ed.; CRC Press:

[41]

Dixon, D. J.; Johnston, K. P.; Bodmeier, R. A. AIChE J. 1993, 39, 127139.

[42]

Reverchon, E. The 5th Meeting on Supercritical Fluids Materials and

Boca Raton, 1991.

Natural Products Processing, Nice (France), 1998; p 221236.

143

You might also like