Você está na página 1de 54

121

8 Automotive catalytic converter


8.1 Introduction
Today, the overwhelming majority of vehicles is equipped with internal combustion
engines that use fossil derived fuels. This on-road traffic represents the main source of
pollutant emissions of carbon monoxide, volatile organic compounds (VOC, unburnt
hydrocarbons), and nitrogen oxides as is shown in Table 8.1. The initiative to introduce
legislation for the limitation of exhaust gas emissions from vehicles was taken in Cali-
fornia in 1966. At first, improvements of the internal combustion process were sufficient
to meet the demands. The continuous decrease of the limits by American legislation
required the aftertreatment of the exhaust gas by automotive catalytic converters intro-
duced in 1976 [195]. In 1985, exhaust gas pollution limits were introduced in Europe,
which consequently also led to the utilization of automotive catalytic converters.

NOx CO VOC SO2


Total emissions [1000 tons] 11932 40964 13807 9386
Road traffic’s share in emissions 40% 56% 31% 4%

Table 8.1: Pollutant emissions in Europe in 1997 [196].

The most frequently used design of a catalytic converter is a monolithic structure (Fig.
8.1), which is coated with a washcoat that supports the catalyst material. The devel-
opment of such a catalytic converter is a complex process involving the optimization
of different physical and chemical parameters. Simple properties such as monolith
length, cell density and metal loading of the catalyst influence the performance of the
converter. Numerical simulation using detailed models for the transport and chem-
ical processes are expected to accelerate the design and optimization of automotive
catalytic converters.

Figure 8.1: Sketch of a three-way catalytic converter; courtesy of J. Eberspächer GmbH


& Co.

In this Chapter, the modeling and numerical simulation of steady-state and transient
processes in automotive catalytic converters are discussed. First the focus is on a
122 8. AUTOMOTIVE CATALYTIC CONVERTER

typical three-way catalytic converters (3WCC) as commonly used in passenger cars.


The computational tools, developed in this work, will be applied to study the steady-
state behavior in the single channel of the catalytic monolith as well as the transient
behavior of the whole monolith. In the second part, a DeNOx catalyst is investigated.
Here, the focus is on the application of different transport models. In particular, the
effect of simplified models for the description of washcoat diffusion and flow field will
be discussed.

8.2 Three-way catalyst


Today, three-way catalytic converters are used extensively to reduce the pollutant emis-
sions of internal combustion engines. The role of the 3WCC is the complete oxidation
of carbon monoxide, formed in the combustion process, and unburned hydrocarbons to
harmless water and carbon dioxide14 , and the reduction of nitrogen oxides to molecular
nitrogen. These three functions of the catalytic converter can chemically be written as

CO + 1
2
O2 → CO2

Cn Hm + (n+ m4 ) O2 → n CO2 + m
2
H2 O

CO + NO → CO2 + N2 .

The simultaneous conversion of all these harmful species can only be ensured if the
exhaust gas composition is very close to the stoichiometric ratio. This requirement
is technically controlled by the lambda-sensor and accompanying electronics for en-
gine control [32]. Meanwhile, the performance of 3WCC is almost perfect after the
catalytic reaction is ignited. However, at low operating temperatures, below 570 K,
almost no conversion of the pollutant emissions occurs. Therefore, current research
and development focuses on the reduction of this start-up period.
The majority of automotive catalytic converters have a monolithic structure made
of ceramics or metals. The walls of the monolith channels are coated with the wash-
coat, usually alumina, that supports the noble metal such as platinum, palladium and
rhodium. The monoliths consist of numerous parallel channels with a diameter of ap-
proximately 1 mm to have a large catalytic surface area. For the design of a catalytic
converter, several chemical and physical properties of both the catalyst and the ex-
haust gas must be considered, for instance the cell geometry (length and diameter of
the channels, wall thickness), the composition of the noble metal and catalyst loading,
the use of catalyst promoters, and the properties of the exhaust gas (temperature,
velocity and chemical composition).
The experimental characterization of the catalytic performance of the converter is
time-consuming and requires an expensive and complex experimental setup. Numerical
simulation offers an interesting alternative for the investigation of the catalytic activity
of a converter as a function of external conditions. This method is also efficient in
analyzing the transient flow and thermal phenomena in the catalytic converter and can
14
The effect of CO2 emissions on global warming is not a topic of this work.
8.2 Three-way catalyst 123

help to understand the complex interactions between the flow field and the catalytic
surface chemistry.
In recent years, several proposals were made for the numerical simulation of cat-
alytic converters [197–202]. In most of these studies, a global model for the chemistry
was used, which neglected the complex network of chemical reactions on the catalytic
surface. An alternate approach is the description of the chemical reactions by a set of
elementary-like reaction steps describing the chemical processes on a molecular level as
discussed in Chapter 3.2 and by Chatterjee et. al for 3WCC [87, 203]. This approach
is superior to any fitted global kinetics because it can be used to predict the catalyst
behavior at different external conditions. However, also the transport processes have
to be described and coupled to the chemical reactions in an accurate manner.
Detailed models for the chemistry and the transport processes are applied in this
study of a 3WCC. The numerical simulations are carried out using the computational
tools discussed in Chapter 4. The numerically predicted conversion of pollutants is
compared with experimentally derived data.
The catalyst investigated is a commercially available tree-way catalyst [87]. The
catalyst contains 50 g/ft3 of rhodium and platinum metal with a Pt/Rh ratio of 5/1.
The noble metals were impregnated on a ceria stabilized γ-alumina washcoat. The
washcoat was supported by a cordierite monolith with a cell density of 62 cells per cm2
(400 cpsi) and a wall thickness of 0.165 mm.
At first, the transport and chemistry in a single channel of the catalytic monolith
is discussed for steady-state conditions. Then the transient behavior of the 3WCC at
real operating conditions is described.

8.2.1 Steady state conditions

Experimental
The experimental determination of conversion as a function of temperature in the
three-way catalyst was conducted in an isothermal laboratory-scale tube reactor.15 A
sample of 22 mm in diameter and 29 mm in length was taken from the catalyst for this
investigation. Table 8.2 summarizes the composition and the species concentrations of
the exhaust gas sample. The oxygen concentration was varied in order to simulate a
stoichiometric, rich and lean exhaust gas mixture. The λox -values are defined as the
inverse of the redox ratio [197, 201, 87]:

XNO + 2XO2
λox = . (8.1)
XCO + 9XC3 H6

With this definition λox equals unity at stoichiometric conditions, and it is larger
(smaller) than unity for lean (rich) conditions. The volumetric flow of the exhaust
gas was 15 l/min at standard conditions (298.15 K). A uniform superficial velocity of
1.35 m/s corresponds to this volumetric flow rate. The reactor was heated with a tubu-
lar furnace with a heat rate of 100 K/h, which led to an almost isothermal sample. The
15
The experiment was carried out by S. Kureti and O. Görke at the University of Karlsruhe.
124 8. AUTOMOTIVE CATALYTIC CONVERTER

species concentration [vol.%]


nearly stoichiometric rich lean
(λox =0.9) (λox =0.5) (λox s=1.8)
CO 1.42 1.42 1.42
O2 0.77 0.9 1.6
C3 H6 0.045 0.045 0.045
NO 0.1 0.1 0.1
N2 balance balance balance

Table 8.2: Composition of the simulated exhaust gas used in the experiment and
simulation.
Noble metal composition Pt/Rh, 5:1
Noble metal loading 50 g/ft3
Active noble metal surface 0.247 m2 /g
Surface ratio Pt(s)/Rh(s) 3:1
Noble metal dispersion 29.1%
Catalyst surface/geometric surface (Fcat/geo ) 70
Channel diameter 1 mm
Channel length 29 mm
Superficial velocity (STP) 1.35 m/s
Mean pore diameter (micropores) 12.29 nm
Porosity (micropores) 27.9 %
Tortuousity 3
Washcoat thickness 100 µm
Active catalyst surface/washcoat volume 7.77·105 m−1

Table 8.3: Catalyst parameters used in the numerical simulation of the 3WCC.

maximum temperature gradient over the sample was reported to be 40 K caused by the
exothermicity of the oxidation reactions. The exit species concentrations were mea-
sured in 20 K steps. The steady state of the system was ensured at each temperature
step.
The CO2 concentration in the inlet exhaust gas was zero due to experimental rea-
sons. Tests have shown that there was no difference in conversion of CO, HC and NO
in measurements whether or not CO2 was added.
After the experimental studies of the temperature-dependent conversion, the sample
was investigated with H2 chemisorption in order to obtain the properties of the active
metal phase. The active metal surface of the catalyst was 28 m2 /g and the dispersion of
the conditioned catalyst was 33%. The calculation of the ratio of active metal surface
area and geometric surface area (Fcat/geo in Eq. 3.28) of the catalyst led to a value
of 70. The ratio of the platinum to rhodium surface was chosen to be 3:1 taken from
literature [204] for a catalyst with a noble metal composition of Pt/Rh with 5:1, which
corresponds to the catalyst used for the investigations.
8.2 Three-way catalyst 125

1.0 1.0

0.8 0.8 experiment


C3H6 - conversion

CO - conversion
experiment
simulation
0.6 eimulation 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 600 800 1000 400 600 800 1000
T [K] T [K]

1.0

0.8 experiment
NO - conversion

simulation
0.6

0.4

0.2

0.0
400 600 800 1000
T [K]

Figure 8.2: Conversion of C3 H6 , CO, and NO at lean conditions as function of temper-


ature [87].

Numerical

It can be assumed that all channels of the monolith behave essentially alike because an
isothermal sample was used in the experiment, and the inlet conditions (flow velocity
and gas composition) did not vary over the various monolith channels. Therefore,
only one channel needs to be analyzed. The single-channel is assumed to be a tube-like
reactor with the flow inside this reactor being laminar. Hence, the single channel of the
monolith can be modeled by the axisymmetric two-dimensional Navier-Stokes equations
with the axial and radial directions as spatial variables as discussed in Chapter 2.2.1.
The flow field simulation is based on the CFD code FLUENT coupled with the chemistry
module DETCHEM as described in Chapter 4.3.1.
The flow enters the computational domain with a known velocity, gas composition
and temperature given by the experimental conditions. A flat profile of the axial
velocity and a vanishing radial velocity are used in the simulation at the inlet boundary.
At the reactor exit, an outlet boundary is applied at which values for all variables are
extrapolated from the interior cells adjacent to the outlet. A structured grid is used
for the simulation. The grid has to be very fine around the catalyst entrance and the
catalytic wall to resolve the flow field and also to determine the variations in the species
concentrations due to chemical reactions at the catalytic wall. The total number of
computational cells used for the single channel were 20 cells in radial direction and 72
126 8. AUTOMOTIVE CATALYTIC CONVERTER

1.0 1.0
experiment experiment
0.8 0.8
C3H6 - conversion

CO - conversion
simulation simulation

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 600 800 1000 400 600 800 1000
T [K] T [K]

1.0
experiment
0.8
NO - conversion

simulation

0.6

0.4

0.2

0.0
400 600 800 1000
T [K]

Figure 8.3: Conversion of C3 H6 , CO, and NO at nearly stoichiometric conditions as


function of temperature [87].

cells in axial direction.


A detailed multi-step reaction mechanism is used to model the catalytic reactions
and to calculate the surface mass fluxes. The surface coverage of the species on the
catalytic material is also calculated as a function of the position in the channel. The
mechanism includes only surface chemistry; gas phase chemistry can be neglected be-
cause of the low pressure and temperature, and the short residence time.
The sample exhaust gas mixture is composed of C3 H6 , CO, NO, O2 , N2 with con-
centrations as given in Tab. 8.2. The surface reaction scheme consists of 61 reaction
steps among 27 chemical species, e.g., dissociative oxygen adsorption, non-dissociative
adsorption of C3 H6 , CO and NO, the formation steps of carbon dioxide, water and
nitrogen, and desorption reactions for all species. Some activation energies (e.g., oxy-
gen desorption) are coverage-dependent due to interactions between adsorbed species.
It is assumed that all species are adsorbed competitively. The model also considers
the different adsorption sites (platinum or rhodium) on the metallic catalyst surface.
However, on rhodium, surface reactions only between NO, CO, and O2 are consid-
ered [49, 205]. The kinetic data of the mechanism were taken either from literature or
fits to experimental data. Parts of the surface reaction mechanism were already used
for numerical modeling of catalytic ignition (Chapter 7 and [45]), simulation of total
and partial oxidation of light hydrocarbons on platinum (Chapter 6 and [49]) and mod-
eling the CO-O2 and NO-CO reactions on rhodium [205, 206]. A detailed description
8.2 Three-way catalyst 127

1.0 1.0
experiment experiment
0.8 0.8
C3H6 - conversion

CO - conversion
simulation simulation

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 600 800 1000 400 600 800 1000
T [K] T [K]

1.0
experiment
0.8
NO - conversion

simulation

0.6

0.4

0.2

0.0
400 600 800 1000
T [K]

Figure 8.4: Conversion of C3 H6 , CO, and NO at rich conditions as function of temper-


ature [87].

of the reaction mechanism can be found in reference [203]; the mechanism is given in
Table B.5. In the simulation, the simplified washcoat model (Chapter 3.3.2) is used
with CO as species that determines the effectiveness factor (Eq. 3.49).

Results and discussion

The experimental conditions are applied to the simulation. Table 8.3 summarizes the
input data for the simulation; the concentrations of the incoming exhaust gas are
already given in Tab. 8.2. The gas flows at a uniform velocity into the cylindric tube.
The temperature is set to the furnace temperature.
In Fig. 8.2, the conversion data of CO, C3 H6 , and NO are shown as function of
temperature. The composition of the inlet gas mixture is lean. The conversion of
CO, C3 H6 , and NO starts at 570 K and increases up to 100% for CO and C3 H6 at
770 K and 670 K, respectively. The NO conversion shows a maximum at 630 K and
decreases at higher temperatures. The predicted conversion of all three species agree
well with the experimentally measured data. Especially the temperature behavior
of the NO conversion and the slow increase of the C3 H6 conversion is well-predicted
in the temperature range above 630 K. The simulation shows some CO conversion
at temperatures lower than 570 K. This behavior is probably caused by the missing
reactions for hydrocarbons on rhodium in the mechanism. Therefore C3 H6 cannot
128 8. AUTOMOTIVE CATALYTIC CONVERTER

block the CO oxidation on Rh in contrast to the CO oxidation on Pt.


The conversion data as a function of the temperature for the nearly stoichiometric
mixture are shown in Fig. 8.3. Again, conversion of CO, C3 H6 , and NO starts at 570 K,
but increases more slowly with temperature compared to the lean mixture. Because of
the insufficient amount of O2 in the mixture, CO conversion is not complete. Complete
conversion of C3 H6 is reached at 770 K, which indicates that C3 H6 can compete with
CO for O2 . Also, for temperatures higher than 720 K, NO reduction is completed.
Concerning the conversion of CO and NO and the competition between CO and C3 H6
for O2 , the simulation results agree well with the experimental data. Only C3 H6 rate
shows deviations between 610 K and 690 K. Compared to the experimental data, the
predicted conversion is too high.
The rich mixture, as shown in Fig. 8.4, contains 0.4 vol.% O2 , which leads to a
maximum of CO conversion of only 33%. The conversion of NO reaches 100% for tem-
peratures higher than 800 K. In comparison with the results of the two other mixtures,
the increase of the conversions of CO, C3 H6 and NO with temperature is slower, in
particular for C3 H6 conversion. Regarding the CO and NO conversion, the simulations
agree with the experimentally determined data. However, large deviations exist be-
tween the predicted C3 H6 conversion and the experimental data. These deviations can
be explained by the fact that in the rich regime a wider variety of surface species (e.g.,
partial-oxidation products of C3 H6 ) resides on the catalytic surface, which can reduce
the oxygen coverage and lead to different reaction paths. In the reaction mechanism
used in this study not all surface species possible are included. Further reactions and
surface species have to be included to improve the prediction of the C3 H6 conversion
at rich conditions.
In Fig. 8.5, the mass fractions of C3 H6 , CO, CO2 , and NO for the lean mixture
(Tab. 8.2) within the channel are shown for a temperature of 673 K . The input data
are as given in Tab. 8.3. The mass fraction profiles show that most of the propylene
is converted within the first centimeter. In this axial range, CO is almost completely
converted. The NO conversion is limited to the first centimeter and vanishes further
downstream. This behavior can be explained by the surface coverages. The calculated
coverages of the most relevant surface species on platinum and rhodium are shown as
function of the axial position along the channel in Fig. 8.6. The coverages are defined
in respect of the whole catalytic area, consisting of rhodium and platinum. A strong
coverage variation is revealed on the Pt surface at an axial position of 1.1 cm due to
the decreasing CO concentration in the gas phase. There, the surface state shifts from
a mainly CO(s)-covered to an O(s)-covered state. During this transition the number
of free platinum sites, Pt(s), increases, which allows more NO to be adsorbed and
dissociated. Where the surface finally reaches the O(s) covered state, the number of
free platinum sites is decreased, and the equilibrium of NO dissociation is shifted to
NO(s) resulting in a vanishing NO conversion.
The rhodium surface shifts from a N(s)-covered state to an O(s)-covered state.
The O(s)-covered surface also prevents NO conversion on Rh. In front of the axial
position of 1 cm the surface is mainly covered with N(s) and active for NO conversion.
With increasing reaction temperature, the transition point moves toward the channel
entrance, which reduces the area that is active for NO conversion. In Fig. 8.2 the
8.2 Three-way catalyst 129

C3H6 6.70E-04
0.050
5.47E-04
0.025
r [cm]

4.25E-04
0.000
3.02E-04
-0.025
1.80E-04
-0.050
0.0 0.5 1.0 1.5 2.0 2.5
5.70E-05
z [cm]
CO 1.42E-02
0.050
1.14E-02
0.025
r [cm]

8.52E-03
0.000
5.68E-03
-0.025
2.84E-03
-0.050
0.0 0.5 1.0 1.5 2.0 2.5
3.50E-06
z [cm]
CO2 2.40E-02
0.050
1.92E-02
0.025
r [cm]

1.44E-02
0.000
9.60E-03
-0.025
4.80E-03
-0.050
0.0 0.5 1.0 1.5 2.0 2.5
0.00E+00
r [cm]
NO 1.07E-03
0.050
1.00E-03
0.025
r [cm]

9.38E-04
0.000
8.72E-04
-0.025
8.06E-04
-0.050
0.0 0.5 1.0 1.5 2.0 2.5
7.40E-04
z [cm]

Figure 8.5: Mass fraction profiles in a single channel of the 3-way catalyst monolith
at 673 K; lean conditions (λox =1.8) [87]. Different scales are used in axial and radial
direction for visual clarity.
130 8. AUTOMOTIVE CATALYTIC CONVERTER

100 CO(Pt) 100


N(Rh)

10-1 Pt(s) 10
-1

Rh - surface coverage
Pt - surface coverage

NO(Pt) O(Rh)
-2
10 10-2
O(Pt)
10-3 10-3
OH(Pt) Rh(s)
N(Pt)
10-4 10
-4 NO(Rh)

CO(Rh)
10-5 10-5
C3H6(Pt)

10-6 H2O(Pt) 10-6

10-7 10
-7
0 1 2 0 1 2
z [cm] z [cm]

Figure 8.6: Pt and Rh surface coverages along the catalytic wall in a single-channel of
a 3-way catalyst monolith at 673K; lean conditions (λox =1.8) [87].

resulting decrease of NO conversion with increasing temperature is shown.


A typical feature of the three-way converter is the λ-window behavior. The λ-
window is the narrow range around the stoichiometric air/fuel ratio, in which hy-
drocarbons, CO, and NO are simultaneously converted with high efficiency. Fig. 8.7
presents the predicted conversion for the catalytic converter tested for various exhaust
gas compositions. The chosen temperature of 773 K is a typical catalyst inlet tem-
perature for partial load [32]. The experimental data were already presented in the
Figs. 8.2, 8.3, and 8.4. The simulation predicts the experimental data well in the lean
region. In the rich regime the CO conversion is also predicted well, while there are
major deficiencies for C3 H6 conversion as discussed above.
C3H6, simulation C3H6, experiment
CO, simulation CO, experiment
NO, simulation NO, experiment

0.8
conversion

0.6

0.4

0.2

0 0.5 1 1.5 2
λ ox

Figure 8.7: Experimentally determined and numerically predicted conversion of C3 H6 ,


CO, and NO as function of the fuel/oxygen ratio (λox - window) at 773 K [87].
8.2 Three-way catalyst 131

8.2.2 Transient conditions

The ultimate goal in numerical simulation of automotive catalytic converters is the


prediction of exhaust gas emissions as function of time for varying inlet conditions, i.e.,
the simulation of a driving cycle. Such a simulation must include the calculation of the
transient three-dimensional temperature-field of the monolithic solid structure of the
converter, which results from a complex interaction of a variety of physical and chemical
processes: (1) time-dependent and spatially-varying exhaust gas conditions (tempera-
ture, species concentrations, velocity), (2) heat conduction with spatially varying and
temperature-dependent thermal conductivity in different solid media (ceramic or metal
monolith structure, insulation material, canning), (3) heat transport between exterior
converter wall and the ambient air by conduction, convection, and thermal radiation,
(4) axial and radial mass transport including pore diffusion in the washcoat in the
single monolith channels, (5) the chemical reactions and heat release on the catalyst in
the single-channels, (6) heat transfer between fluid phase, catalyst/washcoat and solid
converter structure. Then, the integration over the chemical conversion in the single
channels leads to the total conversion of the monolith as function of time.
The computer program DETCHEMMONOLITH [92, 93, 81], which has been discussed
in Chapter 4.3.4, can be applied for the numerical simulation of the transient behavior
of automotive catalytic converters with a monolithic structure. All effects mentioned
above can be included.
As an example, a numerical simulation of the start-up period of a 3WCC is per-
formed and discussed now. The hot exhaust gas is fed in the initially cold catalytic
converter, which heats up the converter and eventually ignites the catalytic reactions
leading to the conversion of the harmful pollutants. The converter including the mono-
lithic structure, the fiber insulation, and the metal canning has a cylindric shape with
a total diameter of ∼10 cm and a length of ∼20 cm. A sample exhaust gas composed of
C3 H6 , O2 , CO, NO, and N2 flows at a spatially and temporally constant composition,
temperature (673 K), and superficial velocity (3 m/s) in the initially cold (300 K) cat-
alytic converter. The catalyst parameters are applied as in the investigation described
in Chapter 8.2.1 and [87]) again.
The heat transport model in the numerical simulation uses axial and radial varying,
temperature-dependent thermal conductivity for all three solid materials (monolith,
insulation, canning). The simplified washcoat model is applied, where the efficiency
factor refers to the species CO. The chemistry model is based on the surface reaction
scheme discussed above. Radiative and convective heat losses at the exterior boundary
of the converter are taken into account.
For example, Fig. 8.8 reveals the temperature profile in the solid converter structure
and the CO mass fraction in one of the many single channels at three different time
periods after starting the engine. Five seconds after start-up, only the front end of the
monolith is already warmed up slightly, leading to little CO conversion. After 15 s, the
maximum catalyst temperature already is significantly higher the the inlet gas tem-
perature. Now, considerable chemical reactions occur; the exothermic CO oxidation is
almost complete, although only the front end of the converter is hot. The exothermicity
of the oxidation reactions (CO and C3 H8 ) lead to this peak temperature. At that time,
132 8. AUTOMOTIVE CATALYTIC CONVERTER

t=5s

T (K)
800
767
733
700
667 t = 15 s
633
600
567
533
500
467
433
400
367
333
300

t = 100 s

YCO
0.0142
0

Figure 8.8: Two-dimensional temperature profiles of a 3WCC (left) and CO mass


fractions in a single channel located in the center of the monolith (right) at 5, 15, and
100 s after start-up of the engine; the converter, 20 cm in length and 10 cm in diameter
including insulation and canning, is initially is at 298 K; inlet gas conditions (spatial
and temporal constant): u = 3 m/s, T = 673 K, vol.% of C3 H6 =0.00045, O2 =0.016,
CO=0.0142, NO=0.001, N2 =0.96835.

conversion in the outer channels is still low (not shown), since the temperature is lower
due to the heat loss into the ambient air. A large region of the converter has reached
its operating temperature after 100 s. However, the CO profile in a single-channel in
the center of the catalyst does not differ from the one at 15 s of operating time. These
results elucidate that the overall reaction rate of CO oxidation is already mass transfer
limited.
Similar simulations have been carried out for temporally and spatially varying inlet
conditions [93]. The computation of an complete driving cycle for passenger vehicles
as given by European exhaust gas legislation [196] will take approximately 10 hours
CPU-time on Apple Mac G4 processors. In these simulations, between ten and twenty
representative channels are simulated at each time step in parallel manner, i.e., each
processor solves one channel.
8.3 DeNOx catalyst 133

8.3 DeNOx catalyst


In spite of the enormous achievements in the aftertreatment of exhaust gas emissions,
the worldwide increasing number of vehicles represent a serious environmental problem
due to vehicles’ raw emissions, in particular, carbon dioxide, which has a strong impact
on the greenhouse effect. A more efficient fuel consumption can be realized in Diesel and
lean-operated engines, i.e., in excess of air (oxygen). Here, the problem is the formation
of nitrogen oxides (NOx ).16 Because improvements of the combustion process itself are
not sufficient to meet future legislative limits, the development of a technique for the
aftertreatment of NOx is a hot research topic today [207–209, 88].
Aside from NOx reduction by plasma discharge techniques, all current concepts are
based on the utilization of catalysts. The demand of a reducing agent in the oxygen-
rich mixture can be met by the addition of such a reducing agent, for instance of
hydrocarbons derived from the fuel (HC-SCR17 technique).
In this section, the HC-SCR on a Pt/Al2 O3 catalyst is discussed using propylene as
reducing agent [88,89]. In the experimental18 and numerical investigations, a monolithic
catalyst was used and isothermal and steady state conditions were applied. Therefore,
only one single-channel needs to be analyzed again.
The numerical study also focuses on the application of various transport models for
the description of radial mass transport in the fluid (Chapter 2) and diffusion in the
washcoat (Chapter 3.3). Those models are compared to understand what kind of model
complexity is needed for an accurate description of the DeNOx catalytic converter.
In all cases simulated, a detailed surface reaction mechanism is used. This mecha-
nism consists of 62 reactions among 20 surface species and the gas phase species C3 H6 ,
NO, NO2 , N2 O, N2 , CO, CO2 , H2 , OH, H2 O, and O2 . The reaction kinetics is based on
the concept discussed in Chapter 3.2.2. For more detailed information on the reaction
mechanism the reader is referred to [88, 89]. The mechanism is given in Table B.6.
Gas-phase reactions are neglected because they are not significant at the conditions
(temperature, pressure, residence time), at which the converter is operated.

8.3.1 HC-SCR with C3 H6


In the experiment and simulation, a premixed gas, containing 500 ppm NO, 500 ppm
C3 H6 , and 5 vol.% O2 in nitrogen dilution, flows at a superficial velocity of 0.633 m/s
(standard conditions) in the channels of the monolithic structure. All the catalyst
parameters are given in Tab. 8.4.
Figure 8.9 presents the predicted and experimentally determined conversion of
C3 H6 , NOx , N2 O, and NO2 as function of temperature and at different catalyst lengths.
The simulation applies the plug-flow model (Chapters 2.2.3, 4.3.3 and Reference [88])
with an additional mass transfer coefficient, which accounts for radial mass transport.
Furthermore, the detailed washcoat model is used as discussed in Chapter 3.3.2.
In general, a good agreement between predicted and measured data could be
achieved, the minor differences can also be caused by deviations in the washcoat thick-
16
The 3WCC only works at stoichiometric conditions.
17
HC-SCR = hydrocarbon selective catalytic reduction
18
The experiments were carried out by E. Frank and W. Weisweiler at the University of Karlsruhe.
134 8. AUTOMOTIVE CATALYTIC CONVERTER

Monolith diameter 20 mm
Monolith length 5 - 60 mm
Single-channel diameter 1 mm
Number of channels 201
Catalytic material Pt/Al2 O3
Pt-loading 1 weight-%
Pt-dispersion 17%
Active Pt-surface 0.22 m2 /g
Active catalyst / geometric surface (Fcat/geo ) 60.44
Mean diameter of the Pt-particles 6.5 nm
BET-surface 27 m2 /g
Pore volume 0.7257 cm3 /g
Mean diameter of the micropores 4 nm
Porosity of micropores 39.5%
Mean diameter of the macropores 5.98 µm
Porosity of macropores 10%
Washcoat thickness 100µm

Table 8.4: Characteristic parameter of the catalyst used.

ness and catalyst dispersion in the different catalyst samples used. These parameters
were not experimentally determined for each single sample.
Although NO reduction at lean conditions can be achieved, most of NO (70-80%
selectivity) is converted into N2 O. The emission of laughing gas (N2 O) is currently not
limited by legislation but is is known as harmful greenhouse gas, which really is the
problem with the commercial application of the HC-SCR/Pt system. The reason for
the observed selectivity is the total high coverage of the Pt-surface while N(s)-coverage
is low. This favors the reaction NO(s) + N(s) → N2 O and lowers the probability
of N(s) recombination to molecular nitrogen. The selectivity to N2 formation could
be increased if the use of promoters led to a higher activation energy for nitrogen
desorption.
The correlation between conversion and surface coverage is revealed by Figs. 8.10
and 8.11, in which the species profiles in the single channel and the surface coverage
along the channel wall inside the washcoat are presented. The results are based on a
simulation using the Boundary-Layer model (Chapters 2.2.2 and 4.3.2) and the detailed
washcoat model (Chapter 3.3.2). The oxidation of C3 H6 occurs within the first 2 cm
channel length. The reduction of NO is limited to the first 1.5 cm channel length,
behind that, NO2 formation occurs. While the surface is mainly covered by CO at
the channel entrance, oxygen becomes the primary adsorbate downstream of an axial
position of 1.6 cm. Yet the oxygen poisoned surface is inactive for NO reduction and
leads to NO oxidation instead. Hence, the width of the reaction zone of propylene
oxidation determines the zone, in which NOx can be destroyed.
The axial surface coverage at a certain washcoat depth, 0.45µm in Fig. 8.11, really
is a simplification and does not describe the more complex picture of the interaction of
diffusion and reaction inside the washcoat. Figure 8.12 presents the surface coverages as
8.3 DeNOx catalyst 135

1.0 1.0

0.8 0.8

conversion
0.6
conversion

0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 400 500 600 700 800 900
T [K] T [K]

1.0 1.0

0.8 0.8
conversion

0.6
conversion

0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 400 500 600 700 800 900
T [K] T [K]

1.0 1.0

0.8 0.8
conversion

conversion

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 400 500 600 700 800 900
T [K] T [K]

Figure 8.9: Comparison of experimentally determined and numerically predicted con-


version as function of temperature and catalyst length (from top to bottom: 60 mm,
15 mm, 5 mm). N2 O and NO2 conversion means the fraction of that species formed
referring to the NOx conversion; NOx includes NO and NO2 .
136 8. AUTOMOTIVE CATALYTIC CONVERTER

C3H6 5.00x10
-04

0.050 -04
4.00x10
0.025
-04
r [cm]
3.00x10
0.000
-04
2.00x10
-0.025
-04
1.00x10
-0.050
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 -09
6.58x10
z [cm]

NOx 5.00x10
-04

0.050 -04
4.76x10
0.025
-04
4.52x10
r[cm]

0.000
-04
4.27x10
-0.025
-04
4.03x10
-0.050
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 -04
3.79x10
z [cm]
NO2 1.89x10
-04

0.050 -04
1.51x10
0.025
-04
r [cm]

1.13x10
0.000
-05
7.56x10
-0.025
-05
3.78x10
-0.050
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 +00
0.00x10
z [cm]

Figure 8.10: Mole fraction distribution in a single monolith channel at 570 K; other
conditions as before.

O(s)
100 CO(s)

-1
10 NO(s)
surface coverage

-2 C3H5(s)
10
Pt(s)

10-3

C3H5(s1)
-4
10
0 1 2 3 4 5 6
z [cm]

Figure 8.11: Surface coverage along the catalytic channel wall inside the washcoat,
0.45µm away from the fluid/washcoat boundary, at 570 K; other conditions as before.
8.3 DeNOx catalyst 137

function of axial and radial position inside the washcoat. The CO(s) and O(s) profiles
reveal that the shift from CO(s) to O(s) coverage depends on the radial position inside
the only 100µm thick washcoat layer. Even at the washcoat entrance, a primarily
oxygen covered surface exists in 80µm washcoat depth. The other profiles reveal that
there are axial and radial variations of the surface coverages and, hence, the reaction
rates vary inside the washcoat.

8.3.2 Comparison of transport models


Now the application of different transport models of various complexity for modeling
automotive catalytic converters will be emphasized. In particular, there is a strong
impact of radial mass transport in the fluid and in the washcoat at high space velocities
and at the catalyst entrance.
If radial mass transport is neglected in the fluid model, total conversion is over-
predicted by the numerical simulation. For example, Fig. 8.13 compares a pure Plug-
Flow simulation and a Plug-Flow simulation with mass transport coefficient and de-
tailed washcoat model. At 5 mm channel length, which corresponds to a space velocity
of 216000 h−1 , propylene is already completely gone, which clearly contradicts experi-
mental observation and the simulation with more complex transport models.
Because the gradients of the species concentration are relatively small at the condi-
tions chosen, it is justified to use the PF model with mass transfer coefficient. Figure
8.14 compares simulations using the BL model and the PF model with mass transfer
coefficients, both simulations apply the same detailed washcoat model. There are no
visible differences in the conversion data predicted. In contrast to that, a simulation us-
ing the PF model without mass transfer coefficients fails to predict the right propylene
conversion as shown in Fig. 8.15.
In Chapter 3.3.2, a simple washcoat model based on an effectiveness factor was
discussed. In Fig. 8.16, two simulations are compared, which both use the PF-model
with mass transport coefficients but different washcoat models: In the simple washcoat
model, the effectiveness factor was determined for the propylene species because it is the
most crucial reacting species as discussed above. The detailed washcoat model is based
on the solution of an additional one-dimensional reaction-diffusion equation for each
chemical species (Chapter 3.3.2). While the differences remain relatively small for the
predicted C3 H6 conversion, the predicted NO2 formation considerably differs between
simple and detailed model. This fact leads to a general problem when using effectiveness
factors: After light-off of C3 H6 oxidation, the C3 H6 -consumption is primarily limited
by radial mass transport, while the NO oxidation is still controlled by chemical kinetics.
In such situations, the simplified washcoat model cannot be applied.
Summarizing, it is difficult to a-priori judge whether or not simplified transport
models can be applied. The application of the simplified model has to be checked
from case to case. If simplified models are accurate enough, their application can save
tremendous computing time. For instance, the simplified washcoat model works fine
in the transient 3WCC simulations if the reference species for the effectiveness factor
is chosen according to the species concentrations. As shown here, it fails in a similar
system. Therefore, it is pointed out that the most detailed model should always be
138 8. AUTOMOTIVE CATALYTIC CONVERTER

CO(s) O(s)

1
surface coverage

surface coverage
0.6

0.4
0.5
0 0
0.2 1 1
2 2
0 3
m] z [c 3 100
4 z [c m] 4
r [µ 50 5 50
m] m]
r [µ
5
100 0

NO(s) N(s)

-04
surface coverage

2.0x10
surface coverage

0.1

0 1.0x10-04 0
0.05
1 1
2 +00
2
0 3 0.0x10 0 3
m] m]
4 z [c 4 z [c
r [µ 50 5 r [µ 50
m] m] 5
100 100

N2O(s) NO2(s)
-06
1.0x10
surface coverage
surface coverage

-09
3.0x10
-07
2.0x10
-09 5.0x10
0 0
-09 1 1
1.0x10 2
2
0 3 z [c 3 100
m] m]
4 z [c 4
50
r [µ 50 m]
r [µ
m] 5 5
100 0

Figure 8.12: Surface coverage with CO(s), O(s), NO(s), N(s), N2 O, and NO2 as func-
tion of axial and radial position in the washcoat layer; conditions are as before; the
fluid/washcoat boundary is at r = 0.
8.3 DeNOx catalyst 139

applied as a first step. Especially, if the simplified models involve unknown parameters,
e.g., for the description of the reaction kinetics, it is very risky to fit these parameters
to experimental data without having checked the accuracy of all the transport models.
Otherwise the fitted parameters will compensate for the errors in the transport model,
and conclusions drawn from the simulations are arbitrary.

1.0 1.0

0.8 0.8

conversion
conversion

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 0 2 4 6
T [K] z [cm]

Figure 8.13: Impact of radial mass transport on conversion as function of temperature


at a catalyst length of 5 mm (left) and as function of axial position in the monolithic
channel at a temperature of 570 K; other conditions are chosen as before.

1.0 1.0

0.8 0.8
conversion

conversion

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 2 4 6 0 2 4 6
z [cm] z [cm]

Figure 8.14: Comparison of conversion as function of axial position using the BL model
and PF model with mass transfer coefficients (both simulations include the detailed
washcoat model) at a temperature of 570 K (left) and 800 K (right); other conditions
are chosen as before.
140 8. AUTOMOTIVE CATALYTIC CONVERTER

1.0 1.0

0.8 0.8

conversion
conversion

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 400 500 600 700 800 900
T [K] T [K]

Figure 8.15: Impact of the radial mass transport in the fluid phase. Comparison
of experimentally determined conversion vs. temperature with numerical simulations
using the PF model with and without mass transfer coefficients (both simulations
include the detailed washcoat model); other conditions are chosen as before.

1.0 1.0

0.8 0.8
conversion

conversion

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
400 500 600 700 800 900 0 2 4 6
T [K] z [cm]

Figure 8.16: Comparison of the washcoat models: Numerically predicted conversion


vs. temperature for a 6 cm long catalyst (left) and vs. axial position at 570 K (right)
for simulations using the PF model with mass transfer coefficients; other conditions
are chosen as before.
141

9 Conclusions
Experimental and numerical investigations were carried out to achieve a better under-
standing of the interactions between mass and heat transport and homogeneous and
heterogeneous chemical reactions in catalytic reactors. Gaseous chemically reacting
flows with heterogeneous reactions on solid surfaces are considered. The objective of
this research was the development of models and computer programs for the numerical
simulation of heterogeneous reactive flows.
In the experiments, catalytic reactors were studied for the partial and complete
oxidation of light alkanes at short contact times. The noble metals rhodium and plat-
inum served as catalyst. The conversion of the reactants and product selectivity were
determined by gas chromatography and mass spectrometry; the latter technique can
also be used for transient measurements. The temperature was determined by thermo-
couples. The reactor behavior was studied at varying conditions such as the flow rate,
temperature, dilution by an inert gas, and hydrocarbon/oxygen ratio. The experimen-
tal investigations served as basis for the development of reliable models. Experimental
data from collaborating research groups were also used as further independent sources
to judge the quality of the models applied.
A concept for modeling catalytic reactors was presented, in which all physical and
chemical processes were described as detailed as possible. These models were combined
to understand and, eventually, optimize the behavior of the catalytic reactor. The flow
field was modeled by the multi-dimensional Navier-Stokes equations coupled with an
enthalpy equation and one further governing equation for each chemical species. The
model includes temperature- and composition-dependent transport coefficients. The
chemistry in the gas phase was described by elementary-reaction mechanisms. The
heterogeneous reactions of gas-phase species on the solid surfaces were described by
multi-step reaction mechanisms, which were also based on the elementary processes.
In most previous studies, the numerical simulation of catalytic reactors is based
on simplifying assumptions, either for the flow field or the chemistry. For instance,
the flow field in tubular reactors is frequently assumed to be plug-flow-like, neglecting
any effect of radial mass transport, or the chemical processes are described by global
reactions with kinetic data fit to limited experimental measurements. Those models
can hardly be used for prediction or optimization of the reactor behavior.
In the concept presented in this work, the numerical simulations were based on the
most accurate models available. For instance, a monolithic reactors with rectangular-
shaped channel cross-sections were first modeled by a three-dimensional flow field sim-
ulation. Then, the results of these detailed simulations were compared with those of
simulations applying simpler models. This comparison allowed the potentials and lim-
itations of the simplifying assumptions to be recognized. It was found that the channel
flow in a catalytic monolith can be described by an axisymmetric flow field. Further-
more, axial diffusion can be neglected at high flow rates leading to the Boundary-Layer
equations. The simple Plug-Flow model failed to predict the reactor behavior cor-
rectly. It was also shown that an extended Plug-Flow model, which includs a radial
mass-transfer coefficient, works well for relatively slow reactions such as occurring in
automotive catalytic converters. However, this model cannot be used for very fast
142 9. CONCLUSIONS

reactions as in catalytic combustion monoliths, because no correlation was found to


estimate proper mass-transfer coefficients.
In the models developed, the heterogeneous reactions of the gas-phase species on
catalytic surfaces are described by multi-step reaction mechanisms, based on the mole-
cular processes at the gas-surface interface. The model takes the coverage dependence
of the reaction kinetics into account. The surface coverage with adsorbed species de-
pends on the position in the reactor and, therefore, is computed at any position in
the reactor. This approach is superior to frequently used global reaction schemes, be-
cause it permits the prediction of the reactor behavior at varying external conditions,
and can eventually be used for optimization. The developed surface reaction mecha-
nisms are based on the mean-field approximation, in which the surface is assumed to
be homogeneously covered on a microscopic scale, and the kinetic data are averaged
over microscopic local inhomogeneities. A strategy was presented for the development
of detailed surface reaction mechanisms under the mean-field approximation. The
limitations of this approach were also discussed. The mean-field approximation, for
instance, fails if the concrete lateral interactions of the adsorbates including diffusion
and recrystallization phenomena have to be explicitly taken into account. A more so-
phisticated model is a Monte-Carlo (MC) simulation of the chemical processes on the
surface, coupled with the reactive flow. MC simulations have already been used for
simple reaction systems (e.g., CO oxidation on Pt) and flow configurations (e.g. 1D
stagnation point flows) [70]. However, MC simulations have not yet been coupled with
multi-dimensional flow configurations due to the tremendous computing time needed.
Further research on the heterogeneous kinetics shall focus on the determination of
reaction schemes and kinetic data at atmospheric pressure and on real catalysts to
bridge the pressure and materials gap between most of the surface science experiments
and applied heterogeneous catalysis. High pressure scanning tunnel microscopy and
nonlinear optic techniques such as sum frequency generation vibrational spectroscopy
[5, 6] will be of great use on this path.
The surface of the solid catalyst support is frequently coated with a layer of high
surface area material (washcoat), in which the catalyst is dispersed. Transport of
chemical species inside the washcoat and the interaction of transport and chemical
reactions can be crucial in determining the behavior of the catalytic reactor. Two
models were applied for the description for the chemical and transport processes in
washcoats. The simpler washcoat model calculates an effectiveness factor that limits
the overall reaction rate. This model can be applied if the reaction rate of one chemical
species determines the whole process. In the more sophisticated model, a set of one-
dimensional reaction-diffusion equations is solved in combination with the flow field
description. Both models were coupled with the equations for the description of the
surrounding flow field in the reactor and applied for the numerical prediction of conver-
sion in automotive catalytic converters. The simpler model showed major deficiencies;
the conversion was accurately predicted only for that species, on which the calculation
of the effectiveness factor was based.
Several computational tools have been developed for the numerical simulation
of reactive flows including heterogeneous chemical reactions. The software package
DETCHEM is a module for the application of multi-step reaction mechanisms in CFD
143

codes such as FLUENT. The code DETCHEMCHANNEL simulates the reactive flow in
cylindrical and annular channel with catalytic active walls based on a Boundary-Layer
model. DETCHEMMONOLITH was written for the numerical simulation of the tran-
sient behavior of catalytic monoliths. In that code, the three-dimensional heat balance
equation for the monolithic structure is solved in combination of the simulation of the
reactive flows in the single channels.
The new approach for modeling catalytic reactors and the computational tools were
applied to investigate several catalytic reaction systems that have recently attracted
strong scientific and technological interest.
The catalytic partial oxidation of methane to synthesis gas on rhodium coated
monoliths at short contact times is a promising route for natural gas conversion into
more useful chemicals. The numerical simulation of this autothermally operated reac-
tor revealed fast variations of all transport properties at the catalyst entrance, where
the initially cold methane/oxygen mixture enters a ∼1300 K hot monolith. The con-
version of methane starts right at the catalyst entrance, where a strong competition
between partial and complete oxidation occurs. After 1 mm catalyst length, oxygen is
completely consumed and steam reforming is the determining reaction path. Experi-
mentally determined and numerically predicted conversion and selectivity data agreed
very well. The formation of synthesis gas was studied at varying conditions, e.g., the
syngas selectivity and methane conversion decrease with increasing flow rate. Homo-
geneous gas-phase reactions become significant at pressures above 10 bar. They lead
to a decrease in syngas selectivity and also increase the risks of flames. Light-off of the
reactor was studied by a transient two-dimensional simulation of the whole monolith,
which elucidated the impact of the temperature distribution in the monolithic structure
on overall conversion. Reactor design and scale-up can benefit from the application of
the models and computational tools developed.
The oxy-dehydrogenation of ethane to ethylene was investigated in platinum coated
monoliths at contact times of ∼5 ms. The observed performance of the ethane reac-
tor is the result of coupled heterogeneous and homogeneous chemical processes. The
valuable ethylene product of this reactor results from both homogeneous and heteroge-
neous dehydrogenation of ethane, the relative contributions of each depending on the
reactor conditions. Heat is required to drive this highly endothermic dehydrogenation.
It is concluded that this heat is provided through heterogeneous oxidation reactions
that occur very near the front of the catalytic section of the reactor. Conditions that
result in optimum reactor performance are most strongly characterized by a picture of
heterogeneous oxidation in conjunction with a relatively large contribution from homo-
geneous ethane dehydrogenation. The addition of H2 to the inlet feed of this reactor
results in surface hydrogen oxidation at the expense of surface carbon oxidation, pro-
ducing even more highly localized heat release into the gas-phase. Homogeneous ethane
dehydrogenation is also enhanced, while heterogeneous decomposition and oxidation
to CO and CO2 is diminished. The selectivity to ethylene increases because selectivity
to CO and CO2 is decreased. Desorption of radical species does not play an important
role in the initiation of homogeneous reactions under the conditions studied. Localized
heat release and high peak temperatures in the reactor are responsible for the short
residence times required to initiate the homogeneous chemistry vital to the reactor
144 9. CONCLUSIONS

performance. As such, it is important to incorporate full heat and mass transport in


the simulation in order to capture the spatial distribution of homogeneous reactions,
and an accurate wall temperature profile.
Further catalytic-supported gas-phase partial oxidation reactions of paraffines such
as propane, butane, pentane, and cyclohexane have recently attracted wide interest
also. The reactors used are noble-metal-coated monoliths and Pt/Rh wire gauzes.
The homogeneous and heterogeneous chemical reactions in those systems are strongly
superimposed by heat transfer effects, in particular by fast thermal quenching of the
products, which leads to a highly non-equilibrium selectivity. The synthesis of olefins,
aldehydes, and oxygenates by partial oxidation of these alkanes is not yet understood.
The present work, that dealt with partial oxidation of the light alkanes methane and
ethane, also represents a first step towards a better understanding of the partial oxi-
dation of the higher alkanes. Future work will also consider the more complex three-
dimensional flow fields and the development of reliable reaction mechanisms for those
reactions. This modeling calls not only for heterogeneous reaction schemes for the
oxidation of higher hydrocarbons but also for homogeneous gas-phase reaction mecha-
nisms at fuel-rich conditions and low temperatures.
Light-off of heterogeneous reactions in the catalytic combustion of natural gas de-
mands sophisticated techniques, because the catalytic reaction ignites at temperatures
above 800 K. The ignition temperature can be lowered by adding hydrogen to the
inlet gas, since hydrogen oxidation on the catalyst (Pt, Pd) happens almost at room-
temperature. Hydrogen-assisted ignition of methane oxidation in a platinum coated
monolith was numerically simulated by the solution of the transient three-dimensional
temperature equations of the solid monolith structure coupled with detailed models
for the chemical reactions and flow field in the single channels of the monolith. It was
found that, shortly after light-off, a large amount of methane is converted in the hot
inner channels, while almost no methane is consumed in the colder exterior channels.
The investigated catalytic radiant burner is characterized by a more complex two-
dimensional flow field and several modes of heat transfer, including external heat loss by
thermal radiation, internal surface-to-surface radiation, and heat conduction through
solid reactor walls. Therefore, several days of CPU time were needed for the numerical
simulation of the burner with the CFD code FLUENT coupled with DETCHEM. How-
ever, the efforts were rewarded by a deeper insight into the interaction of flow, heat
transfer, and chemistry in the burner. The predicted temperature, conversion and se-
lectivity at the burner exit as well as the temperature of the radiative burner plate
agreed well with the experimentally determined data. The computational tool can
now be used for burner design and optimization. Future work will also take gas-phase
chemistry into account, which might be necessary to accurately describe the formation
of pollutants such as CO.
The performance of automotive catalytic converters was investigated. Here, mass
transport, heat transfer, and heterogeneous reaction kinetics are superimposed by dif-
fusion of the reactive species in the pores of the washcoat. First, the simultaneous
oxidation of carbon monoxide and propylene as a sample hydrocarbon, and the reduc-
tion of nitrogen oxides was studied at steady-state conditions for a commercially used
three-way catalyst. The numerical simulation of the behavior of a single channel of
145

the monolithic structure is based on a two-dimensional flow field description coupled


with a model for washcoat diffusion and a surface reaction mechanism on platinum and
rhodium consisting of 65 reactions. Conversions of C3 H6 , CO, and NO were studied
as functions of temperature at lean, stoichiometric, and rich conditions. The model,
for instance, explained the increase of NO conversion with increasing temperature at
low temperature and the decrease at higher temperatures by the variation of the sur-
face coverage. A good agreement between experimentally determined and numerically
predicted conversion was achieved. Only in the rich regime, larger deviations were
recorded for the conversion of C3 H6 , which may be caused by partial oxidation steps
missing in the reaction mechanism.
For the first time, a transient two-dimensional simulation of an automotive catalytic
converter was carried out with detailed models for the chemical reactions and the mass
and heat transport. The model and computer code can now be applied for the detailed
numerical simulation of an entire driving cycle that is used by legislation for pollutant
emission control of automotive vehicles. The implementation of models describing
storage features of the converter will be one of the next steps in transient simulation
of automotive catalytic converters.
Different transport models were applied for the description of hydrocarbon selective
catalytic reduction of nitrogen oxides (DeNOx catalyst). The comparison of numeri-
cally predicted and experimentally derived conversion as function of temperature led to
a model discretization. While the Boundary-Layer model and the extended Plug-Flow
model, which includes a radial mass transfer coefficient, can be used as simplified mod-
els for the flow field description, the pure Plug-Flow model fails, because it neglects
the transport limitation by radial mass transfer. Furthermore, the simple washcoat
model, based on an effectiveness factor, cannot be applied because it neglects that the
overall reaction is determined by the reaction rate of various species depending on the
external conditions such as temperature.
While this research seems to have shed considerable light on the physical and chem-
ical processes relevant to the performance of the catalytic reactors discussed, it is also
recognized that chemical mechanisms based on a limited set of experimental data can
be neither unique nor complete in their description of the detailed kinetics. More ex-
periments and kinetic mechanism development are required to achieve a comprehensive
set of rate coefficients. Nevertheless, a point seems to have been reached where opti-
mization of catalytic reactors based on detailed models for the flow field, heat transfer,
and reaction kinetics can be carried out. As a first step, homogeneous reaction sys-
tems have recently been optimized. For instance, optimal temperature profiles led to
an increase in the yields of ketene formation by homogeneous catalytic pyrolysis of
acetic acid [210], and of ethylene by homogeneous oxidative coupling of methane [211].
Currently, the optimization code PRSQP by Schulz [212] is coupled with a model for a
transient one-dimensional stagnation point flow on a catalytic active plate [213].
REFERENCES 147

References
[1] K. Christmann. Introduction to Surface Physical Chemistry. Topics in Physical Chemistry 1. Springer, New
York, 1991.

[2] D.A. Hickman and L.D. Schmidt. Production of syngas by direct catalytic oxidation of methane. Science, 259,
343–346, 1993.

[3] A.S. Bodke, D.A. Olschki, L.D. Schmidt, and E. Ranzi. High Selectivity to Ethylene by Partial Oxidation of
Ethane. Science, 285, 712–715, 1999.

[4] G. Ertl. Elementary Steps in Heterogeneous Catalysis. Angew. Chem. Int. Ed. Engl., 29, 1219–1227, 1990.

[5] K.Y. Kung, P. Chen, F. Wei, Y.R. Shen, and G.A. Somorjai. Sum-frequency generation spectroscopy study of
CO adsorption and dissociation on Pt(111) at high pressure and temperature. Surf. Sci., 463, 627–633, 2000.

[6] U. Metka, M.G. Schweitzer, H.-R. Volpp, J. Wolfrum, and J. Warnatz. In-situ detection of NO chemisorbed on
platinum using infrared-visible Sum-Frequency Generation (SFG). Zeitschr. f. Phys. Chem., 214, 865–888, 2000.

[7] Technology Vision 2020: The U.S. Chemical Industry, 1996. Report published by The American Chemical Society,
American Institute of Chemical Engineers, The Chemical Manufactures Association, The Council for Chemical
Research, The Synthetic Organic Chemical Manufactures Association.

[8] M. Huff and L.D. Schmidt. Production of Olefins by Oxidative Dehydrogenation of Propane and Butane over
Monoliths at Short Contact Times. J. Catal., 149, 127–141, 1993.

[9] A. Beretta, L. Piovesan, and P. Forzatti. An investigation on the role of a Pt/Al2 O3 catalyst in the oxidative
dehydrogenation of propane in annular reactor. J. Catal., 184, 455–468, 1999.

[10] D.A. Goetsch and L.D. Schmidt. Microsecond catalytic partial oxidation of alkanes. Science, 271, 1560–1562,
1996.

[11] D.I. Iordanoglou. Rapid Oxidation of Light Alkanes in Single Gauze Reactors. PhD thesis, Department of
Chemical Engineering and Materials Science, University of Minnesota, Minneapolis, MN, USA, 1998.

[12] R. Lødeng, O.A. Lindvåg, S. Kvisle, H. Reier-Nielsen, and A. Holmen. Short contact time oxidative dehydro-
genation of C2 and C3 alkanes over noble metal gauze catalysts. Appl. Catal. A, 187, 25–31, 1999.

[13] R.P. O’Connor and L.D. Schmidt. Catalytic partial oxidation of cyclohexane in a single-gauze reactor. J. Catal,
191, 245–256, 2000.

[14] R.P. O’Connor, L.D. Schmidt, and O. Deutschmann. Detailed Simulations of the Millisecond Partial Oxidation
of Cyclohexane in Single-Gauze Reactors: Coupled Chemistry and Fluid Dynamics. AIChE J. (submitted).

[15] R.B. Bird, W.E. Stewart, and E.N. Lightfoot. Transport Phenomena. John Wiley & Sons, Inc., New York, 1960.

[16] C.N. Satterfield. Mass Transfer in Heterogeneous Catalysis. MIT Press, Cambridge, MA, 1970.

[17] R. Aris. The Mathematical Theory of Diffusion and Reaction in Permeable Catalysts. Clarendon Press, Oxford,
1975.

[18] S.V. Patankar. Numerical Heat Transfer and Fluid Flow. McGraw-Hill, New York, 1980.

[19] M. Baerns, H. Hofmann, and A. Renken. Chemische Reaktionstechnik. Georg Thieme Verlag Stuttgart, New
York, 1992.

[20] J. Warnatz, R.W. Dibble, and U. Maas. Combustion, Physical and Chemical Fundamentals, Modeling and
Simulation, Experiments, Pullutant Formation. Springer-Verlag, New York, 1996.

[21] F. Keil. Diffusion und Chemische Reaktionen in der Gas-Feststoff-Katalyse. Springer-Verlag, Berlin, 1999.

[22] J. Warnatz. Influence of Transport Models and Boundary Conditions on Flame Structure. In N. Peters and
J. Warnatz, editors, Numerical Methods in Flame Propagation. Friedr. Vieweg and Sohn, Wiesbaden, 1982.

[23] R.J. Kee, J. Warnatz, and J.A. Miller. A FORTRAN Computer Code Package for the Evaluation of Gas Phase
Viscosities, Heat Conductivities, and Diffusion Coefficients. Sandia National Laboratories Report, SAND83-8209,
1983.
148 REFERENCES

[24] G. Dixon-Lewis. Flame structure and flame reaction kinetics. Proc. Roy. Soc. A, 307, 111–135, 1968.

[25] R.J. Kee, G. Dixon-Lewis, J. Warnatz, M.E. Coltrin, and J.A. Miller. A Fortran Computer Code Package
for the Evaluation of Gas-Phase Multicomponent Transport Properties. Sandia National Laboratories Report,
SAND86-8246, 1986.

[26] T.G. Cowling S. Chapman. Mathematical Theory of Non-Uniform Gases. Cambridge University Press, Second
Edition, 1951.

[27] M.W. Chase, C.A. Davis, J.R. Downey, D.J. Frurip, R.A. McDonald, and A.N. Syverud. JANAF Thermochemical
Tables, Third Edition. J. Phys. Chem. Ref. Data, 14, 1985.

[28] R.J. Kee, F.M. Rupley, and J.A. Miller. The Chemkin Thermodynamic Database. Sandia National Laboratories
Report, SAND87-8215, 1987.

[29] A. Burcat. Thermochemical data for combustion. In W.C. Gardiner, editor, Combustion chemistry. Springer,
New York, 1984.

[30] S.W. Benson. Thermochemical Kinetics. John Wiley & Sons, New York, 1976.

[31] S. Gordon and B.J. McBridge. Computer Programm for Calculation of Complex Chemical Equilibrium Compo-
sitions, Rocket Performance, Incident and Reflected Shocks and Chapman-Jouguet Detonations. NASA SP-273,
1971.

[32] E.S. Lox and B.H. Engler. Environmental catalysis. In G. Ertl, H. Knoezinger, and J. Weitkamp, editors,
Handbook of Heterogeneous Catalysis. Wiley-VCH, Weinheim, 1997.

[33] L.L. Raja, R.J. Kee, O. Deutschmann, J. Warnatz, and L.D. Schmidt. A Critical Evaluation of Navier-Stokes,
Boundary-Layer, and PLug-Flow Models of the Flow and Chemistry in a Catalytic-Combustion Monolith. Catal.
Today, 59, 47–60, 2000.

[34] H. Schlichting. Boundary-Layer Theory. 6th ed., McGraw-Hill, New York, 1968.

[35] K.H. Homann. Reaktionskinetik. Steinkopff, Darmstadt, 1975.

[36] S. Arrhenius. Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Säuren. Z. Phys. Chem.,
4, 226, 1889.

[37] D.L. Baulch, C.J. Cobos, R.A. Cox, C. Esser, P. Frank, Th. Just, J.A. Kerr, M.J. Pilling, J. Troe, R.W. Walker,
and J. Warnatz. Evaluated Kinetic Data for Combustion Modelling. J. Phys. Chem. Ref. Data, 23, 850–851,
1994.

[38] W.C. Gardiner, editor. Combustion Chemistry. Springer, New York, 1984.

[39] U. Maas and S. Pope. Simplifying Chemical Kinetics: Intrinsic Low-Dimensional Manifolds in Composition Space.
Combust. Flame, 88, 239–264, 1992.

[40] M.E. Coltrin, R.J. Kee, G.H. Evans, E. Meeks, F.M. Rupley, and J.F. Grcar. SPIN: A Fortran Program for
Modeling One-Dimensional Rotating Disk / Stagnation - Flow Chemical Vapor Deposition Reactors. Sandia
National Laboratories Report, SAND91-8003, 1991.

[41] E. Meeks, R.J. Kee, and D.S. Dandy. Computational Simulation of Diamond Chemical Vapor Deposition in
Premixed C2 H2 /O2 /H2 and CH4 /O2 - Strained Flames. Combust. Flame, 92, 144–160, 1993.

[42] B. Ruf, F. Behrendt, O. Deutschmann, and J. Warnatz. Simulation of reactive flow in filament-assisted diamond
growth including hydrogen surface chemistry. J. Appl. Phys., 79, 7256–7263, 1996.

[43] B. Ruf, F. Behrendt, O. Deutschmann, S. Kleditzsch, and J. Warnatz. Modeling of chemical vapor deposition of
diamond films from acetylene-oxygen flames. Proc. Combust. Inst., 28, (in press), 2000.

[44] S. Kleditzsch and U. Riedel. Sensitivity Studies of Silicon Etching in Chlorine/Argon Plasmas. J. Vac. Sci.
Technol. A, 18, 2130–2136, 2000.

[45] O. Deutschmann, R. Schmidt, F. Behrendt, and J. Warnatz. Numerical modeling of catalytic combustion. Proc.
Combust. Inst., 26, 1747–1754, 1996.

[46] R.E. Hayes and S.T. Kolaczkowski. Introduction to Catalytic Combustion. Gordon and Breach Science Publ.
Amsterdam, 1997.
REFERENCES 149

[47] M.E. Coltrin, R.J. Kee, and F.M. Rupley. SURFACE CHEMKIN (Version 4.0): A Fortran Package for Analyzing
Heterogeneous Chemical Kinetics at a Solid-Surface - Gas-Phase Interface. Sandia National Laboratories Report,
SAND90-8003B, 1990.

[48] U. Metka. Untersuchung von Teilschritten der heterogenen Reaktion von CO mit NO an Platin mittels der
Summenfrequenz-Spektroskopie. Dissertation. Naturwissenschaftlich-Mathematische Gesamtfakultät. Ruprecht-
Karls-Universität Heidelberg, 2000.

[49] D.K. Zerkle, M.D. Allendorf, M. Wolf, and O. Deutschmann. Understanding Homogeneous and Heterogeneous
Contributions to the Platinum-Catalyzed Partial Oxidation of Ethane in a Short-Contact-Time Reactor. J. Catal,
196, 18–39, 2000.

[50] P. Aghalayam, Y.K. Park, and D.G. Vlachos. Construction and Optimization of Complex Surface-Reaction
Mechanisms. AIChE J., 46, 2017–2029, 2000.

[51] C.T. Campbell, G. Ertl, H. Kuipers, and J. Segner. A molecular beam of the adsorption and desorption of oxygen
from a Pt(111) surface. Surf. Sci., 107, 220–236, 1981.

[52] J.A. Dumesic, D.F. Rudd, L.M. Aparicio, J.E. Rekoske, and A. A. Trevino. The Microkinetics of Heterogeneous
Catalysis. American Chemical Society, Washington, DC, 1993.

[53] J.R. Chen and R. Gomer. Mobility of oxygen on the (110) plane of tungsten. Surf. Sci., 79, 413–444, 1979.

[54] S.C. Wang and R. Gomer. Diffusion of hydrogen, deuterium, and tritium on the (110) plane of tungsten. J.
Chem. Phys., 83, 4193–4209, 1985.

[55] D.R. Mullins, B. Roop, S.A. Castello, and J.M. White. Isotope Effects in Surface Diffusion. Hydrogen and
Deuterium on Ni(100). Surf. Sci., 186, 67–74, 1987.

[56] E.G. Seebauer, A.C.F. Kong, and L.D. Schmidt. Adsorption and Desorption of NO, CO and H2 on Pt(111):
Laser-induced Thermal Desorption Studies. Surf. Sci., 176, 134–156, 1986.

[57] E. Shustorovich. Chemisorption phenomena: analytic modeling based on perturbation theory and bond-order
conservation. Surf. Sci. Rep., 6, 1–63, 1986.

[58] E. Shustorovich. Bond making and breaking on transition-metal surfaces: theoretical projections based on bond-
order conservation. Surf. Sci., 176, L863–L872, 1986.

[59] E. Shustorovich and H. Sellers. The UBI-QEP Method: A Practical Theoretical Approach to Understanding
Chemistry on Transition Metal Surfaces. Surf. Sci. Rep., 31, 1–119, 1998.

[60] P. Sautet and J. Paul. Low temperature adsorption of ethylene and butadiene on platinum and paladium surfaces.
Catal. Lett., 9, 245–260, 1991.

[61] R.A. van Santen. Theoretical Heterogeneous Catalysis. World Scientific, Singapore, 1991.

[62] R.A. van Santen and M. Neurock. Theory of surface-chemical reactivity. In G. Ertl, H. Knoezinger, and
J. Weitkamp, editors, Hamdbook of Heterogeneous Catalysis, pages 991–1004. Wiley-VCH, Weinheim, 1997.

[63] T. Wahnström, E. Fridell, S. Ljungström, B. Hellsing, B. Kasemo, and A. Rosén. Determination of the Activation
Energy for OH Desorption in the H2 +O2 Reaction on Polycrystalline Platinum. Surf. Sci., 223, L905–912, 1989.

[64] W.R. Williams, C.M. Marks, and L.D. Schmidt. Steps in the Reaction H2 +O2 =H2 O on Pt: OH Desorption at
High Temperatures. J. Phys.Chem., 96, 5922–5931, 1992.

[65] F. Behrendt, O. Deutschmann, U. Maas, and J. Warnatz. Simulation and sensitivity analysis of the heterogeneous
oxidation of methane on a platinum foil. J. Vac. Sci. Technol. A, 13 (3), 1373–1377, 1995.

[66] O. Deutschmann. Modellierung von Reaktionen an Oberflächen und deren Kopplung mit chemisch reagierenden
Strömungen. Dissertation. Naturwissenschaftlich-Mathematische Gesamtfakultät. Ruprecht-Karls-Universität
Heidelberg, 1996.

[67] L.L. Raja, R.J. Kee, and L.R. Petzold. Simulation of the transient, compressible, gas-dynamic, behavior of
catalytic-combustion ignition in stagnation flows. Proc. Combust. Inst., 27, 2249–2257, 1998.

[68] R. Kissel-Osterrieder, F. Behrendt, J. Warnatz, U. Metka, H.-R. Volpp, and J. Wolfrum. Experimental and
theoretical investigation of the CO oxidation on platinum: Bridging the pressure and material gap. Proc. Combust.
Inst., 28, 1341–1348, 2000.
150 REFERENCES

[69] S.J. Lombardo and A.T. Bell. A Monte Carlo model for the simulation of the temperature-programmed desorption
spectra. Surf. Sci., 206, 101–123, 1988.

[70] R. Kissel-Osterrieder, F. Behrendt, and J. Warnatz. Detailed modeling of the oxidation of CO on platinum: A
monte-carlo model. Proc. Combust. Inst., 27, 2267–2274, 1998.

[71] N. Wakao and J.M. Smith. Diffusion in catalyst pellets. Chem. Eng. Sci., 17, 825, 1962.

[72] C. Rieckmann and F.J. Keil. Simulation and experiment of multicomponent diffusion and reaction in three-
diemnsional networks. Chem. Eng. Sci., 54, 3485–3493, 1999.

[73] F.J. Keil. Diffusion and reaction in porous networks. Catalysis Today, 53, 245–258, 2000.

[74] Fluent, Version 5. Fluent Inc., Lebanon, New Hampshire, 1998.

[75] R.J. Kee, F.M. Rupley, and J.A. Miller. CHEMKIN-II: A Fortran Chemical Kinetics Package for the Analysis
of Gas-Phase Chemical Kinetics. Sandia National Laboratories Report, SAND89-8009, 1989.

[76] M.E. Coltrin, H.K. Moffata, R.J. Kee, and F.M. Rupley. CRESLAF (Version 4.0): A Fortran Pogram for
Modeling Laminar, Chemically Reacting, Boundary-Layer Flow in Cylindrical or Planar Channels. Sandia
National Laboratories Report, SAND93-0478, 1993.

[77] R.J. Kee, F.M Rupley, J.A. Miller, M.E. Coltrin, J.F. Grcar, E. Meeks, K.H. Moffat, A.E. Lutz, G. Dixon-Lewis,
M.D. Smooke, J. Warnatz, G.H. Evans, R.S. Larson, R.E. Mitchell, L.R. Petzold, W.C. Reynolds, M. Caracotsios,
W.E. Stewart, P. Glarborg, C. Wang, and O. Adigun. CHEMKIN Collection, Version 3.6. Reaction Design, Inc.,
San Diego, 2000.

[78] CFD Research Corporation, Huntsville, AL, USA. http://www.cfdrc.com/, 02.05.2001.

[79] Recation Design. http://www.cd.co.uk/, 02.05.2001.

[80] K.D. Devine, G.L. Hennigan, S.A. Hutchinson, A.G. Salinger, J.N. Shadid, and R.S.Tuminaro. High Performance
MP Unstructured Finite Element Simulation of Chemically Reacting Flows. Proc. of SC97, San Jose, CA, Nov.
15-21.

[81] O. Deutschmann, C. Correa, S. Tischer, D. Chatterjee, and J. Warnatz. DETCHEM, User Manual, Version
1.4.1. http://reaflow.iwr.uni-heidelberg.de/∼dmann/DETCHEM.html, 2001.

[82] U. Maas. Mathematische Modellierung instationärer Verbrennungsprozesse unter Verwendung detaillierter


Reaktionsmechanismen. Dissertation. Naturwissenschaftlich-Mathematische Gesamtfakultät. Ruprecht-Karls-
Universität Heidelberg, 1989.

[83] F. Behrendt. Simulation laminarer Gegenstromdiffusionsflammen unter Verwendung detaillierter Reaktions-


mechanismen. Dissertation. Naturwissenschaftlich-Mathematische Gesamtfakultät. Ruprecht-Karls-Universität
Heidelberg, 1989.

[84] J. Warnatz. Resolution of Gas Phase and Surface Combustion Chemistry into Elementary Reactions (Invited
Lecture). Proc. Combust. Inst., 24, 553–579, 1992.

[85] O. Deutschmann and L. D. Schmidt. Modeling the partial oxidation of methane in a short-contact-time reactor.
AIChE J., 44, 2465–2477, 1998.

[86] P. Deuflhard, E. Hairer, and J. Zugk. One-step and extrapolation methods for diffrenetial-algebraic systems.
Num. Math., 51, 501–516, 1987.

[87] J. Braun, T. Hauber, H. Többen, P. Zacke, D. Chatterjee, O. Deutschmann, and J. Warnatz. Influence of Physical
and Chemical Parameters on the Conversion Rate of a Catalytic Converter: A Numerical Simulation Study. SAE
paper, 2000-01-0211, 2000.

[88] D. Chatterjee. Detaillierte Modellierung von Abgaskatalysatoren. Dissertation. Naturwissenschaftlich-Mathema-


tische Gesamtfakultät. Ruprecht-Karls-Universität Heidelberg, 2001.

[89] D. Chatterjee, O. Deutschmann, and J. Warnatz. Simulation of HC-SCR using detailed models for chemistry and
transport, publication in preparation.

[90] Fluent Version 4.4. Fluent Inc., Lebanon, New Hampshire, 1997.
REFERENCES 151

[91] S. Tischer and O. Deutschmann. Modeling the oxy-dehydrogenation of ethane in annular recators. publication
in preparation.

[92] S. Tischer, C. Correa, and O. Deutschmann. Transient three-dimensional simulation of a catalytic combustion
monolith using detailed models for heterogeneous and homogeneous reactions and transport phenomena. 1st In-
ternatl. Conference on Structured Catalysts and Reactors, October 21-24, 2001, Delft/The Netherlands, accpeted
for publication in Catal. Today.

[93] C. Correa, S. Tischer, and O. Deutschmann. Transient three-dimensional simulation of automotve catalytic
converters. publication in preparation.

[94] A.C. Hindmarsh. Odepack, a systematized collection of ode solvers. In R.S. Stepleman et al., editor, Scientific
Computing, pages 55–64. North-Holland, Amsterdam, 1983.

[95] M. Baerns. Oxidative Coupling of Methane for the Utilization of Natural Gas. In H.I. de Lasa et al., editor,
Chemical Reactor Technology for Environmental Safe Reactors and Products, pages 283–316. Kluwer Academic
Publishers, 1993.

[96] J.A. Labinger. Apporaches to catalytic methane conversion in academic and industrial research: comparison,
competition, collaboration. QUIMICA, 48, 27, 1993.

[97] M. Belgued, H. Amariglio, P. Pareja, A. Ameriglio, and J. Saint-Just. Low Temperature Catalytic Homologation
of Methane on Platinum, Ruthenium and Cobolt. Catal. Today, 13, 437–445, 1992.

[98] F. Solymosi, A. Erdhelyi, and J. Cserenyi. A comparative study on the activation and reactions of methane on
supported metals. Catal. Lett., 16, 399–405, 1992.

[99] V.R. Choudhary, A.M. Rajput, and V.H. Rane. Low-Temperature Catalytic Selective Partial Oxidation of
Methane to CO and H2 over Ni/Yb2 O3 . J. Phys. Chem., 96, 8686, 1992.

[100] S.C. Reyes, E. Igelsia, and C.P. Kelkar. Kinetic-transport models of bimodal reaction sequences-I. Homogeneous
and heterogeneous pathways in oxidative coupling of methane. Chem. Eng. Sci., 48, 2643, 1993.

[101] M. Wolf, O. Deutschmann, F. Behrendt, and J. Warnatz. Simulation of the Oxygen-free Methane Conversion to
Higher Hydrocarbon Fuels Using a Platinum Catalyst with a Catalytically Active Carbonaceous Overlayer. In
Natural Gas Conversion V, Studies in Surface Science and Catalysis 119, pages 271–276. Elsevier, Amsterdam,
1998.

[102] R.W. Borry, E.C. Lu, Y.-H. Kim, and E. Iglesia. Non-oxidative catalytic conversion of methane with continuous
hydrogen removal. In Natural Gas Conversion V, Studies in Surface Science and Catalysis, pages 403–410.
Elsevier, Amsterdam, 1998.

[103] M. Wolf, O. Deutschmann, F. Behrendt, and J. Warnatz. Kinetic model of an oxygen-free methane conversion
on a platinum catalyst. Catal. Lett., 61, 5, 1999.

[104] L. Li, R. Borry, and E. Iglesia. Reaction-transport simulations of non-oxidative methane conversion with con-
tinuous hydrogen removal - homogeneous-heterogeneous reaction pathways. Chem. Eng. Sci., 56, 1869–1881,
2001.

[105] K. Weissermel and H.-J. Arpe. Industrielle organische Chemie. VCH Weinheim, 1988.

[106] M. Prettre, Ch. Eichner, and M. Perrin. Trans. Faraday Soc., 43, 335, 1946.

[107] D. Dissanayake, M.P. Rosynek, K.C. Kharas, and J.H. Lundsford. Partial oxidation of methane to carbon
monoxide and hydrogen over a nickel/alumina catalyst. J. Catal., 132, 117–127, 1991.

[108] A.T. Ashcroft, A.K. Cheetham, J.S. Ford, M.L.H. Green, C.P. Grey, A.J. Murrell, and P.D.F. Vernon. Selective
oxidation of methane to synthesis gas using transition catalysts. Nature, 344, 319, 1990.

[109] P.D.F. Vernon, M.L.H. Green, A.K. Cheetham, and A.T. Ashcroft. Catal. Lett., 6, 181, 1990.

[110] P.D.F. Vernon, M.L.H. Green, A.K. Cheetham, and A.T. Ashcroft. Partial oxidation of methane to synthesis gas
and carbon dioxide as an oxidizing agent for methane conversion. Catal. Today, 13, 417–426, 1992.

[111] W.J.M. Vermeiren, E. Blomsma, and P.A. Jacobs. Catalytic and thermodynamic approach of the oxyreforming
reaction of methane. Catal. Today, 13, 427–436, 1992.
152 REFERENCES

[112] V.R. Choudhary, A.M. Rajput, and B. Prabhakar. Low temperature oxidative conversion of methane to syngas
over nickel oxide - calcium oxide catalyst. Catal. Lett., 15, 363–370, 1992.

[113] V.R. Choudhary, A.M. Rajput, and B. Prabhakar. Nonequilibrium oxidative conversion of methane to carbon
monoxide and hydrogen with high selectivity and productivity over nickel/alumina at low temperatures. J. Catal.,
139, 326–328, 1993.

[114] V.R. Choudhary, A.M. Rajput, and B. Prabhakar. Low temperature oxidative conversion of methane to syngas
over cobalt/rare earth oxide catalysts. Catal. Lett., 16, 269–272, 1992.

[115] V.R. Choudhary, A.M. Rajput, and B. Prabhakar. Low temperature selective conversion of methane to carbon
monoxide and hydrogen over cobalt-magnesium oxide catalysts. Appl. Catal., 90, L1–L5, 1992.

[116] J.A. Lapszewicz and X.-Z. Jiang. Investigation of the mechanism of partial oxidation of methane to synthesis
gas. Prep. Am. Chem. Soc. Div. Pet. Chem., 37, 252–260, 1992.

[117] Y. Matsumura and J.B. Moffat. Partial Oxidation of Methane to Carbon Monoxide and Hydrogen with Molecular
Oxygen and Nitrous Oxide over Hydroxyapatite Catalysts. Catal. Lett., 148, 323, 1994.

[118] O.V. Buyevskaya, D. Wolf, and M. Baerns. Rhodium - catalyzed partial oxidation of methane to CO and h2 .
transient studies on its mechanism. Catal. Lett., 29, 249–260, 1994.

[119] K. Walter, O.V. Buyevskaya, D. Wolf, and M. Baerns. Rhodium-catalyzed partial oxidation of methane to CO
and H2 . In situ DRIFTS studies on surface intermediates. Catal. Lett., 29, 261–270, 1994.

[120] D. Wolf, M. Höhenberger, and M. Baerns. External Mass and Heat Transfer Limitations of the Partial Oxidation
of Methane over Pt/MgO - Consequences for Adiabatic Reactor Operation. Ind. Eng. Chem. Res., 36, 3345–3353,
1997.

[121] D. Wolf, M. Barre-Chassonnery, M. Höhenberger, A. van Veen, and M. Baerns. Kinetics of the Water-Gas shift
reaction and its Role in the Conversion of Methane to Syngas over a Pt/MgO Catalyst. Catal. Today, 40, 147–156,
1998.

[122] D.A. Hickman and L.D. Schmidt. Steps in CH4 Oxidation on Pt and Rh Surfaces: High Temperature Reactor
Simulation. AIChE J., 39 (7), 1164–1177, 1993.

[123] K. Heitnes, S. Lindeberg, O.A. Rokstad, and A. Holmen. Catalytic partial oxidation of methane to synthesis gas
using monolithic reactors. Catalysis Today, 21, 471–480, 1994.

[124] K. Heitnes, S. Lindeberg, O.A. Rokstad, and A. Holmen. Catalytic partial oxidation of methane to synthesis gas.
Catalysis Today, 24, 211–216, 1995.

[125] J.C. Jalibert, M. Fathi, O.A. Rokstad, and A. Holmen. Synthesis gas production by partial oxidation of metahne
from the cyclic gas-solid reaction using promoted cerium oxide. In Natural Gas Conversion VI, Studies in Surface
Science and Catalysis 136 (E. Iglesia, J.J. Spivey, T.H. Fleisch (eds.), pages 301–306. Elsevier, Amsterdam, 2001.

[126] E.P.J. Mallens, J.H.B. Hoebink, and G.B. Marin. The reaction mechanism of the partial oxidation of methane to
synthesis gas: A transient kinetic study over rhodium and a comparison with platinum. J. Catal., 167, 43–46,
1997.

[127] M. Schwiedernoch. Untersuchung zur katalytischen Partialoxidation von Methan. Diplomarbeit, Fakultät Chemie
der Ruprecht-Karls-Universität Heidelberg, 2000.

[128] O. Deutschmann, R. Schwiedernoch, L.I. Maier, and D. Chatterjee. Natural Gas Conversion in Monolithic
Catalysts: Interaction of Chemical Reactions and Transport Phenomena. In Natural Gas Conversion VI, Studies
in Surface Science and Catalysis 136 (E. Iglesia, J.J. Spivey, T.H. Fleisch (eds.), pages 215–258. Elsevier,
Amsterdam, 2001.

[129] L.D. Schmidt. University of Minnesota, Minneapolis, MN, USA, personal communication, 1997.

[130] A.S. Bodke, S.S. Bharadwaj, and L.D. Schmidt. The Effect of Ceramic Supports on Partial Oxidation of Hydro-
carbons over Noble Metal Coated Monoliths. J. Catal., 179, 138–149, 1998.

[131] J. Frauhammer, L. von Hippel, D. Arntz, and G. Eigenberger. A new reactor concept for endothermic high-
temperature reactions. Chem. Eng. Sci., 54, 3661–3670, 1999.

[132] G. Kolios, J. Frauhammer, and G. Eigenberger. Autothermal fixed-bed reactor concepts. Chem. Eng. Sci., 55,
5945–5967, 2000.
REFERENCES 153

[133] G. Veser, J. Frauhammer, and U. Friedle. Syngas formation by direct oxidation of methane - Reaction mechanisms
and new reactor concepts. Catal. Today, 61, 55–64, 2000.

[134] G. Kolios, J. Frauhammer, and G. Eigenberger. A simplified procedure for the optimal design of autothermal
reactors for endothermic high-temperature reactions. Chem. Eng. Sci., 56, 351–357, 2001.

[135] L. Basini, A. Guarinoni, K. Aasberg-Petersen, and J.H. Bak-Hansen. Catalytic partial oxidation of natural gas
at elevated pressure and low residence time. Proc. EuropCat 4, Rimini/Italy, September, 1999.

[136] L.D. Schmidt, O. Deutschmann, and C.T. Goralski, Jr. Modeling the Partial Oxidation of Methane to Syngas
at Millisecond Contact Times. In Natural Gas Conversion V, Studies in Surface Science and Catalysis, pages
685–692. Elsevier, Amsterdam, 1998.

[137] C.R.H. de Smet, M.H.J.M. de Croon, R.J. Berger, G.B. Marin, and J. C. Schouten. An experimental reactor to
study the intrinsic kinetcis of the partial oxidation of methane to synthesis gas in the presence of heat-transport
limitations. Appl. Catal. A, 187, 33, 1999.

[138] G. Veser and J. Frauhammer. Modelling steady state and ignition during catalytic methane oxidation in a
monolith reactor. Chem. Eng. Sci., 55, 2271–2286, 2000.

[139] D.A. Hickman and L.D. Schmidt. Synthesis Gas Formation by Direct Oxidation of Methane over Pt Monoliths.
J. Catal., 138, 267–282, 1992.

[140] O. Deutschmann and L.D. Schmidt. Two-dimensional modeling of partial oxidation of methane on rhodium in a
short contact time reactor. Proc. Combust. Inst., 27, 2283–2291, 1998.

[141] O. Deutschmann, L.D. Schmidt, and J. Warnatz. Simulation of reactive flow in a partial oxidation reactor with
detailed gas phase and surface chemistry models. In F. Keil, W. Mackens, H. Vos̈, and J. Werther, editors,
Scientific Computing in Chemical Engineering II. Computational Fluid Dynamics, Reaction Engineering, and
Molecular Properties, pages 368–375. Springer, 1999.

[142] S. Tummala, L.D. Schmidt, and O. Deutschmann. publication in preparation.

[143] V. Karbach. Validierung eines detaillierten Reaktionsmechanismus zur Oxidation von Kohlenwasserstoffen bei
hohen Temperaturen. Diplomarbeit, Fakultät Chemie der Ruprecht-Karls-Universität Heidelberg, 1997.

[144] K.L. Hohn, P.M. Witt, M.B. Davis, and L.D. Schmidt. Methane Coupling to Acetylene over Pt Coated Monoliths
at Millisecond Contact Times. Catal. Lett., 54, 113–118, 1998.

[145] D. Papadias, L. Edsberg, and P. Björnbom. Simplified Method of Effectiveness Factor Calculations for Irregular
Geometries of Washcoats; A General Case in a 3D Concentration Field. Catal. Today, 60, 11, 2000.

[146] M. Huff and L. D. Schmidt. Ethylene Formation by Oxidative Dehydrogenation of Ethane over Monoliths at Very
Short Contact Times. J. Phys. Chem., 97, 11815–11822, 1993.

[147] A.G. Dietz III and L.D. Schmidt. Effect of pressure on three catalytic partial oxidation reactions at millisecond
contact times. Catal. Lett., 33, 5–29, 1995.

[148] C. Yokoyama, S.S. Bharadwaj, and L.D. Schmidt. Platinum-tin and platinum-copper catalysts for autothermal
oxidative dehydrogenation of ethane to ethylene. Catal. Lett., 38, 181–188, 1996.

[149] P. Witt and L.D. Schmidt. Effectof flow rate on the partial oxidation of methane and ethane. J. Catal., 163,
465–475, 1996.

[150] D.W. Flick and M.C. Huff. Acetylene formation during the catalytic oxidative dehydrogenation of ethane over a
Pt-coated monoliths at short contact times. Catal. Lett., 47, 91–97, 1997.

[151] A.S. Bodke, D. Henning, L.D. Schmidt, S.S. Bharadwaj, and J. Siddall. Oxidative Dehydrogenation of Ethane
at Millisecond Contact Times: Effect of Hydrogen Addition. J. Catal., 191, 62–74, 2000.

[152] D.W. Flick and M.C. Huff. Oxidative dehydrogenation of ethane over a Pt-coated monolith versus Pt-loaded
pellets: Surface area and thermal effects. J. Catal., 178, 315–327, 1998.

[153] D.W. Flick and M.C. Huff. Oxidative dehydrogenation of ethane over supported chromium oxide and Pt modified
chromium oxide. Appl. Catal. A, 187, 13–24, 1999.

[154] E. Morales and J.H. Lundsford. Oxidative dehydrogenation of ethane over a lithium-promoted magnesium oxide
catalyst. J. Catal., 118, 255–265, 1989.
154 REFERENCES

[155] G.A. Martin, A. Bates, V. DuCarme, and C. Mirodatos. Oxidative conversion of methane and C2 hydrocarbons
on oxides: homogeneous versus heterogeneous processes. Appl. Catal., 47, 287–297, 1989.

[156] R. Burch and E.M. Crabb. Homogeneous and heterogeneous contributions to the catalytic oxidative dehydro-
genation of ethane. Appl. Catal. A, 97, 49–65, 1999.

[157] M. Baerns and O. Buyevskay. Simple chemical processes based on low molecular-mass alkanes as chemical
feedstocks. Catal. Today, 45, 13–22, 1998.

[158] R. Lødeng, O.A. Lindvåg, S. Kvisle, H. Reier-Nielsen, and A. Holmen. Oxidative dehydrogenation of ethane over
Pt and Pt/Rh gauze catalyst at very short contact times. In Natural Gas Conversion V, Studies in Surface
Science and Catalysis 119, pages 641–646. Elsevier, Amsterdam, 1998.

[159] A. Beretta, G. Groppi, L. Majocchi, and P. Forzatti. Potentialities and draw-backs of the experimental approach
to the study of high T and high GHSV kinetics. Appl. Catal. A, 187, 49–60, 1999.

[160] A. Beretta, P. Forzatti, and E. Ranzi. Production of olefins via oxidative dehydrogenation of propane in autother-
mal conditions. J. Catal., 184, 469–478, 1999.

[161] M. Huff and L.D. Schmidt. Elementary Step Model of Ethane Oxidative Dehydrogenation on Pt-Coated Mono-
liths. AIChE J., 42, 3484–3497, 1996.

[162] M.C. Huff, I.P. Androulakis, J.H. Sinfelt, and S.C. Reyes. The Contribution of Gas-Phase Reactions in the
Pt-Catalyzed Conversion of Ethane-Oxygen Mixtures. J. Catal, 191, 46–45, 2000.

[163] D.K. Zerkle, M.D. Allendorf, M. Wolf, and O. Deutschmann. Modeling of On-Line Catalyst Addition Effects in
a Short Contact Time Reactor. Proc. Combust. Inst., 28, 1365–1372, 2000.

[164] J. Siddall. Dow Chemical Company, personal communication, 2000.

[165] N.M. Marinov, W.J. Pitz, C.K. Westbrook, M.J. Castaldi, and S.M. Senkan. Modeling of aromatic and polycyclic
aromatic hydrocarbon formation in premixed methane and ethane flames. Combust. Sci. Technol., 116, 211–287,
1996.

[166] M.D. Allendorf and D.K. Zerkle, 2001.

[167] G. Stahl and J. Warnatz. Numerical investigation of time-dependent properties and extinction of strained
methane- and propane-air flamelets. Combust. Flame, 85, 285–299, 1991.

[168] A.E. Lutz, R.J. Kee, and J.A. Miller. A Program for Predicting Homogeneous Gas-phase Chemical Kinetics in
a Closed System with Sensitivity Analysis. Sandia National Laboratories Report No. SAND87-8248, 1987.

[169] C.K. Law F.N. Egolfopoulos, D.L. Zhu. Experimental and numerical determination of laminar flame speeds:
Mixtures of C2 - hydrocarbons with oxygen and nitrogen. Proc. Combust. Inst., 23, 471–478, 1990.

[170] R.J. Kee, J.F. Grcar, M.D. Smooke, and J.A. Miller. A FORTRAN Program for Modeling Steady Laminar
One-Dimensional Flames. Sandia National Laboratories Report No. SAND85-8240, 1985.

[171] P. Dagaut, M. Cathonnet, and J.C. Boettner. Kinetics of ethane oxidation. Int. J. Chem. Kin., 23, 437–455,
1989.

[172] M. Rinnemo, O. Deutschmann, F. Behrendt, and B. Kasemo. Experimental and Numerical Investigation of the
Catalytic Ignition of Mixtures of Hydrogen and Oxygen on Platinum. Combust. Flame, 111, 312–326, 1997.

[173] J. Warnatz, M.D. Allendorf, R.J. Kee, and M.E. Coltrin. A Model of Elementary Chemistry and Fluid Mechanics
in the Combustion of Hydrogen on Platinum Surfaces. Combust. Flame, 96, 393–406, 1994.

[174] G.B. Fisher and J.L. Gland. The Interaction of Water with the Pt(111) Surface. Surf. Sci., 94, 446–455, 1980.

[175] A.B. Anton and D.C. Cadogan. Surf. Sci. Lett., 239, L548, 1990.

[176] D. Knight. Humphry Davy: Science & Power. Blackwell Publishers, Cambridge, MA, 1992.

[177] G.B. Kauffman. Johann Wolfgang Döbereiners Feuerzeug. On the sesquicentennial Anniversary of his Death.
Platinum Metals Rev., 43, 122–128, 1999.

[178] R.A. Dalla Betta. Catalytic combustion gas turbine systems: the preferred technology for low emissions electric
power generation and co-generation. Catal. Today, 35, 129–135, 1997.
REFERENCES 155

[179] P. Forzatti and G. Groppi. Catalytic combustion for the production of energy. Catal. Today, 54, 165–180, 1999.

[180] K.-I. Fujimoto, F.H. Ribeiro, M. Avalos-Borja, and E. Iglesia. Structure and Reactivity of PdOx /ZrO2 Catalysts
for Methane Oxidation at Low Temperatures. J. Catal., 179, 431–442, 1998.

[181] G. Groppi, C. Cristiani, and P. Forzatti. Preparation and characterization of hexaaluminate materials for high -
temperature catalytic combustion. Catal., 13, 85–113, 1997.

[182] G. Saraccoa and V. Specchia. Catalytic filters for the abatement of volatile organic compounds. Chem. Eng. Sci.,
55, 897–908, 2000.

[183] L.D. Pfefferle and W.C. Pfefferle. Catalysis in Combustion. Catal. Rev.-Sci. Eng., 29, 219–267, 1987.

[184] K.E. Brenan, S.L. Campbell, and L.R. Petzold. Numerical Solution of Initial-Value Problems in Differential-
Algebraic Equations. 2nd ed., SIAM, Philadelphia, 1998.

[185] O. Deutschmann, F. Behrendt, and J. Warnatz. Modelling and Simulation of Heterogeneous Oxidation of Methane
on a Platinum Foil. Catal. Today, 21, 461–470, 1994.

[186] W.M. Kays and M.E. Crawford. Convective Heat and Mass Transfer. McGraw-Hill, New York, 1980.

[187] R.E. Hayes and S.T. Kolaczkowski. A study of Nusselt and Sherwood numbers in a monolith reactor. Catalysis
Today, 47, 295–303, 1999.

[188] M. Frenklach. GRI-Mech – An Optimized Detailed Chemical Reaction Mechanism for Methane Combustion. GRI
Report No. GRI-95/0058, 1995.

[189] G. Veser and L.D. Schmidt. Ignition and Extinction in the Catalytic Oxidation of Hydrocarbons over Platinum.
AIChE J., 42, 1077–1087, 1996.

[190] G. Veser, J. Frauhammer, L.D. Schmidt, and G. Eigenberger. Catalytic ignition during methane oxidation on
platinum: Experiments and modeling. In Dynamics of Surfaces and Reaction Kinetics in Hetergeneous Catalysis,
Studies in Surface Science and Catalysis 109, pages 273–284. Elsevier, Amsterdam, 1997.

[191] O. Deutschmann, L. Maier, U. Riedel, J. Warnatz, A.H. Stroemanna, and R.W. Dibble. Hydrogen asssisted
catalytic combustion of methane on platinum. Catal. Today, 59, 141–140, 2000.

[192] D.L. Baulch, C.J. Cobos, R.A. Cox, C. Esser, P. Frank, Th. Just, J.A. Kerr, M.J. Pilling, J. Troe, R.W. Walker,
and J. Warnatz. Evaluated Kinetic Data for Combustion Modelling. J. Phys. Chem. Ref. Data, 21, 11, 1992.

[193] C.T. Goralski and L.D. Schmidt. Lean catalytic combustion of alkanes at short contact times. Catal. Lett., 42,
47–55, 1996.

[194] J. Redenius, L.D. Schmidt, and O. Deutschmann. Millisecond Catalytic Wall Reactors: I. Radiant Burner.

[195] R.M. Heck and R.J Farrauto. Catalytic Air Pollution Control. Van Nostrand Reinhold, New York, 1995.

[196] The European Commission. http://europa.eu.int/comm/transport, 02.05.2001.

[197] G.C. Koltsakis, P.A. Konstantinidis, and A.M. Stamatelos. Development and application range of mathematical
models for 3-way catalytic converters. Appl. Catal. B: Environmental, 12, 161–191, 1997.

[198] T. Kirchner and G. Eigenberger. On the dynamic behavior of automotive catalysts. Catal. Today, 28, 3, 1997.

[199] G.P. Ansell, P.S. Bennett, J.P. Cox, J.C. Frost, P.G. Gray, A.-M. Jones, R.R. Rajaram, A.P. Walker, M. Litorell,
and G. Smedler. The development of a model capable of predicting diesel lean NOx catalyst performance under
transient conditions. Appl. Catal. B, 10, 183, 1996.

[200] A.L. Boehman. Numerical Modeling of NO Reduction over Cu-ZSM-5 under Lean Conditions. SAE paper,
970752, 1997.

[201] C.N. Montreuil, S.C. Williams, and A.A. Adamczyk. Modeling current generation catalytic converters; laboratory,
experiments and kinetic parameter optimization-steady state kinetics. SAE-paper, 980881, 1992.

[202] S.-J. Jeong and W.-S. Kim. A numerical approach to investigate transient thermal and conversion characteristics
of automotive catalytic converter. SAE-paper, 980881, 1998.
156 REFERENCES

[203] D. Chatterjee, O. Deutschmann, and J. Warnatz. Detailed surface reaction mechanism in a 3-way catalyst.
Faraday Discussions, accepted for publication.

[204] E. Rogermond, N. Essayem, R. Frety, V. Perrichon, M. Primet, M. Chevrier, C. Gouthier, and F. Mathis.
Characterization of model three-way catalysts. J. Catal., 186, 414, 1999.

[205] S.H. Oh, G.B. Fisher, J.E. Carpenter, and D.W. Goodman. Comparative Kinetic Study of CO-O2 and CO-NO
Reactions over Single Crystal and Supported Rhodium Catalysts. J. Catal., 100, 360, 1986.

[206] E.I. Altman and R.J. Gorte. A temperature-Programmed Desoprtion study of NO on Rh Particles Supported on
α-Al2 O3 (0001). J. Catal., 113, 185, 1988.

[207] A. Fritz and V. Pichon. The current state of research on automotive lean NOx catalysis. Appl. Catal., 13, 1,
1997.

[208] R.J. Farrauto and E.M. Heck. Catalytic converter: State of the art and perspectives. Catal. Today, 51, 351,
1999.

[209] Competenznetzwerk Katalyse. http://www.connecat.de/, 02.05.2001.

[210] V. Schulz, O. Deutschmann, L.D. Schmidt, and J. Warnatz. Process optimization of reactive systems modeled by
elementary reactions. In F. Keil, W. Mackens, H. Vos̈, and J. Werther, editors, Scientific Computing in Chemical
Engineering II. Simulation, Image Processing, Optimization, and Control, pages 354–361. Springer, 1999.

[211] M.v. Schwerin, O. Deutschmann, , and V. Schulz. Process optimization of reactive systems by partiallly reduced
sqp methods. Computer & Chemical Engineering, 24, 89–97, 2000.

[212] V. Schulz. Reduced SQP methods for large-scale optimal control problems in DAE with application to path plan-
ning problems for satellite mounted robots. Dissertation. Naturwissenschaftlich-Mathematische Gesamtfakultät.
Ruprecht-Karls-Universität Heidelberg, 1996.

[213] S. Großhans, O. Deutschmann, V. Schulz, and H.G. Bock. Optimal control of a reactive stagnation point flow on
a catalytic surface. publication in preparation.
157

A Gas-phase reaction mechanisms


In this section, the homogeneous reaction mechanisms are listed which are applied in
the Chpaters 5 and 6. Because gas phase chemistry is not significant in the applications
of Chapter 7, the mechanisms used are not listed here; it is referred to the literature
as given in the text.
In the schemes, the symbol denotes that the kinetic data of the reverse reaction
are calculated using Eq. 3.10. The kinetic data are given according to Eq. 3.15. The
units are: A [mol, cm, s], β [-], and Ea [kJ mol−1 ]. The lines beginning with the keywords
Low and Troe below the reaction equation refer to Troe parameters as described in
References [37, 81] for the given reaction; the data behind the reaction equation refer
to k∞ , behind Low to k0 , and behind Troe to the Troe parameters a, T ∗∗∗ , T ∗ , and T ∗∗ .
Efficiency coefficients (ηik in Eq. 3.17) for third body reactions, i.e. M(*) is part of the
reaction equation, are unity unless specified differently at the end of the table.

A.1 Partial oxidation of methane


This mechanism was developed for the numerical simulation of the homogeneous gas-
phase reactions in partial oxidation of methane as described in Chapter 5, where the
origin of the mechanism is discussed.

Reaction A β Ea
1. O2 + H
OH + O 2.000 · 1014 0.0 70.3
2. H2 + O
OH + H 5.060 · 104 2.7 26.3
3. H2 + OH
H2 O + H 1.000 · 108 1.6 13.8
4. OH + OH
H2 O + O 1.500 · 109 1.1 0.4
5. H + O2 + M(1)
HO2 + M(1) 2.300 · 1018 −0.8 0.0
6. HO2 + H
OH + OH 1.500 · 1014 0.0 4.2
7. HO2 + H
H2 + O2 2.500 · 1013 0.0 2.9
8. HO2 + H
H2 O + O 3.000 · 1013 0.0 7.2
9. HO2 + O
OH + O2 1.800 · 1013 0.0 −1.7
10. HO2 + OH
H2 O + O2 6.000 · 1013 0.0 0.0
11. HO2 + HO2
H2 O2 + O2 2.500 · 1011 0.0 −5.2
12. OH + OH + M(1)
H2 O2 + M(1) 3.250 · 1022 −2.0 0.0
13. H 2 O2 + H
H2 + HO2 1.700 · 1012 0.0 15.7
14. H2 O2 + OH
H2 O + HO2 5.400 · 1012 0.0 4.2
15. CO + OH
CO2 + H 6.000 · 106 1.5 −3.1
16. CO + HO2
CO2 + OH 1.500 · 1014 0.0 98.7
17. CO + O + M(1)
CO2 + M(1) 7.100 · 1013 0.0 −19.0
18. CO + O2
CO2 + O 2.500 · 1012 0.0 200.0
19. CHO + M(1)
CO + H + M(1) 7.100 · 1014 0.0 70.3
20. CHO + O2
CO + HO2 3.000 · 1012 0.0 0.0
21. CH2 O + M(1)
CHO + H + M(1) 5.000 · 1016 0.0 320.0
22. CH2 O + H
CHO + H2 2.300 · 1010 1.1 13.7
23. CH2 O + O
CHO + OH 4.150 · 1011 0.6 11.6
24. CH2 O + OH
CHO + H2 O 3.400 · 109 1.2 −1.9
25. CH2 O + HO2
CHO + H2 O2 3.000 · 1012 0.0 54.7
26. CH2 O + CH3
CHO + CH4 1.000 · 1011 0.0 25.5
27. CH2 O + O2
CHO + HO2 6.000 · 1013 0.0 170.7
158 A. GAS-PHASE REACTION MECHANISMS

Reaction A β Ea
28. CH3 + O
CH2 O + H 8.430 · 1013 0.0 0.0
29. CH4 + M(2)
CH3 + H + M(2) 2.400 · 1016 0.0 439.0
Low: 1.290 · 1018 0.0 380.0
Troe: 0.000 1.350 · 1003 1.0 7830.0
30. CH3 + OH → CH3 O + H 2.260 · 1014 0.0 64.8
31. CH3 O + H → CH3 + OH 4.750 · 1016 −0.1 88.0
32. CH3 + O2 → CH2 O + OH 3.300 · 1011 0.0 37.4
33. CH3 + HO2
CH3 O + OH 1.800 · 1013 0.0 0.0
34. CH3 + HO2
CH4 + O2 3.600 · 1012 0.0 0.0
35. CH3 + CH3 → C2 H4 + H2 1.000 · 1016 0.0 134.0
36. CH3 + CH3 + M(1)
C2 H6 + M(1) 3.610 · 1013 0.0 0.0
Low: 3.630 · 1041 −7.0 11.6
Troe: 0.620 7.300 · 1001 1180.0 0.0
37. CH3 O + M(1)
CH2 O + H + M(1) 5.000 · 1013 0.0 105.0
38. CH3 O + H
CH2 O + H2 1.800 · 1013 0.0 0.0
39. CH3 O + O2
CH2 O + HO2 4.000 · 1010 0.0 8.9
40. CH3 O + O
O2 + CH3 1.100 · 1013 0.0 0.0
41. CH4 + H
H2 + CH3 1.300 · 104 3.0 33.6
42. CH4 + O
OH + CH3 6.923 · 108 1.6 35.5
43. CH4 + OH
H2 O + CH3 1.600 · 107 1.8 11.6
44. CH4 + HO2
H2 O2 + CH3 1.100 · 1013 0.0 103.1
45. C2 H3 + M(1)
C2 H2 + H + M(1) 1.900 · 1014 0.0 166.3
Low: 1.000 · 1042 −7.5 190.4
Troe: 0.350 0.000 · 1000 0.0 0.0
46. C2 H3 + OH
C2 H2 + H2 O 5.000 · 1013 0.0 0.0
47. C2 H3 + H
C2 H2 + H2 1.200 · 1013 0.0 0.0
48. C2 H3 + O2
CH2 O + CHO 2.300 · 1012 0.0 −2.3
49. C2 H4 + M(1)
C2 H2 + H2 + M(1) 7.500 · 1017 0.0 332.0
50. C2 H4 + M(1)
C2 H3 + H + M(1) 0.850 · 1018 0.0 404.0
51. C2 H4 + H
C2 H3 + H2 0.540 · 1015 0.0 62.9
52. C2 H4 + O
CHO + CH3 2.420 · 106 2.1 0.0
53. C2 H4 + OH
C2 H3 + H2 O 2.200 · 1013 0.0 24.9
54. C2 H4 + H + M(1) → C2 H5 + M(1) 3.970 · 109 1.3 5.4
Low: 6.980 · 1018 0.0 3.2
Troe: 0.760 4.000 · 1001 1025.0 0.0
55. C2 H5 + M(1) → C2 H4 + H + M(1) 8.200 · 1013 0.0 166.8
Low: 3.400 · 1017 0.0 139.6
Troe: 0.750 9.700 · 1001 1379.0 0.0
56. C2 H5 + H
CH3 + CH3 3.000 · 1013 0.0 0.0
57. C2 H5 + O2
C2 H4 + HO2 1.100 · 1010 0.0 −6.3
58. C2 H5 + CH3
C2 H4 + CH4 1.140 · 1012 0.0 0.0
59. C2 H5 + C2 H5
C2 H4 + C2 H6 1.400 · 1012 0.0 0.0
60. C2 H6 + H
C2 H5 + H2 1.400 · 109 1.5 31.1
61. C2 H6 + O
C2 H5 + OH 1.000 · 109 1.5 24.4
62. C2 H6 + OH
C2 H5 + H2 O 7.200 · 106 2.0 3.6
63. C2 H6 + HO2
C2 H5 + H2 O2 1.700 · 1013 0.0 85.9
64. C2 H6 + O2
C2 H5 + HO2 6.000 · 1013 0.0 217.0
65. C2 H6 + CH3
C2 H5 + CH4 1.500 · 10−7 6.0 25.4

H2 H2 O O2 N2 CO CO2 CH4
M(1) 1.00 6.50 0.40 0.40 0.75 1.50 3.00
M(2) 1.00 6.50 0.40 0.40 0.75 1.50 0.66
A.2 Oxy-dehydrogenation of ethane 159

A.2 Oxy-dehydrogenation of ethane


This mechanism was developed for the numerical simulation of the homogeneous gas-
phase recations in oxy-dehydrogenation as described in Chapter 6, where the origin of
the mechanism [49, 166] is discussed.

Reaction A β Ea
1. OH + H2
H + H2 O 2.140 · 108 1.5 14.4
2. O + OH
O2 + H 2.020 · 1014 −0.4 0.0
3. O + H2
OH + H 5.060 · 104 2.7 26.3
4. H + O2 + M(1)
HO2 + M(1) 4.520 · 1013 0.0 0.0
Low: 1.050 · 1019 −1.2 0.0
5. H + O 2 + H2
HO2 + H2 4.520 · 1013 0.0 0.0
Low: 1.520 · 1019 −1.13 0.0
6. H + O 2 + H2 O
HO2 + H2 O 4.520 · 1013 0.0 0.0
Low: 2.100 · 10223 −2.4 0.0
7. OH + HO2
H2 O + O2 2.130 · 1028 −4.8 14.6
8. OH + HO2
H2 O + O2 9.100 · 1014 0.0 45.9
9. H + HO2
OH + OH 1.500 · 1014 0.0 4.2
10. H + HO2
H2 + O2 8.450 · 1011 0.7 5.2
11. H + HO2
O + H2 O 3.010 · 1013 0.0 7.2
12. O + HO2
O2 + OH 3.250 · 1013 0.0 0.0
13. OH + OH
O + H2 O 3.570 · 104 2.4 −8.8
14. H + H + M(2)
H2 + M(2) 1.000 · 1018 −1.0 0.0
15. H + H + H2
H2 + H2 9.200 · 1016 −0.6 0.0
16. H + H + H2 O
H2 + H2 O 6.000 · 1019 −1.2 0.0
17. H + OH + M(3)
H2 O + M(3) 2.210 · 1022 −2.0 0.0
18. H + O + M(3)
OH + M(3) 4.710 · 1018 −1.0 0.0
19. O+O+M
O2 + M 1.890 · 1013 0.0 −7.5
20. HO2 + HO2
H2 O2 + O2 4.200 · 1014 0.0 50.1
21. HO2 + HO2
H2 O2 + O2 1.300 · 1011 0.0 −6.8
22. OH + OH + M
H2 O2 + M 1.240 · 1014 −0.4 0.0
Low: 3.040 · 1030 −4.6 8.4
Troe: 0.470 1.000 · 1002 2.0 · 1003 1.0 · 1015
23. H2 O2 + H
HO2 + H2 1.980 · 106 2.0 10.2
24. H2 O2 + H
OH + H2 O 3.070 · 1013 0.0 17.6
25. H2 O2 + O
OH + HO2 9.550 · 106 2.0 16.6
26. H2 O2 + OH
H2 O + HO2 2.400 · 100 4.0 −9.0
27. CH3 + CH3 + M(4)
C2 H6 + M(4) 9.220 · 1016 −1.2 2.7
Low: 1.140 · 1036 −5.3 7.1
Troe: 0.405 1.120 · 1003 69.6 1.0 · 1015
28. CH3 + H + M(4)
CH4 + M(4) 2.140 · 1015 −0.4 0.0
Low: 3.31 · 1030 −4.0 8.8
Troe: 0.000 1.0 · 10−15 1.0 · 10−15 40.0
29. CH4 + H
CH3 + H2 2.200 · 104 3.0 36.6
30. CH4 + OH
CH3 + H2 O 4.190 · 106 2.0 10.7
31. CH4 + O
CH3 + OH 6.920 · 108 1.6 35.5
32. CH4 + HO2
CH3 + H2 O2 1.120 · 1013 0.0 103.1
33. CH3 + HO2
CH3 O + OH 7.000 · 1012 0.0 0.0
34. CH3 + HO2
CH4 + O2 3.000 · 1012 0.0 0.0
35. CH3 + O
CH2 O + H 8.000 · 1013 0.0 0.0
36. CH3 + O2
CH3 O + O 1.450 · 1013 0.0 122.2
37. CH3 + O2
CH2 O + OH 2.510 · 1011 0.0 61.3
38. CH3 O + H
CH3 + OH 1.000 · 1014 0.0 0.0
39. CH3 + OH
3-CH2 + H2 O 3.000 · 106 2.0 10.5
40. CH3 + OH
CH2 O + H2 5.480 · 1013 0.0 12.5
41. CH3 + OH
CH2 O + H2 2.250 · 1013 0.0 18.0
42. CH3 + H
3-CH2 + H2 9.000 · 1013 0.0 63.2
43. CH3 + M
3-CH2 + H + M 1.960 · 1016 0.0 382.5
160 A. GAS-PHASE REACTION MECHANISMS

Reaction A β Ea
44. CH2 O + H + M(5)
CH3 O + M(5) 5.400 · 1011 0.5 10.9
Low: 1.50 · 1030 −4.8 23.0
Troe: 0.758 9.400 · 1001 1555 4200
45. CH3 O + H
CH2 O + H2 2.000 · 1013 0.0 0.0
46. CH3 O + OH
CH2 O + H2 O 1.000 · 1013 0.0 0.0
47. CH3 O + O
CH2 O + OH 1.000 · 1013 0.0 0.0
48. CH3 O + O2
CH2 O + HO2 6.300 · 1010 0.0 10.9
49. CH2 O + OH
CHO + H2 O 2.000 · 1013 0.0 0.0
50. CH2 O + H
CH2 O + H 2.000 · 1014 0.0 0.0
51. CH2 O + O
CO2 + H + H 5.000 · 1013 0.0 0.0
52. CH2 O + O
CO + OH + H 3.000 · 1013 0.0 0.0
53. CH2 O + O2
CO2 + H + OH 5.000 · 1012 0.0 0.0
54. CH2 O + O2
CO2 + H2 O 3.000 · 1013 0.0 0.0
55. 3-CH2 + OH
CH2 O + H 2.500 · 1013 0.0 0.0
56. 3-CH2 + CO2
CH2 O + CO 1.100 · 1011 0.0 4.2
57. 3-CH2 + O
CO + H + H 5.000 · 1013 0.0 0.0
58. 3-CH2 + O
CO + H2 3.000 · 1013 0.0 0.0
59. 3-CH2 + O2
CH2 O + O 3.290 · 1021 −3.3 12.0
60. 3-CH2 + O2
CO2 + H + H 3.290 · 1021 −3.3 12.0
61. 3-CH2 + O2
CO2 + H2 1.010 · 1021 −3.3 6.3
62. 3-CH2 + O2
CO + H2 O 7.280 · 1019 −2.5 7.6
63. 3-CH2 + O2
CHO + OH 1.290 · 1020 −3.3 1.2
64. 3-CH2 + CH3
C2 H4 + H 4.000 · 1013 0.0 0.0
65. 3-CH2 + 3-CH2
C2 H2 + H + H 4.000 · 1013 0.0 0.0
66. 3-CH2 + HCCO
C2 H3 + CO 3.000 · 1013 0.0 0.0
67. CH2 O + OH
CHO + H2 O 3.430 · 109 1.2 −1.9
68. CH2 O + H
CHO + H2 2.190 · 108 1.8 12.6
69. CH2 O + M
CHO + H + M 3.310 · 1016 0.0 338.9
70. CH2 O + O
CHO + OH 1.800 · 1013 0.0 12.9
71. CHO + O2
HO2 + CO 7.580 · 1012 0.0 1.7
72. CHO + M(6)
H + CO + M(6) 1.860 · 1017 −1.0 71.1
73. CHO + OH
H2 O + CO 1.000 · 1014 0.0 0.0
74. CHO + H
CO + H2 1.190 · 1013 0.2 0.0
75. CHO + O
CO + OH 3.000 · 1013 0.0 0.0
76. CHO + O
CO2 + H 3.000 · 1013 0.0 0.0
77. CO + OH
CO2 + H 9.420 · 103 2.2 −9.8
78. CO + O + M
CO2 + M 6.170 · 1014 0.0 12.6
79. CO + O2
CO2 + O 2.530 · 1012 0.0 199.5
80. CO + HO2
CO2 + OH 5.800 · 1013 0.0 96.0
81. C2 H6 + CH3
C2 H5 + CH4 5.500 · 10−1 4.0 34.7
82. C2 H6 + H
C2 H5 + H2 5.400 · 102 3.5 21.8
83. C2 H6 + O
C2 H5 + OH 3.000 · 107 2.0 21.4
84. C2 H6 + OH
C2 H5 + H2 O 7.230 · 106 2.0 3.6
85. C2 H5 + H
C2 H4 + H2 1.250 · 1014 0.0 33.5
86. C2 H5 + H
CH3 + CH3 3.000 · 1013 0.0 0.0
87. C2 H5 + H
C2 H6 7.000 · 1013 0.0 0.0
88. C2 H5 + OH
C2 H4 + H2 O 4.000 · 1013 0.0 0.0
89. C2 H5 + O
CH3 + CH2 O 1.000 · 1014 0.0 0.0
90. C2 H5 + HO2
CH3 + CH2 O + OH 3.000 · 1013 0.0 0.0
91. C2 H5 + O2
C2 H4 + HO2 3.000 · 1020 −2.9 28.3
92. C2 H5 + O2
C2 H4 + HO2 2.120 · 10−6 6.0 39.7
93. C2 H4 + H
C2 H3 + H2 3.360 · 10−7 6.0 7.1
94. C2 H4 + OH
C2 H3 + H2 O 2.020 · 1013 0.0 24.8
95. C2 H4 + O
CH3 + CHO 1.020 · 107 1.9 0.7
96. C2 H4 + O
CH2 CHO + H 3.390 · 106 1.9 0.7
97. C2 H4 + CH3
C2 H3 + CH4 6.620 · 100 3.7 39.7
98. C2 H4 + H + M(4)
C2 H5 + M(4) 1.080 · 1012 0.5 7.6
Low: 1.11 · 1034 −5.0 18.5
Troe: 1.00 1.0 · 10−15 95 200
99. C2 H4 + M
C2 H2 + H2 + M 1.800 · 1013 0.0 318.0
Low: 1.5 · 1015 0.0 232.7
A.2 Oxy-dehydrogenation of ethane 161

Reaction A β Ea
100. C2 H3 + H + M(5)
C2 H4 + M(5) 6.100 · 1012 0.3 1.2
Low: 9.80 · 1029 −3.86 13.9
Troe: 0.782 2.080 · 1002 2663 6095
101. C2 H3 + H
C2 H2 + H2 4.000 · 1013 0.0 0.0
102. C2 H3 + O
CH2 CO + H 3.000 · 1013 0.0 0.0
103. C2 H3 + O2
CH2 O + CHO 1.700 · 1029 −5.3 27.2
104. C2 H3 + O2
CH2 CHO + O 3.500 · 1014 −0.6 22.0
105. C2 H3 + O2
C2 H2 + HO2 2.120 · 10−6 6.0 39.7
106. C2 H3 + OH
C2 H2 + H2 O 2.000 · 1013 0.0 0.0
107. C2 H3 + CH3
C2 H2 + CH4 2.000 · 1013 0.0 0.0
108. C2 H3 + C2 H3
C2 H4 + C2 H2 1.450 · 1013 0.0 0.0
109. C2 H2 + OH
CH2 CO + H 2.180 · 10−4 4.5 −4.2
110. C2 H2 + OH
CH2 CO + H 2.000 · 1011 0.0 0.0
111. C2 H2 + OH
CH3 + CO 4.830 · 10−4 4.0 −8.4
112. C2 H2 + O
3-CH2 + CO 6.120 · 106 2.0 7.9
113. C2 H2 + O
HCCO + H 1.430 · 107 2.0 7.9
114. C2 H2 + O2
HCCO + OH 4.000 · 107 1.5 125.9
115. C2 H2 + H + M(4)
C2 H3 + M(4) 3.110 · 1011 0.6 10.8
Low: 2.25 · 1040 −7.27 27.6
Troe: 1.00 1.0 · 10−15 675 1.0 · 1015
116. CH2 CHO + H
CH2 CO + H2 4.000 · 1013 0.0 0.0
117. CH2 CHO + O
CH2 O + CHO 1.000 · 1014 0.0 0.0
118. CH2 CHO + OH
CH2 CO + H2 O 3.000 · 1013 0.0 0.0
119. CH2 CHO + O2
CH2 O + CO + OH 3.000 · 1010 0.0 0.0
120. CH2 CHO + CH3 → C2 H5 + CO + H 4.900 · 1014 −0.5 0.0
121. CH2 CHO
CH2 CO + H 3.950 · 1038 −7.6 188.8
122. CH2 CO + O
CO2 + 3-CH2 1.750 · 1012 0.0 5.6
123. CH2 CO + H
CH3 + CO 7.000 · 1012 0.0 12.6
124. CH2 CO + H
HCCO + H2 2.000 · 1014 0.0 33.5
125. CH2 CO + O
HCCO + OH 1.000 · 1013 0.0 33.5
126. CH2 CO + OH
HCCO + H2 O 1.000 · 1013 0.0 8.4
127. 3-CH2 + CO + M(5)
CH2 CO + M(5) 8.100 · 1011 0.5 18.9
Low: 1.90 · 1033 −5.11 29.7
Troe: 0.591 2.750 · 1002 1226 5185.0
128. HCCO + O
H + CO + CO 8.000 · 1013 0.0 0.0
129. HCCO + O2
CHO + CO + O 2.500 · 108 1.0 0.0
130. HCCO + O2
CO2 + CHO 2.400 · 1011 0.0 −3.6
131. HCCO + HCCO
C2 H2 + CO + CO 1.000 · 1013 0.0 0.0

H2 O H2 CH4 CO2 CO
M(1) 0.00 0.00 10.00 3.80 1.90
M(2) 0.00 0.00 1.00 1.00 1.00
M(3) 6.40 1.00 1.00 1.00 1.00
M(4) 5.00 2.00 1.00 3.00 2.00
M(5) 5.00 1.00 1.00 1.00 1.00
M(6) 5.00 1.87 2.81 3.00 1.87
162 B. SURFACE REACTION MECHANISMS

B Surface reaction mechanisms


In this section, all heterogeneous reaction mechanisms are listed. The mechanisms are
given application by application, even though the same reactions frequently occur in
different schemes.
The development of detailed reaction mechanisms is a long process (Chapter 3.2.3).
Mechanisms have continously to be revised after new experimental or theoretical inves-
tigations. For instance, the reaction mechanism for the partial oxidation of methane
to synthesis gas, developed by Hickman and Schmidt in the early 1990s (Tab. B.1),
was recently revised due to new experimental observations as discussed in Chapter 5;
the revised mechanism is given in Tab. B.2.
The heterogeneous reaction schemes of hydrogen, carbon monoxide, light hydro-
carbons, and oxygen on platinum catalysts has been becoming increasingly detailed
in the last few years. For instance, the mechanism for complete oxidation of methane
on platinum, developed in 1995 (Tab. B.4), describes the decomposition of methane
on the catalyst in a very simple manner. The mechanism for oxy-dehydrogenation of
ethane to ethylene, developed five years later (Tab. B.3), now includes more than a
dozen steps for methane decomposition and formation.
In the tables, the symbol denotes that the kinetic data of the reverse reaction
are calculated by Eq. 3.26. Species with the suffix (s) are adsorbed species. The
species named Pt(s) and Rh(s) denote uncovered surface sites available for adsorption
on platinum and rhodium, respectively.
The kinetic data are given according to Eq. 3.24. The units are: A [mol, cm, s], S 0 ,
β and µ [-], Ea and  [kJ mol−1 ]. S 0 denotes the initial (uncovered surface) sticking
coefficient. If the reaction kinetics exhibits an additional dependence on surface cov-
erage (Eq. 3.24), the line under the reaction equation names the species, to which the
dependence is referred, and gives the kinetic parameters µ and . µ and  are zero for
all other reactions.
B.1 Partial oxidation of methane on rhodium - I 163

B.1 Partial oxidation of methane on rhodium - I


This mechanism, developed by Hickman and Schmidt in 1993 [122], is used in the
first numerical simulations of partial oxidation of methane on rhodium catalysts as
discussed in Chapter 5).

Reaction A / S0 β /µ Ea / 
I Adsorption
1. H2 + Rh(s) + Rh(s) → H(s) + H(s) 8.400 · 1018 1.0 0.0
θRh(s) -1.0
2. O2 + Rh(s1) + Rh(s1) → O(s1) + O(s1) 1.300 · 1017 1.0 0.0
θRh(s1) -1.0
3. CH4 + Rh(s) + Rh(s) → CH3 (s) + H(s) 1.100 · 1018 1.0 21.0
θRh(s) -1.0
4. H2 O + Rh(s) → H2 O(s) 4.600 · 109 1.0 0.0
5. CO + Rh(s) → CO(s) 1.200 · 1010 1.0 0.0
II Desorption
6. H(s) + H(s) → Rh(s) + Rh(s) + H2 3.000 · 1021 0.0 75.4
θH(s) -1.0
7. O(s1) + O(s1) → Rh(s1) + Rh(s1) + O2 3.000 · 1021 0.0 293.3
θO(s1) -1.0
8. H2 O(s) → H2 O + Rh(s) 1.000 · 1013 0.0 45.2
9. CO(s) → CO + Rh(s) 4.000 · 1013 0.0 132.4
θCO(s) 44.0
10. OH(s) → OH + Rh(s) 8.100 · 1011 0.0 142.5
III Surface reactions
11. O(s1) + H(s) → OH(s) + Rh(s1) 4.200 · 1021 0.0 83.8
12. OH(s) + Rh(s1) → O(s1) + H(s) 6.000 · 1021 0.0 21.0
13. H(s) + OH(s) → H2 O(s) + Rh(s) 1.800 · 1026 0.0 33.5
14. H2 O(s) + Rh(s) → H(s) + OH(s) 3.000 · 1023 0.0 155.0
15. OH(s) + OH(s) → H2 O(s) + O(s) 2.400 · 1024 0.0 62.8
16. O(s) + Rh(s1) → Rh(s) + O(s1) 3.700 · 1021 0.0 0.0
17. C(s) + O(s1) → CO(s) + Rh(s1) 3.000 · 1022 0.0 62.8
18. CO(s) + Rh(s1) → C(s) + O(s1) 6.000 · 1019 0.0 167.6
19. CO(s) + O(s1) → CO2 + Rh(s) + Rh(s1) 6.000 · 1020 0.0 104.7
20. CH3 (s) + Rh(s) → CH2 (s) + H(s) 3.700 · 1021 0.0 0.0
21. CH2 (s) + Rh(s) → CH(s) + H(s) 3.700 · 1021 0.0 0.0
22. CH(s) + Rh(s) → C(s) + H(s) 3.700 · 1021 0.0 0.0

The suffixes (s) and (s1) denote different adsorption sites, the latter one available for oxygen adsorption only.

Reaction steps 3. and 20.-22. represent only one recation in the original mechanism being first order in Rh(s). Splitting
up this reaction in an adsorption step, again being first order in Rh(s), and fast consecutive decomposition steps (20.-22.)
does not change any results.
164 B. SURFACE REACTION MECHANISMS

B.2 Partial oxidation of methane on rhodium - II


This mechanism, developed in 2000 [142], is used for the more recent numerical simu-
lation of partial oxidation of methane on rhodium catalysts (Chapter 5).

Reaction A / S0 β /µ Ea / 
I Adsorption
1. H2 + Rh(s) + Rh(s) → H(s) + H(s) S0 = 1.0 · 10−2 0.0 0.0
2. O2 + Rh(s) + Rh(s) → O(s) + O(s) S0 = 1.0 · 10−2 0.0 0.0
3. CH4 + Rh(s) → CH4 (s) S0 = 8.0 · 10−3 0.0 0.0
4. H2 O + Rh(s) → H2 O(s) S0 = 1.0 · 10−1 0.0 0.0
5. CO2 + Rh(s) → CO2 (s) S0 = 1.0 · 10−5 0.0 0.0
6. CO + Rh(s) → CO(s) S0 = 5.0 · 10−1 0.0 0.0
II Desorption
7. H(s) + H(s) → Rh(s) + Rh(s) + H2 3.000 · 1021 0.0 77.8
8. O(s) + O(s) → Rh(s) + Rh(s) + O2 1.300 · 1022 0.0 355.2
9. H2 O(s) → H2 O + Rh(s) 3.000 · 1013 0.0 45.0
10. CO(s) → CO + Rh(s) 3.500 · 1013 0.0 133.4
11. CO2 (s) → CO2 + Rh(s) 1.000 · 1013 0.0 21.7
12. CH4 (s) → CH4 + Rh(s) 1.000 · 1013 0.0 25.1
III Surface reactions
13. H(s) + O(s) → OH(s) + Rh(s) 5.000 · 1022 0.0 83.7
14. OH(s) + Rh(s) → H(s) + O(s) 3.000 · 1020 0.0 37.7
15. H(s) + OH(s) → H2 O(s) + Rh(s) 3.000 · 1020 0.0 33.5
16. H2 O(s) + Rh(s) → H(s) + OH(s) 5.000 · 1022 0.0 106.4
17. OH(s) + OH(s) → H2 O(s) + O(s) 3.000 · 1021 0.0 100.8
18. H2 O(s) + O(s) → OH(s) + OH(s) 3.000 · 1021 0.0 224.2
19. C(s) + O(s) → CO(s) + Rh(s) 3.000 · 1022 0.0 97.9
20. CO(s) + Rh(s) → C(s) + O(s) 2.500 · 1021 0.0 169.0
21. CO(s) + O(s) → CO2 (s) + Rh(s) 1.400 · 1020 0.0 121.6
22. CO2 (s) + Rh(s) → CO(s) + O(s) 3.000 · 1021 0.0 115.3
23. CH4 (s) + Rh(s) → CH3 (s) + H(s) 3.700 · 1021 0.0 61.0
24. CH3 (s) + H(s) → CH4 (s) + Rh(s) 3.700 · 1021 0.0 51.0
25. CH3 (s) + Rh(s) → CH2 (s) + H(s) 3.700 · 1024 0.0 103.0
26. CH2 (s) + H(s) → CH3 (s) + Rh(s) 3.700 · 1021 0.0 44.0
27. CH2 (s) + Rh(s) → CH(s) + H(s) 3.700 · 1024 0.0 100.0
28. CH(s) + H(s) → CH2 (s) + Rh(s) 3.700 · 1021 0.0 68.0
29. CH(s) + Rh(s) → C(s) + H(s) 3.700 · 1021 0.0 21.0
30. C(s) + H(s) → CH(s) + Rh(s) 3.700 · 1021 0.0 172.8
31. CH4 (s) + O(s) → CH3 (s) + OH(s) 1.700 · 1024 0.0 80.3
32. CH3 (s) + OH(s) → CH4 (s) + O(s) 3.700 · 1021 0.0 24.3
33. CH3 (s) + O(s) → CH2 (s) + OH(s) 3.700 · 1024 0.0 120.3
34. CH2 (s) + OH(s) → CH3 (s) + O(s) 3.700 · 1021 0.0 15.1
35. CH2 (s) + O(s) → CH(s) + OH(s) 3.700 · 1024 0.0 158.4
36. CH(s) + OH(s) → CH2 (s) + O(s) 3.700 · 1021 0.0 36.8
37. CH(s) + O(s) → C(s) + OH(s) 3.700 · 1021 0.0 30.1
38. C(s) + OH(s) → CH(s) + O(s) 3.700 · 1021 0.0 145.5
B.3 Oxy-dehydrogenation of ethane on platinum 165

B.3 Oxy-dehydrogenation of ethane on platinum


This mechanism is used for the numerical simulation of oxy-dehydrogenation of ethane
on platinum catalysts as described in Chapter 6. The development of the mechanism
is discussed in [49].

Reaction A / S0 β /µ Ea / 
I Adsorption
1. H + Pt(s) → H(s) S 0 = 1.0 · 10−0 0.0 0.0
2.. H2 + Pt(s) + Pt(s) → H(s) + H(s) S 0 = 4.6 · 10−2 0.0 0.0
θPt(s) -1.0
3. H2 + C(s) → CH2 (s) S 0 = 4.0 · 10−2 0.0 29.7
θC(s) -4.6
4. O + Pt(s) → O(s) S 0 = 1.0 · 10−0 0.0 0.0
5. O2 + Pt(s) + Pt(s) → O(s) + O(s) 1.891 · 1021 −0.5 0.0
6. OH + Pt(s) → OH(s) S 0 = 1.0 · 10−0 0.0 0.0
7. H2 O + Pt(s) → H2 O(s) S 0 = 7.5 · 10−1 0.0 0.0
8. CO + Pt(s) → CO(s) S 0 = 8.4 · 10−1 0.0 0.0
9. CO2 + Pt(s) → CO2 (s) S 0 = 5.0 · 10−3 0.0 0.0
10. CH3 + Pt(s) → CH3 (s) S 0 = 1.0 · 10−0 0.0 0.0
11. CH4 + C(s) → CHCH3 (s) S 0 = 7.0 · 10−9 0.0 23.0
θC(s) -47.5
12. CH4 + Pt(s) + Pt(s) → CH3 (s) + H(s) S 0 = 9.0 · 10−4 0.0 72.2
13. CH4 + O(s) + Pt(s) → CH3 (s) + OH(s) 5.000 · 1018 0.7 42.0
θO(s) -8.0
14. CH4 + OH(s) + Pt(s) → CH3 (s) + H2 O(s) S0 = 1.0 · 10−0 0.0 10.0
15. C2 H2 + Pt(s) → C2 H2 (s) S0 = 5.0 · 10−2 0.0 0.0
16. C2 H4 + Pt(s) → C2 H4 (s) S0 = 1.5 · 10−2 0.0 0.0
17. C2 H5 + Pt(s) → C2 H5 (s) S0 = 1.0 · 10−0 0.0 0.0
18. C2 H6 + Pt(s) + Pt(s) → C2 H6 (s) S0 = 1.5 · 10−2 0.0 0.0
II Desorption
19. H(s) → H + Pt(s) 6.000 · 1013 0.0 254.4
θH(s) 5.0
20. H(s) + H(s) → Pt(s) + Pt(s) + H2 3.700 · 1021 0.0 67.4
θH(s) 10.0
21. CH2 (s) → C(s) + H2 7.690 · 1013 0.0 25.1
θC(s) -50.0
22. O(s) → O + Pt(s) 1.000 · 1013 0.0 358.8
θO(s) 94.1
23. O(s) + O(s) → Pt(s) + Pt(s) + O2 3.700 · 1021 0.0 227.4
θO(s) 188.3
24. OH(s) → OH + Pt(s) 5.000 · 1013 0.0 251.1
θO(s) 167.4
25. H2 O(s) → H2 O + Pt(s) 4.500 · 1012 0.0 41.8
26. CO(s) → CO + Pt(s) 2.500 · 1016 0.0 146.0
θCO(s) 33.0
27. CO2 (s) → CO2 + Pt(s) 1.000 · 1013 0.0 27.1
28. CH3 (s) → Pt(s) + CH3 1.000 · 1013 0.0 163.0
29. CH3 (s) + H(s) → CH4 + Pt(s) + Pt(s) 1.500 · 1020 0.0 50.0
30. CH3 (s) + H2 O(s) → CH4 + OH(s) + Pt(s) 2.500 · 1020 0.0 23.0
θH(s) 50.0
31. CH3 (s) + OH(s) → CH4 + O(s) + Pt(s) 3.700 · 1021 0.0 85.9
32. C2 H2 (s) → Pt(s) + C2 H2 1.000 · 1012 0.0 58.6
33. C2 H4 (s) → Pt(s) + C2 H4 1.000 · 1013 0.0 50.2
34. CHCH3 (s) → C(s) + CH4 1.000 · 1010 0.0 25.5
θC(s) -47.5
35. C2 H5 (s) → Pt(s) + C2 H5 1.000 · 1013 0.0 173.0
36. C2 H6 (s) → Pt(s) + Pt(s) + C2 H6 1.000 · 1013 0.0 20.9
166 B. SURFACE REACTION MECHANISMS

Reaction A / S0 β /µ Ea / 
III Surface reactions
37. H(s) + O(s) → OH(s) + Pt(s) 1.280 · 1021 0.0 11.2
38. OH(s) + Pt(s) → H(s) + O(s) 7.390 · 1019 0.0 77.3
θO(s) 73.2
39. H(s) + OH(s) → H2 O(s) + Pt(s) 2.040 · 1021 0.0 66.2
40. H2 O(s) + Pt(s) → H(s) + OH(s) 1.150 · 1019 0.0 101.4
θO(s) -167.4
41. OH(s) + OH(s) → H2 O(s) + O(s) 7.400 · 1020 0.0 74.0
42. H2 O(s) + O(s) → OH(s) + OH(s) 1.000 · 1020 0.0 43.1
θO(s) -240.6
43. C(s) + O(s) → CO(s) + Pt(s) 3.700 · 1019 0.0 0.0
44. CO(s) + Pt(s) → C(s) + O(s) 3.700 · 1019 0.0 236.5
θCO(s) 33.0
45. CO(s) + O(s) → CO2 (s) + Pt(s) 3.700 · 1019 0.0 117.6
θCO(s) 33.0
46. CO2 (s) + Pt(s) → CO(s) + O(s) 3.700 · 1019 0.0 173.3
θO(s) -94.1
47. CO(s) + OH(s) → CO2 (s) + H(s) 2.000 · 1019 0.0 38.7
θCO(s) 30.0
48. CO2 (s) + H(s) → CO(s) + OH(s) 2.000 · 1019 0.0 28.3
49. CH3 (s) + Pt(s) → CH2 (s) + H(s) 1.262 · 1022 0.0 70.3
50. CH2 (s) + H(s) → CH3 (s) + Pt(s) 3.090 · 1022 0.0 0.0
θH(s) 2.8
51. CH2 (s) + Pt(s) → CH(s) + H(s) 7.314 · 1022 0.0 58.9
θC(s) -50.0
52. CH(s) + H(s) → CH2 (s) + Pt(s) 3.090 · 1022 0.0 0.0
θH(s) 2.8
53. CH(s) + Pt(s) → C(s) + H(s) 3.090 · 1022 0.0 0.0
θH(s) 2.8
54. C(s) + H(s) → CH(s) + Pt(s) 1.248 · 1022 0.0 138.0
55. C2 H6 (s) + O(s) → C2 H5 (s) + OH(s) + Pt(s) 3.700 · 1021 0.0 25.1
56. C2 H5 (s) + OH(s) + Pt(s) → C2 H6 (s) + O(s) 1.350 · 1030 0.0 77.4
57. C2 H4 (s) → CHCH3 (s) 1.000 · 1013 0.0 83.3
58. CHCH3 (s) → C2 H4 (s) 1.000 · 1013 0.0 75.3
59. C2 H5 (s) + H(s) → C2 H6 (s) 3.700 · 1021 0.0 41.8
60. C2 H6 (s) → C2 H5 (s) + H(s) 7.000 · 1012 0.0 57.7
61. CH3 (s) + CH3 (s) → C2 H6 (s) 1.000 · 1021 0.0 14.5
62. C2 H6 (s) → CH3 (s) + CH3 (s) 1.000 · 1013 0.0 89.0
63. C2 H5 (s) + Pt(s) → CHCH3 (s) + H(s) 1.000 · 1022 0.0 54.4
64. CHCH3 (s) + H(s) → C2 H5 (s) + Pt(s) 1.000 · 1021 0.0 29.3
65. CHCH3 (s) + Pt(s) → CCH3 (s) + H(s) 2.000 · 1022 0.0 99.1
66. CCH3 (s) + H(s) → CHCH3 (s) + Pt(s) 3.700 · 1021 0.0 75.3
67. CHCH3 (s) + Pt(s) → CHCH2 (s) + H(s) 3.700 · 1021 0.0 128.5
68. CHCH2 (s) + H(s) → CHCH3 (s) + Pt(s) 3.700 · 1021 0.0 57.3
69. C2 H4 (s) + Pt(s) → CHCH2 (s) + H(s) 3.700 · 1021 0.0 112.7
70. CHCH2 (s) + H(s) → C2 H4 (s) + Pt(s) 3.700 · 1021 0.0 33.5
71. CHCH2 (s) + Pt(s) → CCH2 (s) + H(s) 3.700 · 1021 0.0 121.3
72. CCH2 (s) + H(s) → CHCH2 (s) + Pt(s) 3.700 · 1021 0.0 51.7
73. CCH3 (s) + Pt(s) → CH3 (s) + C(s) 3.700 · 1021 0.0 46.9
θC(s) -50.0
74. CH3 (s) + C(s) → CCH3 (s) + Pt(s) 3.700 · 1021 0.0 46.0
75. C2 H2 (s) → CCH2 (s) 1.000 · 1013 0.0 61.5
76. CCH2 (s) → C2 H2 (s) 1.000 · 1013 0.0 4.2
77. CCH3 (s) → CHCH2 (s) 1.000 · 1013 0.0 176.0
78. CHCH2 (s) → CCH3 (s) 1.000 · 1013 0.0 128.6
79. C2 H2 (s) + Pt(s) → CCH(s) + H(s) 3.700 · 1021 0.0 133.5
80. CCH(s) + H(s) → C2 H2 (s) + Pt(s) 3.700 · 1021 0.0 66.9
81. CCH(s) + Pt(s) → CH(s) + C(s) 3.700 · 1021 0.0 125.1
82. CH(s) + C(s) → CCH(s) + Pt(s) 3.700 · 1021 0.0 121.3
B.4 Oxidation of hydrogen, carbon monoxide, and methane on platinum 167

B.4 Oxidation of hydrogen, carbon monoxide, and methane


on platinum
This mechanism is used for the numerical simulation of the catalytic combustion of
methane on platinum catalysts as described in Chapter 7. The development of the
mechanism is discussed in [45, 66].

Reaction A / S0 β /µ Ea / 
I AdsorPtion
1. H2 + Pt(s) + Pt(s) → H(s) + H(s) S 0 = 4.6 · 10−2 0.0 0.0
θPt(s) -1.0
2. H + Pt(s) → H(s) S 0 = 1.0 · 10−0 0.0 0.0
3. O2 + Pt(s) + Pt(s) → O(s) + O(s) 1.800 · 1021 −0.5 0.0
4. O2 + Pt(s) + Pt(s) → O(s) + O(s) S 0 = 2.3 · 10−2 0.0 0.0
5. CH4 + Pt(s) + Pt(s) → CH3 (s) + H(s) S 0 = 1.0 · 10−2 0.0 0.0
θPt(s) 0.3
6. O + Pt(s) → O(s) S 0 = 1.0 · 10−0 0.0 0.0
7. H2 O + Pt(s) → H2 O(s) S 0 = 7.5 · 10−1 0.0 0.0
8. CO + Pt(s) → CO(s) S 0 = 8.4 · 10−1 0.0 0.0
θPt(s) 1.0
9. OH + Pt(s) → OH(s) S 0 = 1.0 · 10−0 0.0 0.0
II DesorPtion
10. H(s) + H(s) → Pt(s) + Pt(s) + H2 3.700 · 1021 0.0 67.4
θH(s) 6.0
11. O(s) + O(s) → Pt(s) + Pt(s) + O2 3.700 · 1021 0.0 213.0
θO(s) 60.0
12. H2 O(s) → H2 O + Pt(s) 1.000 · 1013 0.0 40.3
13. OH(s) → OH + Pt(s) 1.000 · 1013 0.0 192.8
14. CO(s) → CO + Pt(s) 1.000 · 1013 0.0 125.5
15. CO2 (s) → CO2 + Pt(s) 1.000 · 1013 0.0 20.5
16. O(s) + H(s)
OH(s) + Pt(s) 3.700 · 1021 0.0 11.5
III Surface reactions
17. H(s) + OH(s)
H2 O(s) + Pt(s) 3.700 · 1021 0.0 17.4
18. OH(s) + OH(s)
H2 O(s) + O(s) 3.700 · 1021 0.0 48.2
19. CO(s) + O(s) → CO2 (s) + Pt(s) 3.700 · 1021 0.0 105.0
20. C(s) + O(s) → CO(s) + Pt(s) 3.700 · 1021 0.0 62.8
21. CO(s) + Pt(s) → C(s) + O(s) 1.000 · 1018 0.0 184.0
22. CH3 (s) + Pt(s) → CH2 (s) + H(s) 3.700 · 1021 0.0 20.0
23. CH2 (s) + Pt(s) → CH(s) + H(s) 3.700 · 1021 0.0 20.0
24. CH(s) + Pt(s) → C(s) + H(s) 3.700 · 1021 0.0 20.0

Reaction 3. and 4. represent alternative competing pathways.


168 B. SURFACE REACTION MECHANISMS

B.5 Mechanism for a three-way catalyst


This mechanism is used for the numerical simulation of the chemical reactions in a
three-way catalyst as described in Chapter 8.2. The propylene species presents the
hydrocarbon species in the exhaust gas. The development of the mechanism is discussed
in [203, 88]. While species adsorbed on platinum are denoted by the suffix (s), species
adsorbed on rhodium are denoted by the suffix (Rh).

Reaction A / S0 β /µ Ea / 
C3 H6 /CO - oxidation on Pt
I Adsorption
1. O2 + Pt(s) + Pt(s) → O(s) + O(s) S 0 = 7.0 · 10−2 0.0 0.0
2. C3 H6 + Pt(s) + Pt(s) → C3 H6 (s) S 0 = 9.8 · 10−1 0.0 0.0
3. C3 H6 + O(s) + Pt(s) → C3 H5 (s) + OH(s) S 0 = 5.0 · 10−2 0.0 0.0
θPt(s) -0.9
4. H2 + Pt(s) + Pt(s) → H(s) + H(s) S 0 = 4.6 · 10−2 0.0 0.0
θPt(s) -1.0
5. H2 O + Pt(s) → H2 O(s) S 0 = 7.5 · 10−1 0.0 0.0
6. CO2 + Pt(s) → CO2 (s) S 0 = 5.0 · 10−3 0.0 0.0
7. CO + Pt(s) → CO(s) S 0 = 8.4 · 10−1 0.0 0.0
II Desorption
8. O(s) + O(s) → O2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 232.2
θO(s) 90.0
9. C3 H6 (s) → C3 H6 + Pt(s) + Pt(s) 1.00 · 1013 0.0 72.7
10. C3 H5 (s) + OH(s) → C3 H6 + O(s) + Pt(s) 3.70 · 1021 0.0 31.0
11. H(s) + H(s) → H2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 67.4
θH(s) 6.0
12. H2 O(s) → Pt(s) + H2 O 1.00 · 1013 0.0 40.3
13. CO(s) → CO + Pt(s) 1.00 · 1013 0.0 136.4
θCO(s) 33.0
14. CO2 (s) → CO2 + Pt(s) 1.00 · 1013 0.0 27.1
III Surface reactions
IIIa C3 H5 (s)-oxidation (global reaction)
15. C3 H5 (s) + 5O(s) → 3C(s) + 5OH(s) 3.70 · 1021 0.0 95.0
IIIb C3 H6 (s) -decomposition
16. C3 H6 (s) → CC2 H5 (s) + H(s) 1.00 · 1013 0.0 75.4
17. CC2 H5 (s) + H(s) → C3 H6 (s) 3.70 · 1021 0.0 48.8
18. CC2 H5 (s) + Pt(s) → C2 H3 (s) + CH2 (s) 3.70 · 1021 0.0 108.2
19. C2 H3 (s) + CH2 (s) → CC2 H5 (s) + Pt(s) 3.70 · 1021 0.0 3.2
20. C2 H3 (s) + Pt(s) → CH3 (s) + C(s) 3.70 · 1021 0.0 46.0
21. CH3 (s) + C(s) → C2 H3 (s) + Pt(s) 3.70 · 1021 0.0 46.9
IIIc CHx -decomposition
22. CH3 (s) + Pt(s) → CH2 (s) + H(s) 1.26 · 1022 0.0 70.4
23. CH2 (s) + H(s) → CH3 (s) + Pt(s) 3.09 · 1022 0.0 0.0
24. CH2 (s) + Pt(s) → CH(s) + H(s) 7.00 · 1022 0.0 59.2
25. CH(s) + H(s) → CH2 (s) + Pt(s) 3.09 · 1022 0.0 0.0
26. CH(s) + Pt(s) → C(s) + H(s) 3.09 · 1022 0.0 0.0
27. C(s) + H(s) → CH(s) + Pt(s) 1.25 · 1022 0.0 138.0
IIId C2 Hx -oxidation
28. C2 H3 (s) + O(s) → CH3 CO(s) + Pt(s) 3.70 · 1019 0.0 62.3
29. CH3 CO(s) + Pt(s) → C2 H3 (s) + O(s) 3.70 · 1021 0.0 196.7
θO(s) -45.0
30. CH3 (s) + CO(s) → CH3 CO(s) + Pt(s) 3.70 · 1021 0.0 82.9
31. CH3 CO(s) + Pt(s) → CH3 (s) + CO(s) 3.70 · 1021 0.0 0.0
θCO(s) -33.0
B.5 Mechanism for a three-way catalyst 169

Reaction A / S0 β /µ Ea / 
IIIe CHx -oxidation
32. CH3 (s) + O(s) → CH2 (s) + OH(s) 3.70 · 1021 0.0 36.6
33. CH2 (s) + OH(s) → CH3 (s) + O(s) 3.70 · 1021 0.0 25.1
34. CH2 (s) + O(s) → CH(s) + OH(s) 3.70 · 1021 0.0 25.1
35. CH(s) + OH(s) → CH2 (s) + O(s) 3.70 · 1021 0.0 25.2
36. CH(s) + O(s) → C(s) + OH(s) 3.70 · 1021 0.0 25.1
37. C(s) + OH(s) → CH(s) + O(s) 3.70 · 1021 0.0 224.8
IIIf H, OH, H2 O - reactions
38. O(s) + H(s) → OH(s) + Pt(s) 3.70 · 1021 0.0 11.5
39. OH(s) + Pt(s) → O(s) + H(s) 5.77 · 1022 0.0 74.9
40. H(s) + OH(s) → H2 O(s) + Pt(s) 3.70 · 1021 0.0 17.4
41. H2 O(s) + Pt(s) → H(s) + OH(s) 3.66 · 1021 0.0 73.6
42. OH(s) + OH(s) → H2 O(s) + O(s) 3.70 · 1021 0.0 48.2
43. H2 O(s) + O(s) → OH(s) + OH(s) 2.35 · 1020 0.0 41.0
IIIg CO-oxidation
44. CO(s) + O(s) → CO2 (s) + Pt(s) 3.70 · 1020 0.0 108.0
θCO(s) 33.0
θNO(s) -90.0
45. CO2 (s) + Pt(s) → CO(s) + O(s) 3.70 · 1021 0.0 165.1
θO(s) -45.0
46. C(s) + O(s) → CO(s) + Pt(s) 3.70 · 1021 0.0 0.0
θCO(s) -33.0
47. CO(s) + Pt(s) → C(s) + O(s) 3.70 · 1021 0.0 218.5
θO(s) -45.0
NO - reduction on Pt
I Adsorption
48. NO + Pt(s) → NO(s) 8.50 · 10−1 0.0 0.0
II Desorption
49. NO(s) → NO + Pt(s) 1.00 · 1016 0.0 140.0
50. N(s) + N(s) → N2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 113.9
θCO(s) 75.0
III NO - surface reactions
51. NO(s) + Pt(s) → N(s) + O(s) 5.00 · 1020 0.0 107.8
θCO(s) -3.0
52. N(s) + O(s) → NO(s) + Pt(s) 3.70 · 1021 0.0 128.1
θO(s) 45.0
NO/CO - reactions on Rh
I Adsorption
53. O2 + Rh(s) + Rh(s) → O(Rh) + O(Rh) 1.00 · 10−2 0.0 0.0
θRh(s) -1.0
54. CO + Rh(s) → CO(Rh) 5.00 · 10−1 0.0 0.0
55. NO + Rh(s) → NO(Rh) 5.00 · 10−1 0.0 0.0
II Desorption
56. O(Rh) + O(Rh) → O2 + Rh(s) + Rh(s) 3.00 · 1021 0.0 293.3
57. CO(Rh) → CO + Rh(s) 1.00 · 1014 0.0 132.3
θN(Rh) 41.9
θCO(Rh) 18.8
58. NO(Rh) → NO + Rh(s) 5.00 · 1013 0.0 108.9
59. N(Rh) + N(Rh) → N2 + Rh(S) + Rh(S) 1.11 · 1019 0.0 136.9
θN(Rh) 16.7
III NO/CO-surface reactions
60. CO(Rh) + O(Rh) → CO2 + Rh(s) + Rh(s) 3.70 · 1020 0.0 59.9
61. NO(Rh) + Rh(s) → N(Rh) + O(Rh) 2.22 · 1022 0.0 79.5

Reaction 15. is a complex fast reaction that takes place if sufficient O(s) is available for Reaction 3. Reaction 15. is
first order in O(s) and C3 H5 (s).
170 B. SURFACE REACTION MECHANISMS

B.6 HC-SCR on Pt/Al2 O3


This mechanism is used for the numerical simulation of the HC-SCR in a DeNOx
catalyst as described in Chapter 8.3. The development of the mechanism is discussed
in [88, 89].

Reaction A / S0 β /µ Ea / 
C3 H6 - oxidation
I Adsorption
1. O2 + Pt(s) + Pt(s) → O(s) + O(s) S 0 = 7.0 · 10−2 0.0 0.0
2. C3 H6 + Pt(s) + Pt(s) → C3 H6 (s) S 0 = 9.8 · 10−1 0.0 0.0
3. C3 H6 + O(s) + Pt(s) → C3 H5 (s) + OH(s) S 0 = 5.0 · 10−2 0.0 0.0
θPt(s) -0.9
4. H2 + Pt(s) + Pt(s) → H(s) + H(s) S 0 = 4.6 · 10−2 0.0 0.0
θPt(s) -1.0
5. H2 O + Pt(s) → H2 O(s) S0 = 7.5 · 10−1 0.0 0.0
6. OH + Pt(s) → OH(s) S0 = 1.0 · 10−0 0.0 0.0
7. CO2 + Pt(s) → CO2 (s) S0 = 5.0 · 10−3 0.0 0.0
8. CO + Pt(s) → CO(s) S0 = 8.4 · 10−1 0.0 0.0
II Desorption
9. O(s) + O(s) → O2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 232.2
θO(s) 90.0
10. C3 H6 (s) → C3 H6 + Pt(s) + Pt(s) 1.00 · 1013 0.0 72.7
11. C3 H5 (s) + OH(s) → C3 H6 + O(s) + Pt(s) 3.70 · 1021 0.0 31.0
12. H(s) + H(s) → H2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 67.4
θH(s) 6.0
13. H2 O(s) → Pt(s) + H2 O 1.00 · 1013 0.0 40.3
14. OH(s) → Pt(s) + OH 1.00 · 1013 0.0 251.4
15. CO(s) → CO + Pt(s) 1.00 · 1013 0.0 146.4
θCO(s) 33.0
16. CO2 (s) → CO2 + Pt(s) 1.00 · 1013 0.0 27.1
III Surface reactions
IIIa C3 H5 (s)-oxidation (global reaction)
17. C3 H5 (s) + 5O(s) → 3C(s) + 5OH(s) 3.70 · 1021 0.0 95.0
IIIb C3 H6 (s) -decomposition
18. C3 H6 (s) → CC2 H5 (s) + H(s) 1.00 · 1013 0.0 75.4
19. CC2 H5 (s) + H(s) → C3 H6 (s) 3.70 · 1021 0.0 48.8
20. CC2 H5 (s) + Pt(s) → C2 H3 (s) + CH2 (s) 3.70 · 1021 0.0 108.2
21. C2 H3 (s) + CH2 (s) → CC2 H5 (s) + Pt(s) 3.70 · 1021 0.0 3.2
22. C2 H3 (s) + Pt(s) → CH3 (s) + C(s) 3.70 · 1021 0.0 46.0
23. CH3 (s) + C(s) → C2 H3 (s) + Pt(s) 3.70 · 1021 0.0 46.9
24. CH3 (s) + Pt(s) → CH2 (s) + H(s) 1.26 · 1022 0.0 70.4
25. CH2 (s) + H(s) → CH3 (s) + Pt(s) 3.09 · 1022 0.0 0.0
26. CH2 (s) + Pt(s) → CH(s) + H(s) 7.00 · 1022 0.0 59.2
27. CH(s) + H(s) → CH2 (s) + Pt(s) 3.09 · 1022 0.0 0.0
28. CH(s) + Pt(s) → C(s) + H(s) 3.09 · 1022 0.0 0.0
29. C(s) + H(s) → CH(s) + Pt(s) 1.25 · 1022 0.0 138.0
IIIc CHx -decomposition
IIId C2 Hx -oxidation
30. C2 H3 (s) + O(s) → CH3 CO(s) + Pt(s) 3.70 · 1019 0.0 62.3
31. CH3 CO(s) + Pt(s) → C2 H3 (s) + O(s) 3.70 · 1021 0.0 196.7
θO(s) -45.0
32. CH3 (s) + CO(s) → CH3 CO(s) + Pt(s) 3.70 · 1021 0.0 92.9
33. CH3 CO(s) + Pt(s) → CH3 (s) + CO(s) 3.70 · 1021 0.0 0.0
θCO(s) -33.0
B.6 HC-SCR on Pt/Al2 O3 171

Reaction A / S0 β /µ Ea / 
IIIe CHx -oxidation
34. CH3 (s) + O(s) → CH2 (s) + OH(s) 3.70 · 1021 0.0 36.6
35. CH2 (s) + OH(s) → CH3 (s) + O(s) 3.70 · 1021 0.0 25.1
36. CH2 (s) + O(s) → CH(s) + OH(s) 3.70 · 1021 0.0 25.1
37. CH(s) + OH(s) → CH2 (s) + O(s) 3.70 · 1021 0.0 25.2
38. CH(s) + O(s) → C(s) + OH(s) 3.70 · 1021 0.0 25.1
39. C(s) + OH(s) → CH(s) + O(s) 3.70 · 1021 0.0 224.8
IIIf H, OH, H2 O reactions
40. O(s) + H(s) → OH(s) + Pt(s) 3.70 · 1021 0.0 11.5
41. OH(s) + Pt(s) → O(s) + H(s) 5.77 · 1022 0.0 74.9
42. H(s) + OH(s) → H2 O(s) + Pt(s) 3.70 · 1021 0.0 17.4
43. H2 O(s) + Pt(s) → H(s) + OH(s) 3.66 · 1021 0.0 73.6
44. OH(s) + OH(s) → H2 O(s) + O(s) 3.70 · 1021 0.0 48.2
45. H2 O(s) + O(s) → OH(s) + OH(s) 2.35 · 1020 0.0 41.0
IIIg CO-oxidation
46. CO(s) + O(s) → CO2 (s) + Pt(s) 1.00 · 1018 0.0 108.0
θCO(s) 33.0
θNO(s) -90.0
47. CO2 (s) + Pt(s) → CO(s) + O(s) 3.70 · 1021 0.0 155.1
θO(s) -45.0
48. C(s) + O(s) → CO(s) + Pt(s) 3.70 · 1021 0.0 0.0
θCO(s) -33.0
49. CO(s) + Pt(s) → C(s) + O(s) 3.70 · 1021 0.0 228.5
θO(s) -45.0
NO - reduction and oxidation
I Adsorption
50. NO + Pt(s) → NO(s) 8.50 · 10−1 0.0 0.0
51. NO2 + Pt(s) → NO2 (s) 9.00 · 10−1 0.0 0.0
52. N2 O + Pt(s) → N2 O(s) 2.50 · 10−2 0.0 0.0
II Desorption
53. NO(s) → NO + Pt(s) 1.00 · 1016 0.0 140.0
θO(s) 10.0
54. NO2 (s) → NO2 + Pt(s) 1.00 · 1013 0.0 60.0
55. N(s) + N(s) → N2 + Pt(s) + Pt(s) 3.70 · 1021 0.0 113.9
θCO(s) 75.0
56. N2 O(s) → N2 O + Pt(s) 1.00 · 1013 0.0 54.4
III NO-surface reactions
57. NO + O(s) → NO2 (s) 1.40 · 1010 0.5 97.5
θO(s) 45.0
58. NO2 (s) → NO + O(s) 1.00 · 1013 0.0 98.7
59. NO(s) + Pt(s) → N(s) + O(s) 4.00 · 1021 0.0 107.8
θCO(s) -3.0
60. N(s) + O(s) → NO(s) + Pt(s) 3.70 · 1021 0.0 128.1
θO(s) 45.0
61. NO(s) + N(s) → N2 O(s) + Pt(s) 1.00 · 1021 0.0 90.9
62. N2 O(s) + Pt(s) → NO(s) + N(s) 3.70 · 1021 0.0 66.9

Reaction 17. is a complex fast reaction that takes place if sufficient O(s) is available for Reaction 3. Reaction 17. is
first order in O(s) and C3 H5 (s).
Acknowledgments

First and foremost, I would like to thank Professor Jürgen Warnatz for his continuous
scientific and personal support of my work. He is a great mentor, who not only pointed
me in the right direction but also provided enough freedom for my own thoughts and
creativity.

The research described in this work is unthinkable without a strong research group,
with which I enjoyed working. In particular, I would like to thank Steffen Tischer
for programming the boundary-layer code, Dr. Chrys Correa for programming the
monolith code, and Daniel Chatterjee for modeling automotive catalytic converters.
Renate Schwiedernoch set up the reactor for experimental investigations of short-
contact-time catalysis in an impressively short time. I acknowledge Dr. Luba Maier’s
development of reaction mechanisms and Oliver Grosshans’ work on optimization.
For assistance in various software and hardware problems, I appreciate the help of
Stefan Kleditzsch, Volker Karbach, Tillman Katzenmeier, and Markus Nullmeier.
I also would like to thank my former colleagues Dr. Markus Wolf, Dr. Christian
Taut, and Dr. Ralf Kissel-Osterrieder. I thank Professor Frank Behrendt and Dr.
Uwe Riedel for their long-lasting colleagueship. Ingrid Hellwig and Barbara Werner
are acknowledged for their supportive secretarial work. I would also like to thank
all members of the research group Reactive Flows at the Interdisciplinary Center
for Scientific Computing at Heidelberg University for their collaboration and a very
conducive working environment.

Professor Jürgen Wolfrum, Dr. Hans-Robert Volpp, and Dr. Uwe Metka (PCI,
Heidelberg University) were very kind to provide the space and technical support for
the experimental study. I also like to thank them for a fruitful collaboration in the
Sonderforschungsbereich 359.

Professor Lanny D. Schmidt (University of Minnesota) introduced me with his


enthusiasm to the problems of short-contact-time reactors. I would like to thank
him and all his students for their hospitality and 15 wonderful months I spent as
Postdoc in Minnesota in 1997/98. In particular, I would like to thank Dr. Ashish
Bodke, Dr. Dimitri Iordanoglou, Prof. Keith Hohn, and Dr. Srinivas Tummala for
our collaboration on catalytic partial oxidation. The friendship I shared with them
and Jamease Kowlaczyck allowed me to survive the cold Minnesotan winter. I very
appreciate the still ongoing collaboration between our groups and thank Jeremy
Redenius and Ryan O’Connor for our collaboration on catalytic radiant burners and
partial oxidation of higher alkanes, respectively.
I have always been enjoying the meetings and discussions with Professor Robert J.
Kee (Colorado School of Mines), who gave me numerous valuable advice. I very
appreciate the collaboration in catalytic combustion research with Professor Kee,
Professor Laxminarayan L. Raja (Colorado School of Mines), and Professor Robert
W. Dibble (University of California at Berkeley).

The investigation of oxy-dehydrogenation of ethane is based on a very fruitful


collaboration with Dr. David K. Zerkle (Los Alamos National Laboratory). I would
also like to thank him for the nice summer of 1998 I spent in Los Alamos.

The experimental data for the studies on automotive catalytic converters were
provided by Professor W. Weissweiler, Erik Frank, and Dr. Sven Kureti (University
of Karlsruhe).

I thank Chrys, Steffen, Ryan, Jeremy, and Keith for proof-reading the manuscript.

The research presented in this work was funded by the Deutsche Forschungsge-
meinschaft (Postdoctoral fellowship, SFB 359, Sachbeihilfen), Forschungsverbund
Verbrennungskraftmaschinen, Bundesland Baden-Württemberg, US Department of
Energy, Eberspächer GmbH & Co, and Conoco Inc. I very much appreciate the
financial support.

I would like to thank my parents and my daughter Franziska for their love, and all my
friends for being there.

Você também pode gostar