Sie sind auf Seite 1von 16

Chemical Engineering Science 60 (2005) 42494264

www.elsevier.com/locate/ces
Regeneration of coked catalystsmodelling and verication of coke
burn-off in single particles and xed bed reactors
Christoph Kern, Andreas Jess

Department of Chemical Engineering, University Bayreuth, Universittsstr. 30, 95447 Bayreuth, Germany
Received 14 July 2004; received in revised form 20 December 2004; accepted 27 January 2005
Available online 18 April 2005
Abstract
The regeneration of a coked naphtha reforming catalyst (Pt/ReAl
2
O
3
) was studied by kinetic investigations on the effective rate of
coke burn-off. For temperatures of industrial relevance for the catalyst, i.e., below 550

C (deactivation), the coke burn-off within the


cylindrical particles (d
P
2 mm) is determined by the interplay of chemical reaction and pore diffusion; limitation by external mass
transfer can be excluded for T <750

C. Based on the parameters of the intrinsic kinetics and of the structure of the catalyst (porosity,
tortuosity), the regeneration process is modelled and discussed both on the level of a single particle and in a technical xed bed reactor.
The results of modelling are compared with data from lab-scale investigations (coke proles within the particles) and the performance
data of the regeneration in an industrial xed bed reactor (moving reaction zone); the agreement of calculation and measurement is in
both cases complete.
2005 Elsevier Ltd. All rights reserved.
Keywords: Regeneration; Coke burn-off; Moving reaction zone; Single particle; Fixed bed reactor
1. Introduction
In several renery and petrochemical processes the cat-
alyst deactivates by coke formation, e.g. during cracking
of heavy oil, hydrodesulphurization and reforming of naph-
tha into high octane gasoline. The catalyst must be regen-
erated by continuous or periodic coke combustion. In case
of a xed bed reactor, the regeneration procedure is con-
ducted periodically after a certain time of operation, i.e., the
reactor is shut down for burn-off. During this non-steady-
state, process precautions have to be taken to avoid exces-
sive high temperatures, e.g. in case of naphtha reforming the
PtAl
2
O
3
-catalyst, the latter loses surface and mechanical
resistance beyond temperatures of 550

C (LePage, 1978).
So the air is strongly diluted with nitrogen to avoid damage
to the catalyst.
A mathematical model of the decoking process would be
a helpful tool both with respect to a better understanding of

Corresponding author. Tel.: +49 921 55 7431; fax: +49 921 55 7435.
E-mail address: jess@uni-bayreuth.de (A. Jess).
0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.01.024
the processes of coke burn-off and to determine how to
perform the process both rapidly and safely. Such a model
should be based on experimental data of the intrinsic kinet-
ics. In addition, mass and heat transfer have to be consid-
ered both on the level of a particle as well as of a xed bed
reactor.
Regeneration and deactivation were already extensively
studied with respect to the nature of coke deposits, e.g. coke
formation and decoking on different catalytic sites (metallic
or acidic) by means of temperature programmed oxidation or
similar techniques (e.g. Beltramini et al., 1991; Parera, 1991;
Zhang et al., 1991; Fung et al., 1991). Although these studies
contribute to a better understanding of the basic principles
of decoking, they are not adequate to describe the actual
processes within a particle or a xed bed.
Regeneration of a coked particle involves both chem-
ical reaction and transport processes, since oxygen must
be transported by external mass transfer and pore diffu-
sion to the internal coked surface. As already outlined by
Weisz (Weisz and Goodwin, 1963, 1966), pore diffusion
strongly inuences the effective rate of burn-off, at least for
4250 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
particle diameters and temperatures relevant for industrial
(xed bed) processes (>1 mm, >400

C). So during the


unsteady process, radial gradients of the O
2
-concentration
and with proceeding burn-off also of the carbon content in a
particle are established (Ishida and Wen, 1968; Wen, 1968;
Froment and Bischoff, 1990).
Model calculations of decoking in a xed bed (Westerterp
et al., 1988) show that a moving reaction zone migrates
through the reactor, which leads to an overheating of the
catalyst, if the velocity of the zone is too fast. In practice, the
process is therefore often conducted too slowly, for fear of
damage to the expensive catalyst. Up to now, models of coke
burn-off in a xed bed reactor are often based on several
simplications, e.g. on the assumption that external mass
transfer completely controls the burn-off rate (Westerterp
et al., 1988).
In the present work, the regeneration of a commercial
naphtha reforming catalyst was taken as a representative
model system and was studied by systematic experimental
as well as by theoretical investigations both on the scale of
a single particle and of a technical reactor (details in Kern,
2003; Tang et al., 2004; Tang, 2004). This approach has (to
the our best knowledge) not been taken before in such an
integrated form: (1) The intrinsic kinetics of coke burn-off
as well as the role of external and internal mass transfer
were studied by lab-scale experiments. (2) The coke burn-
off within a single particle was modelled and compared with
radial coke proles measured after a certain stage of burn-
off. (3) Thereafter, the regeneration process within a techni-
cal reactor was modelled and studied for different boundary
conditions. (4) Finally, the results of reactor modelling were
compared with the performance data of a real technical re-
forming xed bed reactor, which were kindly provided by
the MIRO renery (Karlsruhe, Germany).
Some experimental and theoretical results with respect to
the inuence of chemical reaction rate, diffusion and pore
structure on regeneration within single particles were already
presented in detail elsewhere, thereby taking mainly pure
Al
2
O
3
as a model catalyst (Tang et al., 2004; Tang, 2004).
So the focus of this paper is the coke burn-off within a
xed bed.
2. Experimental methods
Before coke burn-off, the catalyst was deactivated up to a
coke content of 20 wt% by passing toluene or n-heptane (as
model hydrocarbons for naphtha) in N
2
/H
2
over the catalyst.
To determine the kinetic parameters of coke burn-off (re-
action order of oxygen, E
A
, k
m,0
; list of symbols and ab-
breviations at the end of this paper) isothermal experiments
were performed in two lab-scale reactors, thereby varying
the initial carbon load, temperature, and O
2
content (details
in Kern, 2003). A tubular reactor was used for an O
2
con-
tent of less than 2 vol% (2 bar, 350550

C, modied res-
idence time t with respect to the mass of fresh catalyst:
0.02 kg h/m
3
). For higher concentrations (210 vol%), a
Berty-type reactor (2 bar, 350500

C, t (related to feed):
0.2 kg h/m
3
) with an internal recycle ratio of about 20 was
used to avoid a temperature runaway, i.e., to ensure isother-
mal conditions.
Supplementary to the isothermal experiments, the kinetic
parameters were also determined from the ignition temper-
atures at O
2
contents up to 100 vol% (see Section 3.1).
In addition to these kinetic investigations the coked and
decoked catalyst was characterized with respect to structural
parameters, which are essential to evaluate the role of inter-
nal mass transfer in the effective burn-off rate:
The specic internal surface area, the porosity, and the
pore diameter of the coked catalyst were measured by N
2
-
adsorption (BET-method, Micromeritics Gemini 539),
The tortuosity t
P
includes a variety of aspects such as
altered diffusion path length, variations in pore diameter,
interconnecting pores, changing cross-sectional areas in
constrictions, and dead end pores. In this work the tor-
tuosity was determined by the pulse eld gradient (PFG)
NMR technique with n-heptane as probe molecule (Kern,
2003; Tang et al., 2004; Tang, 2004; Ren, 2003; Ren et
al., 2000). Thereby also the change of the tortuosity dur-
ing regeneration was determined.
The radial carbon distribution within single (coked and
partly regenerated) particles was measured by SEM/EDX
(JSM840A, Jeol/INCA, Oxford) at the Department of
Materials Processing (University Bayreuth). Therefore, it
must be noted, that the initial radial carbon distribution
was proven to be always homogeneous, which is the re-
sult of the very low reaction rate of deactivation by cok-
ing (for details see Kern, 2003; Tang et al., 2004; Tang,
2004).
Characteristic data of the reforming catalyst used for the
experiments are listed in Table 1.
3. Data evaluation
3.1. Intrinsic kinetics of coke burn-off
In this work, the coke was regarded as pure carbon, so the
small hydrogen content of the coke (4.5 wt%; molar C/H-
ratio of about 1.75, as determined by elementary analysis
(EA 3000, HEKAtech)) was not considered in the data eval-
uation, i.e., the kinetic data given in this paper are all re-
lated to the carbon of the carbonaceous deposits. During the
isothermal measurements, the initial and the actual amount
of carbon was calculated by planimetry fromthe gas analysis
(CO
2
, 05000 ppm) and the volume rate of the off-gas. (Re-
mark: CO was not formed during coke burn-off (<5 ppm),
which reects the activity of Pt for oxidation of CO.) The
carbon concentration within the catalyst is given here as the
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4251
Table 1
Characteristic data of the reforming catalyst
Designation (company) E-802 (Engelhard, USA)
Composition in wt%: Pt/Re/Cl (rest Al
2
O
3
) 0.26/0.50/1
Catalyst density, j
cat
(fresh catalyst) 1400 kg/m
3
Catalyst bulk density, j
B
770 kg/m
3
Geometry: length/diameter of extrudates 28 mm/1.6 mm
Specic internal BET-surface area A
BET
(fresh catalyst) 200 m
2
/g
carbon load L
C
, i.e., as the mass of carbon per mass of fresh
catalyst.
The intrinsic rate of regeneration is given by
r
m
=k
m,C
L
C
c
q
O
2
(1)
with
k
m,C
=k
m,C,0
e
E
A
/RT
. (2)
The experiments with varied oxygen content show, that the
reaction order with respect to the oxygen concentration is
about one (Kern, 2003). For the tubular reactor used here, the
rate constant (assuming plug ow) can then be calculated by
k
m,C
=
ln(1 X
O
2
)
tL
C
. (3)
The Berty-type reactor with a relative high internal recycle
ratio of about 20 can be regarded as an ideal stirred tank
reactor, and k
m,C
is then given by
k
m,C
=
1
L
C
X
O
2
t(1 X
O
2
)
. (4)
For an oxygen content higher than 10 vol%, the isother-
mal operation was hard to realize, and the reactivity of the
coke was determined by using the following non-isothermal
method: The feed gas (N
2
/O
2
-mixture) is passed through
an electrically heated quartz reactor with a xed bed of the
coked catalyst. Starting from room temperature, the sample
is heated with a rate of 10 K/min. At rst, the temperature
in the bed rises according to the heating rate of the oven
until a sudden and pronounced increase in temperature oc-
curs. This break in the temperature curve is considered to be
the ignition temperature. The rate constant is then deduced
from the ignition temperatures at different O
2
-contents (here
10100 vol%) according to the theory of thermal explosion
(heat generation equals dissipation). The underlying theory
and details of this so-called ignition point method are de-
scribed elsewhere (Kern, 2003; Herbig and Jess, 2002; Hein
and Jess, 2000; Herbig, 2002).
3.2. Determination of diffusional parameters
The inuence of pore diffusion is considered by the effec-
tiveness factor p
pore
, i.e., (in case of no inuence of external
diffusion) by the ratio of the effective (measured) rate con-
stant to the (maximum) intrinsic rate constant (d
P
0),
and is given for a rst-order reaction by (see e.g. Baerns
et al., 1987):
p
pore
=
k
m,C,eff
k
m,C
=
tanh [
[

1
[
for [2. (5)
The so-called Thiele modulus (for a coked particle with a
uniform carbon load) is given by
[ =V
m
/A
m
_
k
m,C
L
C
j
P
/D
O
2
,eff
, (6)
whereby V
m
/A
m
is the ratio of the particle volume to the
external particle surface area.
It should be noted that [ depends on the carbon load L
C
,
which changes during burn-off with time and in case of a
resistance of pore diffusion also with the radial position in
the particle. The Thiele-approach as given by Eqs. (5) and
(6) can then not be applied anymore, and numerical sim-
ulations are needed. Nevertheless, the initial effectiveness
factor p
pore,0
is used here (in addition to the results of nu-
merical simulations, where p
pore
and [ are not needed) as
a descriptive measure for the pore diffusion resistance, sim-
ply dened by Eqs. (5) and (6) for L
C
= L
C,0
. So strictly
speaking, [
0
and p
pore,0
are only valid at the start of the
regeneration.
To describe the effective oxygen diffusion within the
porous catalystexpressed by an effective diffusion coef-
cient D
O
2
,eff
it has to be considered that only a portion
of the particle is permeable, and that the path through the
particle is random and tortuous. Both aspects are taken into
account by the porosity c
P
and the tortuosity t
P
(Eq. (7)),
whereby both factors change with the coke content during
regeneration:
D
O
2
,eff
=
c
P
t
P
D
O
2
,pore
=
c
P
/t
P
_
1
D
O
2
,mol
+
1
D
O
2
,Knu
_. (7)
Depending on the pore diameter, the diffusivity in a pore
is the combined diffusivity of the molecular and Knudsen
diffusivity, the latter calculated by
D
O
2
,Knu
=
d
pore
3
_
8RT
M
O
2
. (8)
If also external diffusion has to be considered, Eq. (5) has to
be extended. The overall effectiveness factor p
overall
is then
4252 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
given by
p
overall
=
_
k
m,C
L
C
[A
m
+
1
p
pore
_
1
. (9)
The external mass transfer coefcient [ was calculated based
on the Sherwood-number (Sh =[d
P
/D
O
2
,mol
) as given by
Schlnder (1986).
4. Modelling methodology
4.1. Coke burn-off within a single particle
The radial proles of the oxygen and carbon load within a
single catalyst particle and the time needed to reach a certain
degree of coke burn-off were simulated by the commercial
computer program Presto (solver for differential equations,
CiT GmbH, Rastede, Germany, details in Wulkow et al.,
2001).
Because of the relative high thermal conductivity of the
solid phase, the simulation of decoking in a particle is only
based on the mass balances of O
2
and carbon. Assuming
that axial gradients of the O
2
-content and carbon load can
be neglected in the cylindrical particles (L
p
/d
P
4), the
following mass balances for the gas and solid phase arise:
c
P
oc
O
2
ot
=
d
r dr
_
D
O
2
,eff
r
oc
O
2
or
_
r
m
j
P
, (10)
oL
C
ot
=M
C
r
m
. (11)
The intrinsic rate of O
2
-conversion is
r
m
=
on
O
2
om
cat
=k
m
(T )L
C
c
O
2
. (12)
These differential equations are solved numerically for the
boundary conditions:
oc
O
2
/or =0 for r =0 (particle center) (13)
c
O
2
=c
O
2,G
, for r =r
p
(particle surface) (14)
4.2. Modelling of regeneration in a xed bed reactor
For the modelling of the coke burn-off in an adiabatic
technical xed bed reactor a so-called one-dimensional pseu-
dohomogeneous reactor model with axial mixing was used,
which can be characterized as follows:
Due to the high ratio of the reactor to particle diameter
in case of a technical xed bed (100), radial gradients
of the O
2
-content and temperature are neglected.
Axial dispersion of heat and mass transfer is taken into
account by the use of an effective dispersion and heat
conduction coefcient D
ax,eff
and z
ax,eff
(see Section 5.5).
The temperature difference between the solid and gas
phase can be estimated by
(T
G
T
S
) =A
R
Hj
P
r
m,eff
/:A
V
. (15)
The temperature difference calculated by Eq. (15) for
the maximum temperature technically relevant for reform-
ing (550

C), a high carbon load (10 g


C
/g
cat
and a super-
cial velocity of 0.5 m/s (:=140 W/(m
2
K); from Schlnder,
1986) is then 10 K. So T was neglected, i.e., no thermal
distinction was made between the gas and solid phase.
Based on the aforementioned assumptions the differential
equations for the mass and heat balance of the solid and the
gas phase are
j
B
M
C
oL
C
ot
=j
B
r
m,eff
, (16)
c
dc
O
2
dt
= u
G
c
dc
O
2
dz
+D
ax,eff
d
2
c
O
2
dz
2
j
B
r
m,eff
, (17)
(j
B
c
S
+c j
G
c
p,G
)
oT
ot
= u
G
cj
G
c
p,G
oT
G
oz
+z
ax,eff
o
2
T
G
oz
2
+A
R
H j
B
r
m,eff
. (18)
The effective dispersion coefcients of heat and mass
(z
ax,eff
, D
ax,eff
) in the xed bed were calculated based on
literature correlations (Verein Deutscher Ingenieure, 2002):
D
ax,eff
=D
f b
+
u
e
d
P
2
, (19)
z
ax,eff
=z
f b
+
u
e
j
G
c
p,G
d
P
2
. (20)
z
f b
and D
f b
represent the minimal values of z
ax,eff
and
D
ax,eff
without gas ow:
D
f b
=D
O
2
,mol
(1

1 c), (21)
z
f b
=z
G
K. (22)
Based on literature correlations (Verein Deutscher
Ingenieure, 2002), the factor K is given here for a voidage
of the xed bed of 0.4 by
K =
11k
P
+4
2k
P
+13
. (23)
(Remark: Eq. (23) is only valid for K <20.)
Therefore k
P
is the ratio of the thermal conductivity of
gas and solid phase. For a typical value of the thermal con-
ductivity z
S
of porous Pt/Al
2
O
3
particles of 0.2W/(mK)
(Baerns et al., 1987), k
P
can be estimated here by
k
P
=
z
S
z
G
(500

C)
=
0.2 (W/mK)
0.05 (W/mK)
=4. (24)
So for the given catalyst and reaction conditions, the value
of K (Eq. (23)) is here about 2.3.
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4253
The boundary conditions for the solution of differential
equations (16)(18) by the above-mentioned computer pro-
gram Presto are:
For z =0 (reactor inlet):
u
G
c(c
O
2
,0
c
O
2
) =cD
ax,eff
oc
O
2
oz
, (25)
u
G
c j
G
c
p,G
(T
0
T ) =z
ax,eff
oT
oz
(26)
and for z =L (reactor outlet):
oc
O
2
/oz =0, (27)
oT/oz =0. (28)
5. Results and discussion
5.1. Structural parameters of catalyst
As shown by Kern (2003), the inuence of the carbon load
on the porosity c
P
and tortuosity t
P
of the catalyst particle
is approximately given by the following linear relationships:
c
P
=0.65 1.3 L
C
, (29)
t
P
=2.59 5.4 L
C
. (30)
So for a typical initial load L
C
of 0.15 g
C
/g
cat
, the c
P
de-
creases and t
P
increases by 30% compared to the fresh cat-
alyst. (More details in Kern, 2003; Tang, 2004; Tang et al.,
2004.)
5.2. Intrinsic coke reactivity
Results on the intrinsic kinetics of deactivation and re-
generation of the catalyst were already presented elsewhere
(Kern, 2003; Jess and Kern, 2001; Jess et al., 1999; Ren et
al., 2002). The main results are:
Hydrogen strongly inhibits the rate of coke formation,
but has no inuence on the reactivity of the carbonaceous
deposits.
Compared to toluene, the coke formation rate is much
lower in case of heptane.
The reactivity of the coke does not depend on the condi-
tions of coke formation (H
2
-pressure, feedstock for coke
formation), although the time needed to reach a certain
carbon load is quite different.
Two sorts of coke are formed on the metal (Pt/Re) and
the acidic sites (Al
2
O
3
), respectively, whereby the latter
is much less reactive. In the beginning of decoking, the
small amount of metal coke (e.g. 2% of the total carbon
for L
C,0
= 0.15 g
C
/g
cat
) is rapidly burned off; for the
modelling of the decoking process, only the second dom-
inating (with respect to the amount) and less reactive type
of coke on alumina was considered.
0.0011 0.0012 0.0013 0.0014 0.0015 0.0016 0.0017
1/T in 1/K
k
m
,
C

i
n

m
3
/
(
k
g
.

s
)
Ignition point method
Tubular reactor
Berty type reactor
10
1
10
0
10
-1
10
-2
10
-3
10
-4
10
-5
400 500 C
Fig. 1. Intrinsic kinetics of coke-burn-off.
0.01
0.10
1.00
0.0006 0.0012 0.0018
Reciprocal temperature in 1/K
E
f
f
e
c
t
i
v
e
n
e
s
s

f
a
c
t
o
r
L
C,0
= 14.5 g
C
/100 g
Kat
Technically relevant
temperature range

pore,0

overall
C
550 750
300 1000
Fig. 2. Inuence of temperature on the pore and overall effectiveness
factor: comparison of measurement and calculation (u
e
= 0.2 m/s (at
500

C), p =1 bar, y
O
2
=0.5 vol%, c
P
=0.46, t
P
=1.7, d
pore
=10 nm,
t =5 (kg s)/m; line: calculation).
The intrinsic rate of O
2
-conversion (of the less reactive
coke) is given in the whole range from 1% to 100% oxy-
gen by (Fig. 1):
r
m
=
on
O
2
om
cat
=k
m,C
(T )L
C
c
O
2
(31)
with
k
m,C
(T ) =1.6 10
6
(m
3
/kg s) e
107,000/RT
.
The reactivity of the coke is practically independent of
the carbon load (if the small amount on the metal sites is
neglected).
5.3. Inuence of internal mass transfer
Fig. 2 shows the result of an experiment with an initial car-
bon load of 0.145 g
C
/g
cat
. The bed was heated from 300

C
up to 600

C with a small rate (0.51

C/ min), so stationary
conditions can be assumed. Because of the small O
2
-content
4254 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
Table 2
Parameters for the modelling of coke burn-off within a single catalyst
particle
Frequency factor, k
m,C,0
1.6 10
6
m
3
/(s kg)
Activation energy, E
A
107,000 J/mol
Density of the catalyst, j
cat
1400 kg/m
3
Porosity, c

P
0.53
Tortuosity, t

P
3.10
Average pore diameter, d
pore
10 nm
Effective diffusion coefcient, D
O
2
,eff
a
4 10
7
m
2
/s
a
Typical values for a coke load of 10 g
C
/100 g
cat
and in case of
D
O
2
,eff
for a temperature of 500

C and 1 bar.
(0.5 vol%) the carbon load of the catalyst during the exper-
iment was considered to be constant and equivalent to the
initial value. Fig. 2 indicates that the (overall) effectiveness
factor calculated by Eq. (9) (based on the Thiele modulus
[
0
for L
C
=L
C,0
) and the measured data are consistent. It
should be noted here, that the pore diameter was used as the
tting parameter in the calculation. If the pore diameter is
set to 10 nm, a good agreement with the experimental data
is obtained (see Fig. 2), which is well in the pore diameter
distribution (225 nm; see Kern, 2003).
For temperatures below 750

C, the overall effectiveness


factor almost equals the one with respect to pore diffusion,
i.e., p
overall
0.9p
pore
, which clearly indicates that the inu-
ence of external mass transport can be neglected, at least for
the particles used here and technically relevant temperatures
(<550

C with respect to deactivation).


5.4. Regeneration of a single coked particles
5.4.1. Numerical solutions
Based on Eqs. (10)(14), the decoking process within
a single particle was numerically simulated. The essential
parameters of the modelling are given in Table 2.
The calculated proles of the oxygen concentration and
the carbon load within a single particle during the regener-
ation process are shown in Fig. 3 for two typical burn-off
temperatures (450 and 550

C). In addition, the initial effec-


tiveness factor p
0
(calculated with the initial Thiele modulus
[
0
for L
C
= L
C,0
) as a measure for the magnitude of the
pore diffusion resistance is given.
As expected, the inuence of pore diffusion on the rate of
coke combustion increases with temperature, and gradients
of the oxygen concentration and the carbon load occur. This
effect is even more pronounced in case of a very high tem-
perature of 650

C (right side of Fig. 4). Therefore it must


be noted that 650

C is unrealistic for a technical regenera-


tion (T
max
of reformer catalyst is about 550

C), but is also


shown here as an instructive comparison. A relative sharp
reaction front within the particle is then formed, which mi-
grates from the outer surface to the centre of the particle.
In case of a very low temperature of 350

C (Fig. 4, left),
diffusion limitations disappear, but the regeneration time is
then very long (>4 days) compared to about 8 h in case of
a technically realistic temperature of 450

C.
The strong inuence of pore diffusion with increasing
temperature is also reected by the (initial) effectiveness
factor p
0
: At 350

C, p
0
is 98%, whereas for 450, 550 and
650

C values of 79%, 39%, and 17% are reached.


Finally, the coked catalyst was regenerated at different
temperatures up to a burn-off degree of about 50%. The
radial carbon proles in the decoked particles were measured
by SEM/EDX. In Fig. 5, two selected proles and also the
experimental times to reach 50% burn-off are compared with
the numerically calculated data. For the simulation (550 and
700

C), the values of c


P
and t
P
of the entirely decoked
catalyst were used, as a pronounced practically carbon-free
shell was formed, which dominates the diffusion resistance.
The calculated and measured values both of the C-proles
and the time needed for regeneration are in good agreement.
The scattering of the experimental data is evoked from the
fact that three C-proles were measured for each sample.
The pronounced development of a C-free outer shell for a
temperature of 700

C is also reected by the photograph of


the cross section of the particles. The white shell obviously
indicates the outer C-free zone.
The thickness of the C-free zone (as determined by
the photographs) is about 0.15 mm (700

C) and less than


0.1 mm (550

C, not shown, see Tang, 2004), which is also


in good agreement with the modelled values of 0.15 mm
(r/r
0
=0.81 for 700

C) and about zero for 550

C ( Fig. 5).
Fig. 6 depicts that the burn-off time, dened here as the
time to reach 90% coke burn off, is constant up to a certain
initial carbon load (1 g per 100 g of (fresh) catalyst for the
given example of a temperature of 450

C), and then strongly


increases, which reects the increasing inuence of pore
diffusion with increasing carbon load.
5.4.2. Analytical (closed) solution
A description of the coke burn-off process within a single
particle by well-known closed solutions like the homoge-
neous model or the shrinking core model is only reasonable
for the border cases of complete control by chemical reac-
tion or by pore diffusion, respectively.
In order to describe quantitatively the burn-off process
without the need of an elaborated numerical solution (see
Section 5.4.1), an advanced closed solution was developed.
This model is slightly more complicated than the two afore-
said border cases, and includes the inuence of the pore
diffusion as well as the inuence of the intrinsic kinetics
on the effective rate of coke burn-off. The objectives to de-
rive such a closed solution in addition to the results of the
numerical calculations were as follows: (1) The numerical
solutions can be proved and veried, at least for selected
border cases. (2) If the closed solution ts quite well the
numerical correct solution, the former can be taken as the
basis and input for the modelling of the regeneration of a
whole xed bed reactor. By this means, the closed solution
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4255
0
20
40
60
80
100
L
C

/

L
C
,
0

i
n

%
C
O
2

/

C
O
2
,
0

i
n

%
relative particle radius r/r
0
0
20
40
60
80
100
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
relative particle radius r/r
0
55 min
115min
200 min
325 min
440 min
55 min
115 min
200 min
325 min
440 min
20 min
40 min
135 min
210 min
20 min
40 min
75 min
135 min
210 min
T = 550 C;
pore,0
= 39 % T = 450 C;
pore,0
= 79 %
75min
Fig. 3. Radial proles of O
2
-content (relative to the gas phase) and carbon content (relative to the initial value) in a single cylindrical particle at different
times (modelled results for L
C,0
=10 wt%, y
O
2
=2 vol%, d
P
=1.6 mm, p =1 bar, d
pore
=10 nm, c
P
and t
P
are calculated by Eqs. (29) and (30)).
0 0.2 0.4 0.6 0.8 1
10 min
30 min
65 min
125 min
200 min
125 min
200 min
relative particle radius r/r
0
0 0.2 0.4 0.6 0.8 1
relative particle radius r/r
0
T = 650 C;
pore,0
= 17 %
0
20
40
60
80
100
,
0
20
40
60
80
100
500 min
1100 min
2000 min 6300 min
3400 min
500 min
1100 min
2000 min
3400 min
6300 min
T = 350 C;
pore,0
= 98 %
65 min
30 min
10 min
L
C

/

L
C
,
0

i
n

%
C
O
2

/

C
O
2
,
0

i
n

%
Fig. 4. Radial proles of O
2
-content (relative to the gas phase) and carbon content (relative to the initial value) in a single cylindrical particle at different
times (modelled results for L
C,0
=10 wt%, y
O
2
=2 vol%, d
P
=1.6 mm, p =1 bar, d
pore
=10 nm, c
P
and t
P
calculated by Eqs. (29) and (30)).
4256 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
0.0 0.2 0.4 0.6 0.8 1.0
0
20
40
60
80
100
Relative particle radius r /r
i 0
T = 550 C
B
C,0
= 0.21 g
C
/g
cat
y
O
2
= 1 vol.-%
P = 1.1 bar
Burn-off degree: 50 %
t
Exp
= 2.2 hr
Model (d
pore
= 20 nm, t
model
= 2.2 h)
0
20
40
60
80
100
0.15 mm
T = 700 C
B
0
= 0.21 g
C
/g
cat
y
O
2
= 1 vol.-%
P = 1.2 bar
Burn-off degree: 50 %
t
Exp
= 1.7 h
Model (d
pore
= 20 nm, t
model
= 1.6 h)
1 0 3 2 5 4 7 6 8
mm
The thickness of the carbon -free outer
shell is about 0.15 mm
T = 700 C
L
C,0
R
e
l
a
t
i
v
e

c
o
n
t
e
n
t

o
f

c
o
k
e

L
C

/

L
C
,
0

[
%
]
L
C,0
Fig. 5. Comparison of measured and calculated radial coke proles in a partly regenerated PtAl
2
O
3
particle (regeneration conditions: xed bed,
y
O
2
=1 vol%, p =11.2 bar, PtAl
2
O
3
-cylinders: d
P
=1.6 mm, L
p
=28 mm; c
P
=0.5, t
P
=3.5).
0
200
400
600
800
1000
0.1 1 10 100
Initial carbon load L
C,0
in g
C
/100g
Kat
B
u
r
n
-
o
f
f

t
i
m
e

i
n

m
i
n

(
X
C

=

9
0

%
)
T = 450 C
no influence of pore
diffusion
Fig. 6. Inuence of initial carbon load on the regeneration time (up
to a burn-off degree of 90%) for a temperature of 450

C (conditions:
y
O
2
= 2 vol%, d
P
= 1.6 mm, p = 1 bar, d
pore
= 10 nm, c
P
and t
P
are
calculated by Eqs. (29) and (30)).
for the level of a particle simplies the numerical solution
of the coke burn-off in a technical xed bed reactor (see be-
low Section 5.5), and thus bridges the coke burn-off within
a particle and the regeneration on the macroscopic scale of
a technical reactor.
The closed solution, which was nally selected (for details
see Kern, 2003) is the so-called shrinking core model with
inuence of the chemical reaction (subsequently denoted
as the combined model) and is based on the assumption
of two consecutive steps (Fig. 7):
1. O
2
-diffusion through an entirely regenerated shell of the
particle ranging from the outer surface up to a dened
reaction front at r =r
RF
:
n
O
2
=2r
RF
L
p
D
O
2
,eff
_
oc
O
2
or
_
r=r
RF
. (32)
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4257
r
r
RF
P
L
C
= L
C,0
L
C
= 0 Reaction-
front (RF)
c
O
2
,G
c
O
2
c
O
2
, RF
r
L
Fig. 7. Combined model (shrinking core with inuence of chemical
reaction).
2. Chemical reaction without pore diffusion resistance in
the remaining core (0 <r <r
RF
) with a constant carbon
load of L
C,0
:
n
O
2
=r
2
RF
L
p
k
m,C
L
C,0
c
O
2
,RF
j
cat
. (33)
So according to this model, the inuence of pore diffusion
is restricted to a carbon-free shell (Eq. (32)), whereas in
reality (at least for medium temperatures and not too low
degrees of burn-off) carbon is present in this outer zone. This
leads to an underestimation of the carbon conversion by the
model. To the contrary, the assumption that L
C
in the core
(Eq. (33)) is still equivalent to the initial value overestimates
the burn-off rate compared to reality, where both L
C
and
c
O
2
decrease in the core of the particle. As shown below,
these two effects compensate each other quite well.
The concentration gradient at r = r
RF
can be calculated
based on the stationary mass balance for the carbon free
shell:
0 =
o
2
c
O
2
or
+
1
r
oc
O
2
dr
. (34)
The boundary conditions are
c
O
2
=c
O
2
,RF
for 0 <r r
RF
, (35)
c
O
2
=c
O
2
,G
for r =r
p
. (36)
(The latter condition implies that the external mass transfer
resistance can be neglected.)
Eq. (34) then yields:
_
oc
O
2
or
_
r=r
RF
=
(c
O
2
,G
c
O
2
,RF
)
ln(r
RF
) ln(r
p
)
1
r
RF
. (37)
Based on Eqs. (35), (32) and (33), the oxygen concentration
at r =r
RF
(relative to the one in the gas phase) is given by
c
O
2
,RF
c
O
2
,G
=
_
1
r
2
RF
k
m,C
L
C,0
j
cat
2D
O
2
,eff
ln
_
r
RF
r
p
_
_
1
. (38)
The correlation between the amount of coke converted per
unit time and the position of the reaction front r
RF
can be de-
duced from the following considerations: (1) The converted
amount of coke corresponds to the oxygen diffusing into the
particle. (2) V
C
represents the volume of the core, where
coke is still present at the initial level L
C,0
. This leads to
on
C
ot
= n
O
2
=
j
cat
L
C,0
M
C
oV
C
ot
=
2r
RF
L
p
j
cat
L
C,0
M
C
or
RF
ot
. (39)
Based on Eqs. (38) and (39), the velocity of the reaction
front and the change of the average coke load are then given
by
or
RF
ot
=
0.5M
C
c
O
2
,G
1
k
m,C
r
RF

r
RF
L
C,0
j
cat
D
O
2
,eff
ln
_
r
RF
r
p
_, (40)
o L
C
ot
=2L
C,0
r
ZG
r
2
p
or
ZG
ot
. (41)
Integration of Eq. (41) leads to:
L
C
=L
C,0
r
2
RF
/r
2
p
, (42)
r
RF
=r
2
p
_
L
C
/L
C,0
. (43)
Finally, Eqs. (42), (43), and (40) lead to the rate of change
of the average carbon load:
oL
C
ot
=
M
C
c
O
2
,G
1
k
m,C
L
C

r
2
p
j
cat
4D
O
2
,eff
ln
_
L
C
L
C,0
_. (44)
Integration of Eq. (44) leads to
t =
1
C
1
(C
2
L
C
ln(L
C
/L
C,0
)
+ln(L
C
/L
C,0
) +C
2
ln(L
C
/L
C,0
)) (45)
with
C
1
=M
C
k
m,C
c
O
2
,G
, (46)
C
2
=
r
2
p
k
m,C
j
P
4D
O
2
,eff
. (47)
So Eq. (45) can be used to calculate the burn-off degree
X
C
(=1 L
C
/L
C,0
).
It should be mentioned that two classical border cases can
be derived from Eq. (44):
4258 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
0
20
40
60
80
100
0 200 400 600 800
Regeneration time in min

B
u
r
n
-
o
f
f

d
e
g
r
e
e

X
C

i
n

%
numerical solution
combined model
650 C
450 C
400 C
350 C
500 C
Fig. 8. Comparison of the numerical solution and the closed solution
(combined model) (conditions: L
C,0
= 10 g
C
/100 g
cat
; c
P
/t
P
= 0, 17;
p =1 bar; y
O
2
=2 vol%).
1. At low temperatures (r
2
p
k
m,C
j
P
/D
O
2
,eff
1), Eq. (44)
reduces to the homogeneous model:
o L
C
ot
=M
C
k
m,C
c
O
2
,G
L
C
. (48)
In this case, the effective reaction rate of the coke combus-
tion is only determined by the chemical reaction without any
inuence of pore diffusion.
2. At high temperatures (r
2
p
k
m,C
j
cat
/D
O
2
,eff
1), Eq. (44)
is reduced to
o L
C
ot
=
4D
O
2
,eff
M
C
c
O
2
,G
r
2
p
j
cat
ln(L
C
/L
C,0
)
. (49)
In this case, the effective reaction rate is entirely controlled
by the O
2
-diffusion through the carbon-free shell of the par-
ticle. This classical so-called shrinking core model (without
inuence of chemical reaction) represents the case that the
reaction is conned to a front. In contrast to the shrinking
core model with inuence of chemical reaction (combined
model) used here, the O
2
-content gets zero on that front, so
that no reaction occurs within the coked core of the particle.
Fig. 8 shows the inuence of the time on the burn-off de-
gree at different temperatures. The agreement of the exact
numerical solution and the approximation by the closed so-
lution (Eq. (44), (combined model)) is very good.
The effective reaction rate r
m,eff
according to the com-
bined model is given by
r
m,eff
=
o L
C
M
C
ot
=k
m,C,eff
L
C
c
O
2
,G
. (50)
Insertion of Eq. (44) into Eq. (50) yields
r
m,eff
=
c
O
2
,G
1
k
m,C
L
C

r
2
p
j
cat
4D
O
2
,eff
ln
_
L
C
L
C,0
_. (51)
This expression for the effective reaction rate was used for
the modelling of the coke burn-off in a xed bed reactor (see
Table 3
Parameters of the modelling of coke burn-off in a xed bed reactor
(assumed to be constant)
Particle diameter, d
P
1.6 mm
Bulk density of xed bed, j
B
770 kg/m
3
Porosity of xed bed, c 0.4
Heat capacity of solid phase, 1000 J/(kg K)
c
S
(500

C)
Heat capacity of gas phase, 30 J/(mol K)
c
p,G
(500

C, 1 bar)
Density of gas phase, 16 mol/m
3
j
G
(480

C, 1 bar)
Effective axial heat conductivity 0.54W/(mK)
of xed bed, z
ax,eff
a
Effective axial diffusion coefcient 7.7 10
5
m
2
/s
of xed bed, D
ax,eff
a
a
The effective axial heat conductivity and the effective axial diffusion
coefcient of the xed bed are calculated by correlations published in
Verein Deutscher Ingenieure (2002) (see also Section 4.2).
Section 5.5) and builds the link between the microscopic
and the macroscopic scale of the regeneration process. (Note
that if also the external diffusion would play a role, Eq. (51)
could be extended by simply adding the term 1/([A
m
) in
the denominator of Eq. (51).)
5.5. Regeneration of a coked xed bed
5.5.1. Numerical solutions
Based on the numerical solution of the Eqs. (15)(28) and
(51) (parameters in Table 3), the regeneration of a coked
xed bed was modelled. A typical result within a technical
xed bed reactor of industrial scale (8 m length) is shown in
Figs. 9 and 10 (for details of the rst 6 h). After an induction
period of about one day, a reaction front with a constant
velocity is developed. This velocity can also be deduced by
a mass balance:
Within a time interval t , the reaction front moves in axial
direction by z. The amount of oxygen, which enters the
volume element with the length z within the time interval
t is converted by reaction with coke and thereafter needed
to ll up the void of the bed:
_
u
G
c c
O
2,in
_
t =
_
j
B
L
C,0

M
C
+c c
O
2,in
_
z. (52)
The velocity of the reaction front u
RF
(=z/t ) is then
given by
u
RF
=
u
G
c c
O
2,in
j
B
L
C,0
M
C
+c c
O
2,in

u
G
c c
O
2,in
M
C
j
B
L
C,0
. (53)
(u
G
c represents the supercial velocity u
e
.)
Within the initial period of regeneration, also a heat front
moves through the bed, which heats up the reactor zone
ahead of the moving reaction front from the initial temper-
ature T
0
to the maximum temperature T
max
. The velocity of
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4259
0
0.02
0.04
0.06
0.08
0.1
C
o
k
e

l
o
a
d

o
f

t
h
e

c
a
t
a
l
y
s
t

i
n

k
g
C
/
k
g
c
a
t

0
0.04
0.08
0.12
0.16
O
x
y
g
e
n

c
o
n
c
e
n
t
r
a
t
i
o
n

i
n

m
o
l
/
m
3

400
450
500
550
0 2 4 6 8
Length of reactor in m
T
e
m
p
e
r
a
t
u
r
e

i
n

C
Regeneration time:
1 1.3 h
2 6 h
3 14 h
4 30 h
5 56 h
6 111 h
7 167 h
8 195 h
End of the regeneration
after approx. 200 h
Oxygen breakthrough
after approx. 170 h
1
7
2
3
4
5 6
7
8
1
2
3
4
5
6
7 8
6
3
4
5
8
u
RF
u
RF

u
RF
= 4.7 cm/h
Fig. 9. Modelled proles of carbon load, oxygen-concentration and temperature in a xed bed reactor at different stages of regeneration (conditions: see
Tables 2 and 3; L
C,0
=10 g
C
/100 g
cat
; p =1 bar; y
O
2
=1 vol%; u
e
=0.5 m/s).
the heat front (here about 1.1 m/h) is much higher than the
one of the reaction front (0.05 m/h), and can be calculated
analogous to the reaction front based on a respective heat
balance by
u
HF
=
u
G
c j
G
c
p,G
j
G
c
p,G
+j
B
c
S

u
G
c j
G
c
p,G
j
B
c
S
. (54)
The inuence of the effective axial heat dispersion z
ax,eff
on the spread of the heat front is shown in Fig. 11. The
width of the heat zone is only clearly increased in case of an
(unrealistic) increase of the heat dispersion coefcient z
ax,eff
by a factor of 10 compared to the correct value of about
0.43W/(m K) as calculated by Eq. (20). If the axial heat
dispersion is completely neglected in the model (z
ax,eff
=0),
the width of the reaction zone gets slightly smaller. In other
words, the accuracy of the value of the axial heat dispersion
coefcient calculated by Eq. (20) does not play a role, as it
can certainly be assumed, that the error of the calculation of
z
ax,eff
by Eq. (20) is much smaller than a factor of 10. Note
that even in case of no axial heat dispersion (z
ax,eff
= 0),
the temperature decrease in the heat exchange zone is not a
step function, above all in the rear part (Fig. 11). The reason
is that the input signal (T -increase in the reaction front) is
not an ideal step function.
The inuence of the axial dispersion coefcient of mass
D
ax,eff
on the spread of the reaction zone is even smaller, as
shown in Fig. 12 (see also Kern, 2003)). Even a modelling
with an unrealistic high value of the effective dispersion
coefcient (factor 1000 higher compared to the correct
value as calculated by Eq. (20)) does not lead to a signi-
cant enlargement of the reaction zone. (Remark: If the ax-
ial dispersion of mass is completely neglected in the model
(D
ax,eff
= 0), the width of the reaction zone is the same
as in the case of modelling with the correct value, and
4260 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
Regeneration
time:
1 6 min
2 17 min
3 80 min
4 213 min
5 360 min
400
450
500
550
0 2 4 6 8
Length of reactor in m
T
e
m
p
e
r
a
t
u
r
e

i
n

C
1
2
5
Heat front past 213 min
Temperature
rise
3
4
u
HF
= 112 cm/h
Fig. 10. Modelled temperature proles in a xed bed reactor at the beginning of regeneration (conditions: see Tables 2 and 3; L
C,0
= 10 g
C
/100 g
cat
;
p =1 bar; y
O
2
=1 vol%; u
e
=0.5 m/s).
400
450
500
550
0 2 4 6 8
Length of reactor in m
T
e
m
p
e
r
a
t
u
r
e

i
n

C
u
HF
= 112 cm/h
) K m ( W 0
eff , ax
=
) K m ( W 3 . 4
eff , ax
=
) K m ( W 43 . 0
eff , ax
=
(correct value according to Eq. (20))
Fig. 11. Inuence of z
ax,eff
on the modelled temperature proles of the heat and reaction front (conditions: see Tables 2 and 3; L
C,0
=10 g
C
/100 g
cat
;
p =1 bar; y
O
2
=1 vol%; u
e
=0.5 m/s).
0
0.04
0.08
0.12
0.16
3 4 5 6
Length of reactor in m
O
x
y
g
e
n

c
o
n
c
e
n
t
r
a
t
i
o
n
i
n

m
o
l
/
m
3
s
m
10 73 . 7 D
2
2
eff , ax

.
=
t
mod
= 111 h
(correct value according to Eq. (19))
s
m
10 73 . 7 D
2
5
eff , ax

.
=
Fig. 12. Inuence of effective axial dispersion on the spread of the reaction
zone in a xed bed reactor during regeneration (conditions: see Tables 2
and 3; L
C,0
=10 g
C
/100 g
cat
; p =1 bar; y
O
2
=1 vol%; u
e
=0.5 m/s).
therefore not shown in Fig. 12). So the inuence of the axial
dispersion of mass can be neglected here, because the width
and velocity of the reaction zone, which determine the re-
generation time, are practically independent of the value of
D
ax,eff
. Nevertheless, the modelling was always done with
the correct values for D
ax,eff
(Eq. (19)) and also of z
ax,eff
(Eq. (20)).
In contrast to the small inuence of axial dispersion, the
pronounced inuence of the resistance of pore diffusion on
the width of the reaction front is shown in Fig. 13. In case
of an (unrealistic) small pore diameter of only 1 nm, which
is 10 times lower than the real mean value of the catalyst
used here, the diffusion coefcient is much lower (strong in-
uence of Knudsen diffusion), and the width of the reaction
zone is about 4 m.
With respect to determine the total time needed for coke
burn-off, it should be noted, that about 200 h are needed
compared to the minimum time in case for ideal step func-
tions of carbon load and O
2
-content (innite high reaction
rate). In the latter hypothetical case only 170 h (L
R
/u
RF
=
8 m/4.7 cm/h) are needed. In other words, the regeneration
time is 15% longer than theoretically needed in the absence
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4261
0
0.02
0.04
0.06
0.08
0.1
0 2 4 6 8
Length of reactor in m
C
o
k
e

l
o
a
d

i
n

k
g
C
/
k
g
c
a
t

nm 100 d
pore
=
nm 1 d
pore
=
nm 10 d
pore
=
t
mod
= 85 h
(Low influence of
pore diffusion)
(High influence of
pore diffusion)
Fig. 13. Inuence of pore diffusion on the spread of the reaction zone in
a xed bed reactor during regeneration (conditions: see Tables 2 and 3;
L
C,0
=10 g
C
/100 g
cat
; p =1 bar; y
O
2
=1 vol%; u
e
=0.5 m/s).
400
450
500
550
600
650
0 2 4 6 8
Length of reactor in m
T
e
m
p
e
r
a
t
u
r
e

i
n

C
9 cm/h
3.6 cm/h
0.9 cm/h
t
mod
= 25 h
T
ad
L
C,0
= 0.01 kg
C
/kg
cat
L
C,0
= 0.025 kg
C
/kg
cat
L
C,0
= 0.10 kg
C
/kg
cat
Fig. 14. Inuence of the initial carbon load on the temperature proles
of the heat and reaction front (conditions: see Tables 2 and 3; p =1 bar;
y
O
2
=1 vol%; u
e
=0.1 m/s).
of kinetic limitations, which underlines the need and advan-
tage of accurate modelling.
5.5.2. Temperature effects of regeneration
The adiabatic temperature increase (for a steady-state pro-
cess) is given by
T
ad
=c
O
2,in
A
R
H/j
G
c
p,G
. (55)
In case of the unsteady regeneration process, an unexpected
overheating above this adiabatic end temperature can oc-
cur (Fig. 14). The higher the velocity of the reaction front,
i.e., the lower the carbon load of the bed (see Eq. (53)), the
higher is this unwanted overheating effect (wrong way of be-
haviour). This effect is already described in literature (Wicke
and Vortmeyer, 1959; Emig et al., 1980; Eigenberger, 1983).
To quantify this wrong way of behaviour, the heat balance
of the reaction zone is instructive, whereby the originator of
the balance moves forward with the velocity of the reaction
zone:
Q
R
=Q
G
Q
S
. (56)
The heat ux produced by the coke burn-off is
Q
R
=(u
G
u
RF
)cc
O
2,in
|A
R
H|
=(u
G
u
RF
)c T
ad
j
G
c
p,G
. (57)
The heat ux needed to heat up the gas phase from the inlet
temperature T
0
up to the nal end temperature T
max
is given
by
Q
G
=(u
G
u
RF
)cj
G
c
p,G
(T
max
T
0
). (58)
The heat ux, which enters the reaction zone (from the
viewpoint of the moving observer) by the already heated
solid is given by
Q
S
=u
RF
cj
B
c
S
(T
max
T
0
). (59)
The Eqs. (53), (55)(59) then lead to
T
max
T
ad
=
u
HF
u
RF
cj
G
c
p,G
(cj
G
c
p,G
j
B
c
S
)
u
HF
u
RF
. (60)
So the higher the velocity of the reaction front (the smaller
the difference between u
HF
and u
RF
), the higher is the
overheating of the bed, as it is clearly shown in Fig. 14.
Such a wrong way of behaviour may be also induced
by a non-uniform initial axial carbon load (Fig. 15), here
deliberately calculated for a strong decrease of the carbon
load within the second third of the bed. At rst, i.e., in
the region with a high load of 0.1 g per g catalyst, the adi-
abatic stationary temperature increase is almost reached
(T
max
T
0
= 1.015T
ad
). As soon as the carbon load de-
creases, an overheating of the bed (wrong way behaviour) is
induced, and the temperature increases above the maximum
allowable value of 550

C (catalyst deactivation). Thus, it
must be noted that this scenario is unrealistic in case of naph-
tha reforming, where carbon formationif at allincreases
in axial direction due to the formation of coke precursors.
Nevertheless, Fig. 15 is of theoretical interest and may be
instructive for other decoking processes.
5.5.3. Coke burn-off in a technical reactor
Performance data of the regeneration process of a tech-
nical naphtha reformer were kindly provided by the MIRO-
renery (Karlsruhe, Germany). So the burn-off model could
be nally tested and compared with the regeneration in a
technical xed bed reactor. The respective results are given
in Fig. 16, indicating that the agreement is complete. It must
be noted that this agreement was reached although (1) not
the same, but a similar catalyst with the same geometry is
used in the MIRO renery. (2) The technical coke burn-off
is done at 20 bar, whereas the kinetic parameters were de-
duced from experiments at about 2 bar. (3) The C/H-ratio of
the catalyst of the technical unit is unknown, so the value
obtained in this work was used (1.75 mol/mol) to account for
the fact that hydrogen also contributes about 28% to the heat
generation (for simplication neglected in the modelling re-
sults shown before).
4262 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
0
0.02
0.04
0.06
0.08
0.1
C
o
k
e

l
o
a
d
i
n

k
g
C
/
k
g
c
a
t
400
450
500
550
600
650
0 2 4 6 8
Length of reactor in m
T
e
m
p
e
r
a
t
u
r
e

i
n

C
Regeneration
time:
1 0.4 min
2 12 min
3 70 min
4 110 min
5 118 min
6 125 min
7 140 min
1
2
3
4
5
6
7
1
2 3 4 5 6
7
T
ad
Fig. 15. Modelled proles of carbon load and temperature in a xed bed reactor with non-uniform initial axial carbon distribution (conditions: see Tables
2 and 3; p =1 bar; y
O
2
=1 vol%; u
e
=0.25 m/s).
350
400
450
500
550
0 2000 4000 6000 8000 10000
Length of reactor in mm
T
e
m
p
e
r
a
t
u
r
e

i
n

C
Regeneration
time:
1 7.5 h
2 19 h
3 27 h
1 2 3
T
ad
= 137 K
Length of fixed bed (6700 mm)
modelled
profil
measured
profil
Fig. 16. Comparison of measured and modelled temperature proles in a technical xed bed reactor (MIRO-renery, Karlsruhe, Germany) during
regeneration (conditions: see Tables 2 and 3; p =20 bar; y
O
2
=0, 9 vol%; u
e
=0.26 m/s; L
C,0
=21 g
C
/100 g
cat
).
Obviously the reactivity of carbonaceous deposits in tech-
nical renery processes and in lab-scale experiments is quite
similar, which is also approved by experiments with pure
Al
2
O
3
(Tang, 2004; Tang et al., 2004).
In order to show the possible deviation of the reaction
rate parameters in this work and the real kinetic param-
eters of the coke under technical conditions, Fig. 17 nally
shows the calculated temperature proles in the case, that
the real values of the intrinsic rate constant (MIRO re-
nery) would deviate from the calculation (based on our
lab-scale experiments with a different catalyst) by a factor
of two and 0.5, respectively. According to this sensitivity
analysis, the real intrinsic reactivity seems to be slightly
lower (better agreement in case of a factor of 0.5) than
the one measured here by lab-scale experiments. It must
be nally noted that the uncertainty of the intrinsic rate
constant experimentally determined is also relatively large
(see Fig. 1).
C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264 4263
350
390
430
470
510
550
4000 4500 5000 5500 6000 6500 7000
Length of reactor in mm
T
e
m
p
e
r
a
t
u
r
e

i
n

C
Regeneration time:19 h
modelled
profil
measured
profil C , m
k
C , m
k 5 . 0
C , m
k 2
Fig. 17. Inuence of the intrinsic reaction rate constant on the modelled
temperature prole of the reaction front and comparison with the mea-
sured prole in the technical reactor of the (MIRO-renery in Karlsruhe,
Germany) (conditions: see Tables 2 and 3; p =20 bar; y
O
2
=0, 9 vol%;
u
L
=0.26 m/s; L
C,0
=21 g
C
/100 g
cat
).
6. Conclusions
The kinetics of coke burn-off were determined by sys-
tematic lab-scale experiments. The experiments and theoret-
ical considerations clearly indicate that for carbon loads of
10 g
C
/100 g
cat
and temperatures above 400

C the effective
rate of carbon burn-off is strongly inuenced by pore diffu-
sion. The external gassolid mass transfer can be neglected
for temperatures below 750

C, i.e., for temperatures of in-


dustrial relevance (T <550

C). Based on the kinetic data
and the structural parameters of the catalyst (porosity, tor-
tuosity), the regeneration process was modelled both on the
level of a single particle and in a technical xed bed reactor.
The results of modelling were compared with coke proles
within the particles and the performance data of regenera-
tion of an industrial reactor. The agreement of calculation
and measurement is in both cases complete.
Notation
A
m
external surface area per mass of catalyst, m
2
A
BET
internal surface area per mass of catalyst,
m
2
/kg
A
V
external surface area per volume of catalyst
m
2
/m
3
c
O
2
concentration of O
2
, mol/m
3
c
O
2
,G
concentration of O
2
in the gas phase, mol/m
3
c
p,G
heat capacity of gas phase, J/(mol K)
c
S
heat capacity of solid phase, J/(kg K)
d
P
diameter of particle, m
D
ax,eff
effective axial dispersion coefcient, m
2
/s
D
f b
dispersion coefcient D
ax,eff
without gas
ow, m
2
/s
D
O
2
,Knu
Knudsen diffusion coefcient, m
2
/s
D
O
2
,mol
molecular diffusion coefcient, m
2
/s
D
O
2
,pore
pore diffusion coefcient, m
2
/s
D
O
2
,eff
effective diffusion coefcient, m
2
/s
E
A
activation energy, J/mol
EDX energy dispersive X-ray spectrometer
k
m,C
reaction rate constant, m
3
/(kg s)
k
m,C,0
frequency factor, m
3
/(kg s)
k
P
ratio of thermal conductivities (=z
S
/z
G
)
L
C
carbon load of catalyst (=m
C
/m
cat
),
kg
C
/kg
cat
L
C,0
initial coke content of catalyst, kg
C
/kg
cat
L
p
length of a cylindrical particle, m
m
cat
mass of catalyst, kg
M
C
molecular weight of carbon, kg/mol
M
O
2
molecular weight of oxygen, kg/mol
NMR Nuclear magnetic resonance
n
C
amount of carbon, mol
n
O
2
rate of oxygen, mol/s
p pressure
PFG Pulsed eld gradient (NMR)
PFR Plug-ow reactor
q order of reaction
Q heat ux, J/s
r
m
reaction rate per unit of mass of catalyst,
mol/(kg s)
r
m,eff
effective rate per mass of catalyst, mol/(kg s)
r
p
particle radius, m
R ideal gas law constant (8.314), J/(mol K)
SEM Scanning electron microscopy
Sh Sherwood number, [d
P
/D
mol
t time, s
T temperature,

C, K
T
G
temperature of the gas phase,

C, K
T
S
temperature of the surface of particle,

C, K
T
0
initial temperature,

C, K
T
max
maximum temperature,

C, K
u
e
supercial gas velocity, m/s
u
G
interstitial gas velocity, m/s
u
RF
velocity of the reaction front, m/s
u
HF
velocity of the heat front, m/s
V
C
core volume, m
3
V
m
volume of particle, m
3
/kg
X
O
2
conversion of oxygen
y gas volume fraction
z axial coordinate in reactor, m
Greek letters
: heat transfer coefcient, W/(m
2
K)
[ mass transfer coefcient, m/s
A
R
H heat of reaction, J/mol
t time interval
T
ad
adiabatic temperature increase,

C, K
T
max
maximum adiabatic temperature increase,

C, K
4264 C. Kern, A. Jess / Chemical Engineering Science 60 (2005) 42494264
z interval in axial direction
c voidage of reactor bed
c
P
porosity of particle
p
pore
pore effectiveness factor
p
pore,0
initial pore effectiveness factor
p
overall
overall particle effectiveness factor
z
ax,eff
effective dispersion coefcient of heat,
W/(mK)
z
f b
effective dispersion coefcient of heat z
ax,eff
without gas ow, W/(mK)
z
S
thermal conductivity of porous solid phase,
W/(mK)
z
G
thermal conductivity of gas phase, W/(mK)
j
P
density of catalyst, kg/m
3
j
G
density of gas phase, kg/m
3
j
B
density of reactor bed
t
P
particle tortuosity
t modied resicence time, kg s/m
3
[ Thiele modulus
[
0
initial Thiele modulus
Acknowledgements
Financial support by the Max-Buchner-Forschungsstiftung
(1970) and by the Deutsche Forschungsgemeinschaft (BL
231/25) is gratefully acknowledged. The authors also wish
to thank Engelhard for supplying the catalyst, the De-
partment of Materials Processing (University Bayreuth,
Prof. M. Willert-Porada) for the helpful investigations on
the carbon distribution within the catalyst (SEM/EDX-
measurements), Prof. B. Blmich and his co-workers (In-
stitute of Technical Chemistry and Macromolecular Chem-
istry, University of Aachen) for the fruitful cooperation and
the NMR-measurements of the tortuosity, and nally the
MIRO-renery (Karlsruhe, Germany) for the provision of
performance data of the regeneration of the coked reforming
catalyst in a real technical xed bed reactor.
References
Baerns, M., Hofmann, H., Renken, A., 1987. Chemische Reaktionstechnik.
Lehrbuch der Technischen Chemie, Band I. Georg Thieme Verlag,
Stuttgart New York.
Beltramini, J.B., Wessel, T.J., Datta, R., 1991. Kinetics of deactivation of
bifunctional Pt/Al
2
O
3
-chlorided catalysts by coking. A.I.Ch.E. Journal
37, 845.
Eigenberger, G., 1983. Dynamik und Stabilitt verfahrenstechn. Prozesse.
Chemie Ingenieur Technik 55, 503.
Emig, G., Hofmann, H., Hoffmann, U., Fiand, U., 1980. Experimental
studies on runaway of catalytic xed-bed reactors. Chemical
Engineering Science 35, 249.
Froment, G.F., Bischoff, K.B., 1990. Chemical Reactor Analysis and
Design. second ed. Wiley, New York, USA.
Fung, S.C., Querini, C.A., McCoy, C.J., 1991. Coke and product proles
along catalyst bed in reforming reactions. In: Bartholomew, C.H., Butt,
J.B. (Eds.), Catalyst Deactivation. Elsevier, Amsterdam, p. 135.
Hein, O., Jess, A., 2000. Bestimmung kinetischer Parameter exothermer
Gas/Feststoff-Reaktionen mit der Zndpunktsmethode. Erdl Erdgas
Kohle 116, 18.
Herbig, C., 2002. Zndverhalten exothermer Reaktionssysteme in
statischen und in durchstrmten Laborreaktoren. Doctoral Thesis,
University of Technology, Aachen.
Herbig, C., Jess, A., 2002. Determination of reactivity and ignition
behaviour of solid fuels based on combustion experiments under static
and continuous ow condition. Fuel 81, 2387.
Ishida, M., Wen, C.Y., 1968. Comparison of kinetic and diffusional models
for solidgas reactions. A.I.Ch.E. Journal 14, 311.
Jess, A., Kern, C., 2001. Modelling of the regeneration of a coked xed bed
catalyst based on kinetic studies of coke burn-off. Studies in Surface
Science and Catalysis, vol. 139. Elsevier Science, Amsterdam, p. 447.
Jess, A., Hein, O., Kern, C., 1999. Deactivation and decoking of a naphtha
reforming catalyst. Studies in Surface Science and Catalysis 126, 81.
Kern, C., 2003. Regeneration eines koksbeladenen Katalysators zur
Benzinreformierung. Doctoral Thesis, University of Technology,
Aachen.
LePage, J.-F., 1978. Applied Heterogeneous Catalysis: Design,
Manufacture, Use of Solid Catalysts. d. Technip, Paris, p. 491.
Parera, J.M., 1991. Deactivation and regeneration of PtRe/Al
2
O
3
-
catalysts. In: Bartholomew, C.H., Butt, J.B. (Eds.), Catalyst
Deactivation. Elsevier, Amsterdam, p. 103.
Ren, X.H., 2003. NMR studies of molecular transport in model xed-bed
reactors. Doctoral Thesis, University of Technology, Aachen.
Ren, X.H., Bartusseck, I., Demco, D., Stapf, S., Blmich, B., Kern, C.,
Jess A., 2000. NMR study of changing pore size and tortuosity during
deactivation and decoking of a naphtha reforming catalyst. Proceedings
of the Conference on Magnetic Resonance Applications to Porous
Media. Bologna, October 2000, p. 47.
Ren, X.H., Bertmer, M., Stapf, S., Demco, D., Blmich, B., Kern, C.,
Jess, A., 2002. Deactivation and regeneration of a naphtha reforming
catalyst. Applied Catalysis A: General 228, 39.
Schlnder, E.-U., 1986. Einfhrung in die Wrmebertragung. F. Vieweg
& Sohn, Braunschweig/Wiesbaden.
Tang, D., 2004. Investigations on diffusion and dispersion in reacting
gas/solid-systemscombined use of reaction engineering methods and
NMR-techniques. Doctoral Thesis, University of Technology, Aachen.
Tang, D., Kern, C., Jess, A., 2004. Inuence of chemical reaction rate,
diffusion and pore structure on the regeneration of a coked Al
2
O
3
-
catalyst. Applied Catalysis A: General 272, 187.
Verein Deutscher Ingenieure (Eds.), 2002. VDI-Wrmeatlas.
Berechnungsbltter fr den Wrmebergang, ninth ed., Mh 5, Dee 3,
Dee 4. VDI-Verlag.
Weisz, P., Goodwin, R.D., 1963. Combustion of carbonaceous deposits
within porous catalyst particles I. Diffusion-controled kinetics. Journal
of Catalysis 2, 397.
Weisz, P., Goodwin, R.D., 1966. Combustion of carbonaceous deposits
within porous catalyst particles II, intrinsic burning rate. Journal of
Catalysis 6, 227.
Wen, C.Y., 1968. Noncatalytic heterogeneous solid uid reaction models.
Industrial and Engineering Chemistry 60/9, 34.
Westerterp, K.R., Fontein, H.J., van Beckum, F.P.H., 1988. Decoking of
xed-bed catalytic reactors. Chemical Engineering Technology 11, 367.
Wicke, E., Vortmeyer, D., 1959. Zndzonen heterogener Reaktionen in
gasdurchstrmten Krnerschichten. Berichte Der Bunsen-Gesellschaft-
Physikalische Chemie 63, 145.
Wulkow, M., Gerstlauer, A., Nieken, U., 2001. Modeling and simulation of
crystallization processes using parsival. Chemical Engineering Science
56, 2575.
Zhang, T., Zang, J., Lin, L., 1991. Relation between surface structure
and carbon deposition on PtAl
2
O
3
and PtSn/Al
2
O
3
-catalysts. In:
Bartholomew, C.H., Butt, J.B. (Eds.), Catalyst Deactivation. Elsevier,
Amsterdam, p. 143.

Das könnte Ihnen auch gefallen