Sei sulla pagina 1di 7

2584

Ind. Eng. C h e m . Res. 1994,33, 2584-2590

Coke Formation in the Thermal Cracking of Hydrocarbons. 4.


Modeling of Coke Formation in Naphtha Cracking
Geert C. Reynierst and Gilbert F. Fromenti
Laboratorium uoor Petrochemische Techniek, Universiteit Gent, Krggslaan 281, S 5, B9000 Gent, Belgium

Frank-Dieter Kopinke and Gerhard Zimmermann


Abteilung Hochtemperaturreaktionen a m Institut fur Technische Chernie der Universitat Leipzig,
0-04303 Leipzig, Permoserstrasse 15, Germany

An extensive experimental program has been carried out in a pilot unit for the thermal cracking
of hydrocarbons. On the basis of the experimental information and the insight in the mechanisms
for coke formation in pyrolysis reactors, a mathematical model describing the coke formation
has been derived. This model has been incorporated in the existing simulation tools at the
Laboratorium voor Petrochemische Techniek, and the run length of a n industrial naphtha
cracking furnace has been accurately simulated. In this way the coking model has been validated.
Introduction
One of the main problems in the thermal cracking of
hydrocarbons for olefin production is the formation of a
coke layer on the inner wall of the coils. This leads in
the first place to a decrease of the heat flux to the
reacting gas mixture and second also to an increase of
the pressure drop over the reactor. Periodically the
furnace operation has to be stopped and the coke has
to be burnt off by means of a mixture of steam and air.
In the previous papers in this series (Kopinke et al.,
1988, 1993a,b) relative rates of coke formation from
various types of carbon atoms (aliphatic, aromatic, ...I
were determined. The relative rates were determined
by adding a small but precisely defined amount of a I4Clabeled component to a naphtha feed.
The present paper provides a mechanistic interpretation of the experiments performed by Kopinke et al. and
of additional experimental information obtained in the
pilot plant for thermal cracking of the Laboratorium
voor Petrochemische Techniek.
Experimental information and mechanistic interpretation were combined to develop a model describing
quantitatively the coking process in thermal cracking
coils. The coking model was combined with a kinetic
model for the pyrolysis of liquid feedstocks, a reactor
model and a furnace model. The combination of these
models allows the accurate prediction of the run length
of a furnace for the thermal cracking of naphtha.

Experimental Procedure and Results


The pilot plant for thermal cracking at the Laboratorium voor Petrochemische Techniek has been described elsewhere (Van Damme and Froment, 1982).
The unit is very flexible as to feedstock and operating
conditions. A rigorous procedure was followed in the
coking experimentation reported in the present paper:
1. An oxidative pretreatment of the reactor walls was
applied prior to each experiment. Twice-distilled water
was fed at a rate of 1k g h to the coil which was kept at
600 "C.

* Author to whom correspondence should be addressed.


E-mail: gf@elptes3.rug.ac.be.
' Present address: ENSIM N.V., Noorderlaan 127, B2030
Antwerp, Belgium.

2. Next, the coil was heated, still under a flow of


steam, to the desired temperature profile.
3. Next the hydrocarbon feed was introduced. This
caused a disturbance of the temperature profile, due to
the endothermic reactions. This transient period was
kept as short as possible. Within 15 min the reactor
was again in steady state.
4. The operating conditions were kept constant for a
period of 6 h. During this period, several analyses of
the reactor effluent were carried out. The pressure and
temperature profiles in the reactor were continuously
monitored.
5 . After 6 h the hydrocarbon feed was stopped. This,
again caused an upset in the temperature profile.
Hence, a stabilization period of 15 min was required.
6. Finally, the coke formed during the experiment
was burnt off by means of a mixture of steam and air.
The carbon monoxide and carbon dioxide concentrations
in the reactor effluent were measured continuously by
infrared analyzers, together with the total volumetric
effluent flow. Decoking mainly took place in a reaction
front wandering through the reactor. The temperature
in the reaction front amounted to about 1000 "C.
Forty-five experiments were carried out, with widely
different feedstocks and operating conditions. Feedstocks ranged from CH4 to n- and i-Cs, C2H4, benzene,
and mixtures thereof, naphtha and kerosene. The coil
outlet temperature ranged from 820 to 960 "C, the coil
outlet pressure ranged from 1.5 t o 3.6 bar absolute and
the diluent was either H2O or N2.
From the set of experiments a number of industrially
important conclusions can be drawn:
1. Under identical conditions ethane yields more coke
than propane or n-butane. From the discussion of the
mechanism of coke formation, it will become clear that
this is linked to the type of radicals produced with the
different feedstocks. In ethane cracking, hydrogen,
methyl, vinyl, and ethyl radicals are important. All
these are small and active radicals. In cracking longer
chain paraffins allyl radicals are also formed. Allyl
radicals are more stable and less reactive.
2. Isoparaffinic feedstocks yield more coke than the
corresponding normal paraffinic feedstocks.
3. Steam, used as a diluent, is not inert toward coke.
Experiments were carried out under exactly the same
conditions except that in one case steam dilution was
used and in the other case nitrogen, for the same molar

0888-5885/94/2633-2584$04.50/~ 0 1994 American Chemical Society

*+
- *

Ind. Eng. Chem. Res., Vol. 33,No. 11, 1994 2585

C
CxHy

A,

--.-

Figure 1. Mechanism of filament growth.

flow rate. In all cases the experiment with nitrogen


dilution yielded more coke than the one with steam
dilution. The difference in coke yield is strongly dependent on the type of feedstock used and on the
operating conditions. It is clear, however that steam is
able to remove part of the coke formed at temperatures
exceeding 850 C.
4. The rate of coke formation increases with pressure.
This indicates that bimolecular reactions are important.
In free radical reaction schemes, the main bimolecular
reactions are hydrogen abstractions and addition reactions.

Development of a Coking Model


The experimental observations and a detailed literature survey point to a complex mechanism of coke
formation in thermal cracking. Three mechanisms
contribute to the deposition of a coke layer.
Heterogeneous Catalytic Mechanism. During the
startup of a furnace, the reacting gas mixture is in
contact with the bare reactor walls. Incoloy 800H,
frequently used as the material of construction, contains
31 wt % nickel. By chemisorption, species from the gas
phase form a metal-hydrocarbon complex on the nickel
crystallites at the inner tube wall. This complex
decomposes and yields carbonaceous material on the
metal surface. The carbon migrates through the metal
particle by diffusion and precipitates a t the grain
boundaries. The driving force for the diffusion is a
difference in thermodynamic activity between the carbon at the metal surface and the carbon at the grain
boundaries (Bianchini and Lund, 1989). Diffusion is the
rate-determining step of the process. The precipitated
carbon causes stress in the metal structure. Eventually,
the stress becomes so large that the nickel crystallite
is removed from the metal structure. As more carbon
is deposited, a carbon filament, carrying a metal particle
on top of it, is formed. This is shown in Figure 1. In
the coke layer from an industrial reactor for the thermal
cracking of naphtha, a dense network of filaments has
been observed with filaments having a diameter between 2 and 5 pm and a length between 20 pm and l
mm.
The rate of carbon deposition in this initial phase is
very high. In an electrobalance type of reactor, also
available at the Laboratorium voor Petrochemische
Techniek, initial coking rates of more than 50 g/m2h
have been observed in the thermal cracking of ethane
at 810 C . After 15 min, however, the coking rate
decreases, to reach a n asymptotic value of less than 10
g/(m2h).
The filament network forms a porous structure.
Deficiencies in the graphitic structure act as active sites
on the outer walls of the filaments. On these active sites
coke deposition continues via the heterogeneous noncatalytic mechanism. The influence of the metal par-

Figure 2. Growth of carbon layer.

ticles on the coke formation decreases steadily as the


metal surface becomes covered by a carbon layer. A
remanent catalytic activity is always present, however.
Heterogeneous Noncatalytic Mechanism. This
mechanism is the most important one in the coke
formation. The overall structure of the filament coke
is graphitic. This implies the degradation of the hydrocarbons to an aromatic structure by condensation
and dehydrogenation.
At the gadcoke interface the polyaromatic layer is not
yet completely dehydrogenated. At this surface hydrogen abstraction reactions by free radicals from the gas
phase can occur. Hydrogen, methyl, and ethyl radicals
are the most active species. As a consequence, the
concentration of the active sites a t the coke surface
becomes a function of the gas phase composition. This
explains the experimental observation according t o
which feedstocks generating more active radicals also
yield more coke.
At the free radical positions on the coke surface,
certain gas phase molecules (coke precursors) react via
an addition mechanism. All unsaturated molecules
from the gas phase are potential precursors. In Figure
2, an example of the reaction sequence is shown with
l-hexene as a precursor.
The long aliphatic side chain of these molecules is
subject to decomposition. The remaining part of the
molecule reacts in a few steps to a ring structure, in
which the dehydrogenation reactions proceed very rapidly. In this way the aromatic structure continues to
grow further and the free radical site at the coke surface
is regenerated by further hydrogen abstraction.
This mechanism explains the formation of a deposit
consisting of graphitic layers containing carbon atoms
in sp2 hybridization. The hydrogen content of such a
deposit is very low, in agreement with industrial
observations. Several other reaction sequences, an
example of which is shown in Figure 3, yield crosslinked graphitic structures. The cross-linking of aromatic layers explains why samples of coke layers from
industrial coils are extremely hard and can hardly be
drilled.
It is clear that the number of possible reaction paths
is extremely large. Every molecule from the gas phase
is in principle a potential coke precursor. Since the
number of species in the reaction mixture is very large,
especially when cracking liquid feedstocks, it is impossible to take into account all those reactions in developing a model for the coke formation.
Free radicals contribute to coke formation via termination reactions with the coke macroradical. This

2586 Ind. Eng. Chem. Res., Vol. 33, No. 11, 1994

+ H -

b&+a-.
Figure 3. (a) Cross-linking with sp3
linking with sp2 carbon atom.

contribution is relatively unimportant, however, as is


clear from a comparison of the concentration of the free
radicals in the reaction mixture (typically 10-6-10-3 wt
%) with the coke yield, which is of the order of 0.1 wt
%. Free radicals mainly act in hydrogen abstraction
reactions.
Homogeneous Noncatalytic Mechanism. This
third mechanism implies the formation of polynuclear
aromatics in the gas phase via free radical reactions.
These large molecules grow in the gas phase to tar
droplets that can be liquid or even solid at the conditions
prevailing in a thermal cracking reactor. Part of the
droplets impinge on the tube wall. Some rebound into
the gas phase, but it is more likely that they adhere to
the surface and are incorporated in the coke layer, since
the outer surface of the droplets is not completely
dehydrogenated. Hence, hydrogen abstraction reactions
by gas phase radicals are possible and the coke layer
can grow further.
This mechanism is considered important only in the
cracking of heavy liquid feedstocks, such as atmospheric
or vacuum gas oil, or else when the temperature exceeds
900 C.

Quantitative Formulation of the Rate of Coking


In what follows the modeling of the coking in thermal
cracking reactors will concentrate on the main mechanism, the heterogeneous noncatalytic mechanism. Taking into account all the possible reaction pathways
would Iead t o an unrealistically high number of kinetic
parameters, and their estimation would not be possible
or at least inaccurate.
The number of reactions can be decreased by restricting the number of coke precursors.
Paraffins are the main components in a naphtha
feedstock. These components do not disappear through
addition reactions, so that their direct contribution to
coke formation is low. Moreover, the coking rate is
highest in the high-temperature section of the reactor.

In this section the paraffin content in the reacting


mixture has decreased to a large extent.
For the same reasons, naphthenes can also be neglected as direct precursors t o coke.
Unsaturates are a very important class of coke
precursors. They are reaction products of the pyrolysis
reactions so that their concentration in the hightemperature zone of the reactor is high. Furthermore,
unsaturates are reactive and are good candidates for
radical addition. C4 components, which are present in
high concentration, are important coke precursors.
Longer chain unsaturated components decompose rapidly to smaller components.
Aromatics form a second class of important coke
precursors. The aromatic ring structure is close to the
structure of the coke matrix. Further, (branched)
aromatics are reactive components, especially at the
high temperatures prevailing in thermal cracking coils.
The basis of the modeling is the presence of active
sites at the coke surface. These are in reality radical
positions. The coke is mainly graphitic in structure. The
free radical sites at the surface are all phenylic or
benzylic sites. As an approximation, the activity of
these sites is considered to be equal.
Addition reactions occurring at these sites regenerate
the free radical position, so that the total number of
active sites is constant. What follows then is a sequence
of dehydrogenationand cyclization reactions causing the
incorporation of the carbon atoms of the coke precursor
into the coke layer.
The essential elements of the reaction of a precursor
with an active site at the coke surface are (a) the growth
of the coke layer with a number of carbon atoms equal
t o the carbon number of the precursor and (b) the
regeneration of the active site at the coke surface.
An example of such a reaction with ethylene as a
precursor can be written

P+ C,H4-2C

+ P+ 2H,

and the corresponding rate equation for carbon formation:

The concentration of the free radicals at the coke


surface is unknown. The coke has a macroradical
character, mainly due to hydrogen abstraction by the
radicals from the gas phase. The influence of free
radicals as abstracting species in the coke formation can
be accounted for by introducing a multiplication factor
containing the concentration of the main reaction
products of the two most active radicals. The most
abundant and reactive of these are H and CH3. The
concentration of hydrogen and methane are taken to be
proportional to that of Hand CHf, respectively. The
following expression can then be written for the rate of
coke formation out of an ethylene precursor:

A similar approach for the other precursors leads to


a coke model containing 12 reactions in parallel. The
rate of coke formation may then be expressed as

Ind. Eng. Chem. Res., Vol. 33, No. 11,1994 2687


The precursors are classified into groups depending
upon their characteristic function (double bond, triple
bond, aromatic ring, ...).
The reactions of coke radicals with precursors belonging to the same group are considered t o have the same
activation energy. A reference component is chosen in
each group. The reference factors for the coke formation
out of the other members of the group are related to
that of the reference component through the relative
reactivities for coke formation derived by Kopinke et al.
(1988, 19931, thus reducing the number of independent
parameters to be estimated from the experimental data.

Simulation of the Run Length of a Furnace for


the Thermal Cracking of Naphtha
The coking model is combined with the kinetic model
for the pyrolysis of liquid feedstocks which generates
the local concentrations of Hz, CHI, and the coke
precursors along the coil. The structure of this model
has been described by Clymans and Froment (1984)and
Hillewaert et al. (1988). The rigorous kinetic model
enter in the plug flow continuity and energy equations
for the reactor simulation. The heat flux to the reactor
is generated by a detailed furnace model (Plehiers and
Froment, 1988; Rao et al., 1988). This combination of
models allows an accurate and detailed simulation of a
thermal cracking unit.
Table 1 presents the basic information about the
industrial furnace, the split coil reactors, and the
operating conditions. No further data are required to
set up the simulation model for the radiation section,
except the location of the burners in the walls and the
location of the reactor coils in the furnace. This
information is provided in Figure 4.
The furnace geometry is complex. Because of symmetry only half of the furnace and four reactor coils need
to be simulated. The approach followed for run length
simulations has been described in detail elsewhere
(Plehiers et al., 1990). The continuity equation for
carbon is integrated by incrementing the time in
discrete steps. The coking rate is considered to be
constant in each time interval. For this simulation,
time intervals of 200-300 h were appropriate. In total,
280 h of CPU time was required on a Data General
Eclipse MV15000/8 computer to carry out the complete
run length simulation.
The operating policy of the furnace consisted of
keeping the ethylene production constant in time.
Figure 5 shows the ethylene yield averaged over the four
reactors. It is constant over the whole run length within
0.1 wt % absolute or 0.3%relative. The feed rates were
also kept constant over the run length.
Significant differences were found, however, among
the individual coils. The ethylene yields from the four
coils are shown in Figure 6. The initial differences are
due to a distinct location of each coil with respect to the
burners. This results in differences in the radiative heat
transfer fluxes to each reactor and, hence, in different
naphtha conversions and ethylene yields. The evolution
with time of the ethylene yield from each reactor is not
identical. The growth of the coke layer is not the same
for each coil, and as a consequence the amount of heat
transferred to the reacting process gas and the pressure
level in the reactors evolve in a different way. These
are factors which affect the ethylene yield and selectivity.

Table 1. Basic Information for the Run Length


Simulation
Furnace Data
height (mm)
9090
length (mm)
14400
width (mm)
2500
number of burners
160
thickness of refractory (mm)
230
thickness of insulation (mm)
50
Fuel Conditions
fuel composition (~01%)
methane
95
5
hydrogen
initial fuel gas flow rate (kmoyh)
198.4
air excess (%I
10
uniform side wall firing
Reactor Configuration
8 coils, split coil design
total length (mm)
53890
number of passes
6
passes 1-4: split section
total length (mm)
35920
80.0
internal diameter (mm)
external diameter (mm)
95.6
passes 5,6: single tube section
total length (mm)
17970
internal diameter (mm)
114.3
external diameter (mm)
129.9
Operating Conditions
hydrocarbon flow rate per coil (kgh) 2785
steam dilution (kg/kg)
0.7
coil inlet temperature ("C)
620
coil outlet pressure (bar abs)
1.45
Material Properties
emissivity of furnace wall
0.60
emissivity of tubes
0.95
thermal conductivity (W/(m K))
refractory
0.0193
0.0452
insulation
tubes
-1.257
coke
6.46
specific gravity of coke (kg/m3)
1600

+ 118.0 x 10-6T (K)


+ 111.1x 10-6T(K)
+ 0.04327T(K)

Feedstock Characterization
specific gravity d(15/15)
0.7057
PNA analysis (wt %)
n-paraffins
33.0
isoparafhs
37.0
naphthenes
18.0
aromatics
12.0
boiling range ("C)
IBP
36.0
50%
94.5
FBP
161.0

Figure 7 shows the heat consumption of the furnace.


To keep the ethylene production constant, in spite of
the growing coke layer and its increasing heat transfer
resistance, requires an increase of the heat input. The
ethylene production cost, therefore, rises with time.
Figure 8 gives the evolution of the coil inlet pressures
for the individual reactors. To keep the coil outlet
pressure constant, the inlet pressure has to be raised.
This leads t o a higher pressure level in the reactor,
favoring the secondary (ethylene-consuming) reactions
over the primary (ethylene-producing) reactions. Secondary reactions are mostly bimolecular, while primary
reactions proceed via a monomolecular mechanism.
The most pronounced increase in coil inlet pressure
is found in coil 2, in which it amounts to 0.35 bar, a
23%relative increase with respect to the initial pressure
drop of 1.51 bar.

2688 Ind. Eng. Chem. Res., Vol. 33, No. 11,1994

a~I

a
TLE

3
i

I
I

I
I
I
I
I'

I
I

d
I
I

3I" 'i
I
I
I
I
I
I
I
I

I
I
4

I
I

I
I

I
I

I
1
I

I
d
I.
I
I

4
I

1
I
I

{\

lnlw

P
I

I
I
I
I

1I

II
I

01

I
I
I

+I
I
I

I
I
I
I

t
/

&

I
$ 1I

01

Figure 6. Evolution with time of ethylene yield in the various


coils.

I
I

44

I f

01

I
I

l o

//

I
I
$1
I
I
I

) I t

01

01

-m

II

01

I
I

t1

I
1
I

I t
I
I
I

I
+I
I
I
I

Il

I
I

4 4I I
I
4I 1 @1
I
I
I

\A 1

01

*t

1
1

'. '

0
m

Figure 7. Evolution with time of the total energy input to the


furnace.
Tmrmnw

RL

R2
1

1.1

Downflow

upflow

emir

!,/

3.1

I,/

\I

\I

\I

\I

!I'

\/

'.I

90)

lpoo

1.5-

-w

Figure 8. Evolution with time of inlet pressures.


bil=4hm

s.

Figure 4. (a) Front view of furnace. (b) Top view of furnace


showing burners and coils.
w (n+l

14t

Figure 9. Initial heat flux profiles.

7.
0

Ka

(.ooo

1-

2,000

Ik.(hl

Figure 5. Evolution with time of average ethylene yield.

The initial heat flux profiles are shown in Figure 9.


The heat input is highest for reactors 2 and 4, leading

to a higher temperature level in these reactors and in


a higher feedstock conversion.
Figure 10 shows the evolution of the heat flux profiles
as a function of time for reactor 4. The heat flux
increases in the split tube section of the reactors,
whereas in the single tube section the heat flux decreases with time.
The rise of the heat fluxes in the first part of the

Ind. Eng. Chem. Res., Vol. 33,No. 11, 1994 2689


F
ST

10

2s

16

10

0
0

10

-I*

Figure 10. Evolution with time of heat flux profile for coil 4.

Figure 12. Evolution of profiles of thickness of coke layer in coil


4.

wmi.*ohi

I
B

L
0

10

Io

50

rnDn In)

Figure 11. Evolution with time of profiles of coking rate in coil


4.

reactor results from the requirement of achieving a


higher conversion, so as to keep the ethylene yield
constant, in spite of the increase in pressure. It
compensates for the decrease of the heat fluxes in the
second part of the reactor caused by the coke formation.
The coke layer reaches its maximum thickness in the
second part of the coil.
It should be stressed that these results can only be
obtained from a detailed furnace and reactor simulation.
Assuming a translation of the heat flux profile with
time, as is normally done when performing only reactor
simulation, leads to an overestimate of the coking rates
and to predicted run lengths which are too short.
The evolution of the coking rates with time is shown
in Figure 11 (reactor 4). The coke formation takes place
at the temperature of the gadcoke interface. The coking
rates are high in the second part of the reactor. As a
consequence, the coke layer grows fast there and creates
an additional resistance to heat transfer. Further, the
coke layer also causes a decrease of the tube crosssectional area and this leads to higher linear gas
velocities in this section, causing an improved convective
heat transfer from the interface to the reacting gases.
These two effects lead to a decrease with time of the
interface temperatures in the single tube section of the
reactor. Together with the interface temperatures, the
coking rate also decreases with time.
In the split part of the reactor the coke layer is less
important. Increasing the heat fluxes with time increases the gadcoke interface temperatures and the
coking rates. The coke layer tends to become more
uniform along the tube with time.
The simulated thickness of the coke layer is given in
Figure 12 for reactor 4. The coke reaches its maximum
thickness on the last pass, immediately after the last
lower U-bend in the reactor. At the end of the run the
coke thickness is 8.7 mm, corresponding to a reduction
in diameter of more than 15%.

100
0

,
10

B
1

T
I

so

Figure 13. Profiles of external tube skin temperature, coil 4.


T-lV
1.1oo

mo

1Lo

1
m

2,m

Iu*in)

Figure 14. Evolution with time of external tube skin temperature.

The evolution with time of the coke layer is reflected


in the external wall temperatures. These are shown for
reactor 4 in Figure 13. The temperature rises with time
to meet the heat flux requirements, in spite of the
additional resistance to the heat transfer resulting from
the coke deposition. Initially, the temperatures are
highest in the high-temperature section of the reactor,
where the coking rates are highest and the coke layer
reaches its maximum thickness. The largest increase
in the external wall temperatures is observed in this
section. The maximum value is reached at a reactor
length of 47 m. Initially the maximum value is approximately 1000 "C.
Figure 14 shows the evolution with time of the
maximum external wall temperature for the four simulated reactors. This value should be watched carefully
during operation, since there is a limitation imposed by
the metallurgy. In the present case the limiting value
is 1090 "Cand it is reached on reactor 4 after 1900h or
about 80 days of operation, after which decoking is
necessary. The simulated value is in excellent agreement with the observed plant value, which was 85 days.
The temperature limit is reached on one reactor only.
For the other reactors there still is a margin of some 10

2590 Ind. Eng. Chem. Res., Vol. 33, No. 11, 1994

"C,meaning that their run length could have been


extended for 5-7 days. In commercial operation, it
might be worthwhile to decrease the cracking severity
on reactor 4,so as to prolong the production cycle.

Conclusions
Literature data and experimental observations were
combined to obtain mechanistic insight into coke formation in thermal cracking. Three mechanisms contribute.
Initially a filament layer is formed by interaction of gas
phase species with metal particles from the reactor wall.
The main mechanism is the interaction between active
sites on the coke layer with gas phase species. The
active sites are free radicals and originate from hydrogen abstraction by small active gas phase radicals, such
as hydrogen and methyl radicals. At the free radical
sites unsaturated precursors from the gas phase react
via addition, followed by a set of dehydrogenation and
cyclization reactions finally yielding a graphitic coke
layer. A third mechanism involves the formation of
droplets in the gas phase which then impinge and
adhere t o the surface and are incorporated in the coke
layer. This last contribution is thought to be important
only in the cracking of very heavy liquid feedstocks.
The mechanistic considerations and the experimental
information were used to develop a kinetic model in
which a set of 12 coke precursors form coke via a set of
parallel reactions.
This coking model was combined with a kinetic model
for the cracking, a reactor model and a furnace model
to simulate the run length of a furnace for the thermal
cracking of naphtha. Detailed and accurate information
can be obtained from this simulation. The growth of a
coke layer is accurately simulated, and so is the evolution of the external tube skin temperatures. The
simulated and experimental run lengths agree within
5%.
Simulations of this kind can be used to optimize
furnace operation for various feedstocks and operating
conditions. They can be used as a guide for the
adaptation of the operating variables aiming at prolonging the run length of the furnace.

Literature Cited
Bianchini, E. C.; Lund, C. R. F. Kinetic Implications of Mechanisms Proposed for Catalytic Carbon Filament Growth. J . Catal.
1989,177,
455-466.
Clymans, P. J.; Froment G. F. Computer Generation of Reaction
Paths and Rate Equations in the Thermal Cracking of Normal
and Branched Paraffins. Comput. Chem. Eng. 1984,8(2), 137142.
Figueiredo, J. L.; Orfao, J. J. M. Carbon Deposits on Metal
Catalysts-Mechanisms of Formation and Gasification. Catal.
Today 1989,5,385-393
Hillewaert, L. P.; Dierickx, J. L.; Froment, G. F. Computer
Generation of Reaction Schemes and Rate Equations for
Thermal Cracking. AIChE J . 1988,34(11, 17-24.
Kopinke, F. D.; Zimmermann, G.; Nowak, S. On the Mechanism
of Coke Formation in Steam Cracking-Conclusions from Results obtained by Tracer Experiments. Carbon 1988,26 (21,
117-124.
Kopinke, F. D.; Zimmermann, G.; Reyniers, G.; Froment, G. F.
Relative Rates of Coke Formation from Hydrocarbons in Steam
Cracking of Naphtha. 2. Paraffins, Naphthenes, Mono-, Di-, and
Cycloolefins, and Acetylenes. Ind. Eng. Chem. Res. 1993a,32
(11,56-60.
Kopinke, F.-D.; Zimmermann, G.; Reyniers, G.; Froment, G. F.
Relative Rates of Coke Formation from Hydrocarbons in Steam
Cracking of Naphtha. 3. Aromatic Hydrocarbons. Ind. Eng.
Chem. Res. 1993b,32(ll), 2620-2625.
Plehiers, P. M.; Froment, G. F. Firebox Simulation of Olefin Units.
Presented at the AIChE Spring National Meeting, March 6-10,
1988.
Plehiers, P. M.; Reyniers, G. C.; Froment, G . F. Simulation of the
Run Length of an Ethane Cracking Furnace. Ind. Eng. Chem.
Res. 1990,29 (41, 636-641.
Rao, M. V.; Plehiers, P. M.; Froment, G. F. The Coupled Simulation
of Heat Transfer and Reaction in a Pyrolysis Furnace. Chem.
Eng. Sci. 1988,43(6), 1223-1229.
Van Damme, P. S.; Froment, G. F. Thermal Cracking Computer
Control in Pilot Plants. Chem. Eng. Prog. 1982,78 (91, 77-82.

Received for review December 9, 1993


Revised manuscript received March 9, 1994
Accepted July 6 , 1994@
Abstract published in Advance A C S Abstracts, September
15, 1994.
@

Potrebbero piacerti anche