Sie sind auf Seite 1von 6

Studies in Surface Science and Catalysis

J.J. Spivey, E. Iglesia and T.H. Fleisch (Editors)


o 2001 Elsevier Science B.V. All rights reserved.

313

Ethylene production via Partial Oxidation and Pyrolysis of Ethane


M. Dente 1, A. Berettal, T. Faravelli 1, E. Ranzi l, A. Abbr. 2, M. Notarbartolo3
ICIIC Department Politecnico di Milano, Italy
2Math. Department Politecnico di Milano, Italy
3Technip Rome, Italy
This paper presents the theoretical study of a homogeneous process to produce ethylene by ethane
dehydrogenation. A hydrogen-oxygen diffusive flame in a conventional burner is used as a hot
stream into which the ethane feed is injected. The results indicate that very high selectivities to
ethylene can be reached at ethane conversion greater than 70%. The advantages of the process are
mainly related to the environmental aspects. The high efficiency of the heat exchange, the use of a
clean fuel, like the H2 produced, and the absence of nitrogen drastically reduce the pollutant
emissions. The homogeneous process avoids catalyst poisoning and deactivation. The main
limitation in this configuration is the very fast required mixing between the two streams.
INTRODUCTION
The pyrolysis of hydrocarbon feedstocks in the steam cracking process is the most commonly used
technology to produce ethylene along with other important chemical commodities. In recent years,
the need of increasing the efficiency and reducing the environmental impact of such plants has led
to research in this mature technology. In particular the ethylene production via the oxidative
dehydrogenation of ethane received a great interest (Bodke et al., 1999; Zerkle et al., 2000). The
new process alternative to convert ethane to ethylene is a short contact time reactor (a few
milliseconds) in which the reaction takes place in the presence of a foam monolith of alumina
supported Pt-Sn catalyst. The process can include a premixed Hz addition in the feed. The catalyst
function is to promote the H2 oxidation to H20 avoiding the ethane and ethylene oxidation. CO and
CO2 selectivity significantly decreases and the heat of H2 oxidation reaction allows the endothermic
C2H6 dehydrogenation to occur in a very selective way.
The two stages of the process can be then globally summarized as:
H2 + 89 02
C2H6

--) H20
") C2H4 + H2.

C2H6 + 89 O2 "-) C2H4 + H20


The overall reaction is exothermic, more hydrogen is produced than the required for autothermal
conditions and no additional fuel is needed.
In this scenario, it is then quite evident the advantages offered by such a process. From the
environmental point of view, it reduces the greenhouse gas emissions, using hydrogen as a clean
fuel, while simultaneously the use of pure oxygen avoids the NOx formation. Moreover, the
experimental pilot results seem to indicate that this route is also economically convenient. The
reactor is smaller and simpler than the conventional cracking furnaces and because of the higher
reactor temperatures, selectivities could also be improved. More than 85% ethylene selectivity was
obtained at 70% ethane conversion. On the other hand, higher operation costs are expected due to
the required pure oxygen. These considerations together with correct economic evaluations seem to
make feasible this new approach (Schmidt et al., 2000).
From a scientific point of view, the reaction mechanism is still under debate. The important role
played by gas-phase reactions at this high reactor temperatures has been demonstrated (Beretta et
al., 2000). The catalyst mainly drives the hydrogen oxidation, while the pyrolysis reactions occur in

314
the homogeneous phase. As a matter of fact the same, or even higher, ethylene selectivities at the
same ethane conversions can be reached with a very rapid heating of the ethane feed. The high
conversion obtained at very short contact times in the catalyzed system can be reasonably explained
with active species (radicals) generated by the catalytic surface and released in the gas phase
(Schmidt et al., 2000).
On the basis of these results, it is possible to propose a purely homogeneous system which can
reach the same performance in terms of both
100
selective ethylene production and effective
oxidative
pollutant reduction. This new process, without
=
80
O
catalyst, only involves a primary hydrogen
.,-~
flame followed by ethane injection. Two
60
major problems can be avoided with this
rsis
0
homogeneous approach. From one side, the
~ 40
aging and poisoning of the catalyst, that can
e-I
diminish
its activity with a reduction of
~ 20
ethylene selectivity in favor of CO and CO2
formation.
A second
and
significant
0 ~
0.0001
0.001
0.01
0.1
1 improvement of the homogeneous system is
related to the increase of plant safety. The
Residence time (s)
diffusive hydrogen flame in a conventional
burner allows to avoid the premixing of
Figure 1. Computed comparison between
hydrogen, ethane, and oxygen. The potentially
pyrolysis and oxidative pyrolysis conversion
high adiabatic temperatures of the hydrogen
oxygen system can be easily managed and can be reduced with steam additions in order to reach the
optimum cracking temperature.
MODEL RESULTS
The evaluations of the homogeneous oxidative pyrolysis are computed by using detailed kinetic
models for the pyrolysis (Dente and Ranzi,
1983) and oxidation of hydrocarbons (Ranzi
100[
et al., 2000). The model herein applied is also
pyrolysis
able to predict the formation of heavier
90
species and especially Poly Aromatic .~
Hydrocarbons (PAH) precursors of coke
> 80
formation. This model has been already ~
9 9
extensively
tested
and
validated
by
~
pyrolysis
comparison with experimental data carried a , 70
out in a very wide range of operative
conditions. It has been also used to confirm
60
0
20
40
60
80
100
the relevance of the homogeneous gas-phase
reactions in the short contact time catalytic
C2H6 conversion
reactors (Bodke et al., 1999).
Figure 2. Computed comparison between C2H4
As already discussed in the past (Chen et al.,
1997), the introduction of a small amount of
selectivity in the pyrolysis and oxidative pyrolysis
oxygen in the conventional coils of steam
cases
cracking processes increases the plant
capacity. The heat released from the partial oxidation of hydrocarbons, directly generated inside the
reactor together with very active radicals, like OH, enhances the system reactivity.

315
As an example, fig. 1 shows the conversion of ethane in pyrolysis (C2H6:H20= 2:1 mol) and partial
oxidation systems (C2H6:O2 = 2:1 mol). These data were obtained by simulating an isothermal plug
flow reactor at 1200 K. The presence of oxygen increases the conversion and allows to work with
shorter contact times. Fig. 2 shows the reduction of ethylene selectivity due to the CO and CO2
100
96

.~
7:r,l

.~ 80

92
88

oxt~ 40

84

2o

80 t
0

1300K

ra~

20

40
60
C2H6 conversion

80

100

Figure 3. Computed selectivities of


isothermal ethane pyrolysis

ll00K

_L~

0.0001

:: . . . . . . . . . . . . . . . . . . . .

0.001
0.01
0.1
Residence time (s)

Figure 4. Computed profiles of isothermal


ethane pyrolysis

formation. In purely homogeneous conditions, the introduction of an equimolar amount of hydrogen


(C2H6:O2:H2 = 2:1:2) does not significantly modify the performances of the oxidative pyrolysis.
This is an implicit confirmation that the higher selectivity of ethylene observed by Bodke et al.
(1999) with H2 addition is due to a catalyzed H2/O2 combustion.
In order to investigate the optimal conditions for the pyrolysis it is convenient to observe the results
shown in fig. 3 , where the isothermal pure pyrolysis of ethane is investigatedat different
temperatures. These theoretical results were obtained always with the kinetic scheme here adopted.
_

-T T
H20

H 2

,,. . . . . . . . . . . . . . . . . . . . . . . . . . .

02

_ .............................

--T
.......

C 2 H6

C2H4 + byProducts

Figure 5. Sketch of the process configuration


The higher temperatures of ethane pyrolysis reduce the ethylene selectivity, in favor of larger
amounts of acetylene. The higher temperatures are obviously accompanied by a strong increase in
ethane conversion or, in other words, by significant reductions of the reactor contact times (fig. 4).
Note that at ethane conversion greater than about 60%, ethylene selectivities show a sharp reduction
due to successive condensation reactions. This behavior becomes more evident at lower
temperatures. As already mentioned, the reactor temperature can be achieved by a direct injection of
ethane in a hot stream, with a very effective mixing. The hot stream can be obtained by recycling
and directly burning the same hydrogen produced by the process. Further steam has to be
conveniently added to the hydrogen/oxygen flame to reduce the temperatures. The overall
exothermicity of the reaction produces energy as byproduct of the process.
Fig 5 sketches the very simple process design in which the ethane feed is directly mixed with the
hot flue gases of the H2-O2 flame already mixed with a proper amount of dilution steam.

316

Fig. 6 shows the ethylene selectivities versus ethane conversion in an adiabatic pyrolysis reactor at
atmospheric pressure and different steam to reactant weight ratios, assuming the hot gases at 1600
K and the C2H6 feed at 1000 K.
96
,.O:C,H, = 2:1

92

H~o:c,u,=2.5:1

H20:C,H~= 3:1

,,,,~

"~ 88
~ 84
80

20

40
60
C2H6 conversion

~20
0

80

Figure 6. Computed ethylene selectivity at


different steam/feed ratios

0.02
0.04
0.06
Residence time (s)

0.08

0.I

Figure 7. Computed conversion at different


steam/feed ratios

Considering an instantaneous and ideally perfect mixing, the initial pyrolysis temperatures range
from about 1380 K in the 3"1 (wt) case to about 1320 K in the 2:1 (wt) case. As already discussed in
the isothermal case, also in these conditions, lower temperatures give better initial ethylene
selectivities. On the other side, lower temperatures increase the contact time required to obtain the
same conversion (fig. 7). The 70% ethane conversion, with about 90% selectivity, is achieved in
about 70 ms for the H20:C2H6=3:1 ratio. About 400 ms are required in the H20:C2H6=2.5:1 case.
At the lowest steam to ethane ratio, the initial temperature is too low to convert ethane over 65% in
reasonable contact times.
6
_

. .

...

r..)

~, 4
.>_.
O

_ _

. . . . .

~ .

. . . . . .

1
0

20

40

60

80

C2H 6 conversion

Figure 8. Selectivity of main dehydrogenated


species for H20:C2H6 = 3:1

00

20
C2H 6

40

60

8O

conversion

Figure 9. Selectivity of main dehydrogenated


species for H20:C2H6 = 2.5:1

The successive reactions toward more dehydrogenated and heavier species give rise to acetylene,
butadiene and benzene. When the first aromatic ring is formed successive interactions with mainly
vinyl radical and acetylene as well as cyclopentadiene, produce PAH, precursors of coke. The
model predicts the formation of such intermediates as reported in figs. 8 and 9, respectively for the
3:1 and 2.5:1 cases. Benzene and naphthalene selectivities are always lower than 1% even for
ethane conversion (i.e. residence times) longer than the optimal conditions.
In the homogeneous conditions herein discussed, the real bottleneck is the fluidynamic
effectiveness of the mixing. As a matter of facts, the estimated reaction times are in the order of a
few milliseconds, therefore it is necessary to realize an effective mixing in a very short time.

317
The splitting of the hot gases in two different streams with two successive injections (fig. 10) allows
to relax the stringent boundaries of the characteristic mixing times.
In this case the first ideal mixing temperature is lower and it is possible not only to obtain higher
selectivity but mainly to operate at relatively higher mixing times reducing in this way the crucial
aspects of the injection step. As an example fig. 11 shows the reduction of the maximum reactor
temperature of the 2 stage case when compared with the HzO:C2H6=3:1 ratio. These results are
computed for an initial mixing stage with only 2/3 of the total steam, the remaining steam being
injected, always at 1600 K, after about 20 ms. Figs. 12 and 13 show that the conversion of 70% can
still be achieved in about 90 ms, with a better ethylene selectivity.
The numerical modeling of the adiabatic mixing of the two reacting streams, hot flame gases and
n2

BURNER

H20

MIXING . . . . PYROLYSIS
. . .
__

02

MIXING

PYROLYSIS

'~]

C 2 H6

SEPARATION
I

C2H4 + byProducts

Figure 10. Sketch of the two stage mixing process configuration


C2H6 feed, can be conveniently addressed by the proper use of CFD codes. The presence of
endothermic reactions, the need of correctly characterizing the influence of mutual interactions
between turbulence and chemical reactions, the relative small dimensions of the reactor, and
especially its capability to well represent the mixing phenomena make the Large Eddy Simulation
(LES) technique an attractive alternative to the traditional methods for the numerical solution of the
1400~
resulting flows equations (Abb~ et al.,
1996). The description of the numerical
=3"1
method is beyond the aim of this paper,
~' 1300 Fnevertheless it is worth to highlight the
approach we are using to manage the
2 stag~'
generation
term in the species conservation
"~
9 1200 f
mixing
equations.
The filtered source term ~ in the species
~ 1100 t
conservation equations is:
1000/

le-5

. . . . . . . . . . . . . .

0.0001
0.001
0.01
Contact time Is]

0.1

NR ~ A ^-Ei/RT'r-[NS. vii
~" --" Zi=l/O'/"~iC
I lj=l Y J

where ~ is the filtered density, NS and NR


are respectively the number of species and
Figure 11. Reactor temperature for the 2 stage
reactions, Ai, Ei, vii the frequency factor, the
mixing
activation energy and the stoichiometric
coefficients of the j-th species in the i-th
reaction, and y is the concentration of the chemical species . . . . . .
Several authors employ a large eddy pdf (probability density function) approach, with a great
computational effort (Girimaji et al., 1996).
As a first approximation, it is possible to assume:
~ s Y~,~
e-Ei/RR'I"Hj=I
J ~

e-~l-i~-~,,
x xj--lYj

and to approximate the exponential term with the Taylor expansion around T

318

e -z~T =__e-Z~'~Ru

2[( R
LETi ~/)2 -,

Ei

] T'-~'I Since the temperature field is characterized by a

spectrum similar to the energy one (Girimaji et al., 1996), it is convenient to apply the scale
similarity theory and to take the subgrid-scales temperature fluctuations T-r'similar to turbulent
resolved ones T 'a = ~-7_~2
The results of this computations are only at a preliminary stage but already confirm the validity of
this approach and the critical aspect of the assumption of an ideal mixing.
80
92

60

0
o

= 4o
0
20

[~~
~

2,,,,o
xio,

........

,.Ha= 2:1

88

H,.O:C~.H6= 2:1

84
0.02

0.04
0.06
0.08
Residence time (s)

0.1

Figure 12. Computed conversion at different


steam/feed ratios and for a 2 stage mixing

80

20

40
60
80
C2I-I6 conversion
Figure 13. Computed ethylene selectivity at
different steam/feed ratios and for a 2 stage mixing

CONCLUSION
This paper presents an alternative route for ethylene production from ethane. The process, based on
a hydrogen/oxygen flame in which a C2H6 stream is fed, is autothermal, does not require extra fuel
and overall drastically reduces the environmental impact of conventional steam crackers. The
emissions of greenhouse gases COx and NOx are drastically cut and also the reactor design is
significantly simplified. Moreover the purely homogeneous system makes more safe the process
and it avoids the presence of dangerous explosive mixtures.
However, there still remains the need of a careful investigation on the fluidynamic aspects involved
in the ideal and very rapid mixing, to make technologically feasible the proposed process.
Finally, this approach could be also extended to different and heavier feedstock, like naphthas and
gasoils.
REFERENCES
.
Abba', A., Cercignani, C., Valdettaro, L., Zanini, P. "LES of turbulent thermal convection", Proc. of
the Second ERCOFTAC Workshop on Direct and Large Eddy Simulation, (1996)
Bodke, A.S., Olschki, D.A., Schmidt, L.D., Ranzi, E., Science, 285:712 (1999)
Beretta, A., Ranzi, E., Forzatti, P., Catalysis today, 2238:1 (2000)
Chen, Q., Schweitzer., E.J.A., Van Den Oosterkamp, P.F., Berger, R.J., De Smet C.R.H., Marin,
G.B., Incl. Eng. Chem. Res., 36:3248 (1997)
Dente, M. and Ranzi, E.. Chapter 7 in 'Pyrolysis Theory and Industrial Practice' (L. Albright.
Bryce and Corcoran Eds.) Academic Press San Diego, (1983).
Girimaji, S.S., Zhou, Y., Phys. ofFluids vol. 8, pp. 1224-1236, (1996)
Ranzi, E., Dente, M., Goldaniga, A., Bozzano, G. and Faravelli, T., Lumping Procedures in
Detailed Kinetic Modeling of Gasification, Pyrolysis, Partial Oxidation and Combustion of
Hydrocarbon Mixtures, Prog. Energy Combust. Sci. (2000)
Schmidt, L.D., Siddall, J., Bearden, M., AIChE J., 46, 8:1492 (2000)
Zerkle, D.K., Allendorf, M.D., Wolf, M., Deutchmann, O., Proc. Comb. Inst., 28 (2000)

Das könnte Ihnen auch gefallen