Sei sulla pagina 1di 8

Chemical Engineering Journal 207208 (2012) 473480

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Olens via catalytic partial oxidation of light alkanes over Pt/LaMnO3 monoliths
L. Basini a, S. Cimino b,, A. Guarinoni a, G. Russo b, V. Arca c
a

eni, Rening & Marketing Division, Italy


Istituto Ricerche sulla Combustione CNR, Italy
c
eni Versalis S.p.A., Italy
b

h i g h l i g h t s
" CPO of ethane and n-butane to olens was studied on PtSn/LaMnO3 honeycombs.
" Bench scale testing showed high single pass yields of C2H4 + C3H6 around 55 wt.%.
" Stable reactivity demonstrated for 500 h.o.s. with ethane feed and sacricial H2.
" PtSn/LaMnO3 catalyst guaranteed a net hydrogen production across CPO reactor.
" Products quenching & catalyst overheating issues were identied during scale-up.

a r t i c l e

i n f o

Article history:
Available online 13 July 2012
Keywords:
Olens production
Catalytic partial oxidation
Light alkanes
Pt perovskite
Structured catalyst
Long-term stability

a b s t r a c t
The reactivity of a multi-layered monolith catalyst containing Pt and Sn over LaMnO3/La-c-Al2O3/cordierite, previously studied in a lab-scale plant for producing ethylene via Short Contact Time Catalytic
Partial Oxidation of ethane, has been further and extensively investigated in a bench-scale plant with
higher production capacity. Ethylene yields exceeding 55 wt.% have been achieved and the reactivity performances have been maintained for more than 500 h.o.s. The experiments, while conrming the potential of the technology, have pointed out some weakness in catalyst stability and reactor design. The
bench-scale experimental study has also addressed the reactivity features of n-butane indicating that
ethylene + propylene yields approach 54 wt.% in a wide range of experimental conditions.
2012 Elsevier B.V. All rights reserved.

1. Introduction
Light olens are the most important building blocks for the
polymers and variety of intermediates industry. World demand
of ethylene and propylene is exceeding now 180 MTA (about 2/3
related to ethylene production), with an annual growth of 45%
in the next decade [1,2].
Steam cracking of hydrocarbons has been and still is the main
industrial technology for producing light olens [14]. However,
despite the technological improvements occurred in more than
50 years, steam cracking remains the most energy-consuming process in the petrochemical industry.
It is expected that the possibility to perform oxidative dehydrogenation (ODH) or oxy-cracking of light alkanes through Short
Contact Time Catalytic Partial Oxidation (SCTCPO) would lead
to a novel technology with low capital investment, improved

Corresponding author. Address: Istituto Ricerche sulla Combustione CNR, P.le V.


Tecchio 80, 80125 Napoli, Italy. Tel.: +39 081 7682233; fax: +39 081 5936936.
E-mail address: stefano.cimino@cnr.it (S. Cimino).
1385-8947/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.06.153

energy efciency [26] and reduced NOx and CO2 emissions. In particular, it has been shown that monolithic Pt-based catalysts, operated under autothermal conditions i.e. wherein the feed is
partially combusted to drive the endothermic cracking processcan efciently convert ethane to ethylene, propane and n-butane
to ethylene and propylene, isobutane to propylene, isobutene and
ethylene [320].
Recent experiments performed at laboratory scale utilizing ethane and a patented Pt(Sn)/LaMnO3 catalysts [10,11], have produced olens with yields exceeding 61 wt.% and selectivity above
75 wt.% per pass. It has been estimated that these reactivity features could result in reduced production cost of ethylene with respect to steam cracking [10]. Indeed the advantages resulting from
the high olen yields and the compact reactor system [46,9] could
more than compensate the additional oxygen consumption costs,
not to mention the benet of reducing the CO2 and NOx emissions
by avoiding large heating furnaces.
However, from a technical standpoint, a better denition of several key operating issues is required, including: catalyst activity
and stability as well as reactor design [5,8,9,12,13,16]. In fact, most
of the available experimental results were obtained with small

474

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

scale catalytic reactors operated with diluted streams for a rather


limited time on stream.
In this work we set out to validate previous ethane SCTCPO
performance data utilizing the same PtSn/LaMnO3 catalyst
[10,21] in a pre-pilot scale test rig with a 15 times larger production capacity, specically designed and operated by eni in order
to explore the effect of the main operational parameters (space
velocity, preheating temperature, feed ratio, sacricial fuel ratio,
pressure) on process performance and products distribution. With
the perspective of an industrial application, the target of our work
was to test the stability of the catalyst under relevant operating
conditions (i.e. adiabatic reactor, without feedstock dilution) for
at least 500 h on stream, while assessing key issues related to process scale-up. The catalyst, was fully characterized pre- and posttest by optical microscopy, SEM-BSE and XRD.
In addition, olens production via SCT-CPO from n-butane was
tested for a preliminary assessment of the economics of this process, since the market demand for propylene is increasing faster
than for ethylene [14].
2. Material and methods
2.1. Catalyst preparation
Commercial cordierite honeycombs with straight and parallel
channels of roughly square section (600 cpsi by NGK [22]) were
cut in the shape of disks of 25 mm diameter and 10 mm long;
the monoliths were washcoated by dip-coating in an aqueous slurry of alumina powder (3% La2O3-c-Al2O3, SCFa140-L3 Sasol, 140 m2/g) and pseudobohemite (Disperal, Sasol) to obtain a
nominal thickness of 40 lm (Fig. 1b). A LaMnO3 layer (30% w/w
of the alumina washcoat) was deposited by repeated cycles of
co-impregnation with an equimolar solution of the precursor salts
(La-nitrate and Mn-acetate) and calcination in air at 800 C for 3 h.
Finally, Pt (up to 4% w/w of the active layer, monolith substrate excluded) and Sn as dopant (Sn/Pt atomic ratio = 6) were added to the
structure by co-impregnation of the coated monoliths with a
H2PtCl6 and SnCl2 acid solution. More details on catalyst preparation and characterization can be found elsewhere [11,2123].

Monolith catalysts with identical nominal composition were employed in the CPO of ethane or n-butane.
2.2. Bench-scale CPO plant
The tests were performed in a bench-scale plant with ethane
capacity of 1000 Nl/h (roughly 15 times larger than previous lab
scale [10,11]) and n-butane capacity of 390 Nl/h, composed by ve
main zones:
1.
2.
3.
4.
5.

Feeding and Preheating zone.


Mixing zone.
Reaction zone.
Cooling zone.
Analysis zone.

The reactants ethane, nitrogen, oxygen and hydrogen were supplied by cylinders and their owrates controlled by mass ow controllers (Brooks, Bronkhorst). n-Butane owrate was measured and
controlled by a mass ow meter specic for liquids (Bronkhorst
mini CORI-FLOW).
The hydrocarbon feedstock, nitrogen and hydrogen were conveyed into a single line. Oxygen ew into a second, independent
line. Each line was equipped with an electric pre-heater. Mixing
of feedstock and oxidant streams was performed in a tube in
tube device, located on the top of the vertical steel vessel, featuring a thick internal refractory lining (tight t concentric ceramic
tubes) to limit heat loss in the reaction zone, whose inner diameter
was 26 mm (Fig. 1a). Three electrical resistances surrounded the
reaction zone and were switched on only during the light-off of
the self-sustained SCTCPO reactions.
The hot efuent from the catalytic reactor was transferred
through a water cooled line to heat exchangers and lters before
owing downstream through a back-pressure valve and being
ared.
A side-stream of the efuent was collected and analyzed by an
online GC (HP 7890, equipped with FID and TCD detectors) and a
microGC (Agilent Quad), calibrated to measure CO, CO2, N2, O2,
H2 and hydrocarbons up to C6.

Fig. 1. (a) Schematic of the bench scale SCTCPO mixing, reaction and cooling zones. (b) Optical microscopy and SEM-BSE images of the structured catalyst with PtSn/
LaMnO3/Al2O3 active washcoat layer on a cordierite honeycomb (600 cpsi).

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

Mass balances were calculated by adjusting the volumetric


ow-rate of the efuent to balance C atoms in and out and the
water ow to balance O atoms in and out. The error on N and H
atoms typically resulted in the range 3 vol.%.
Results are generally reported in terms of C-atom selectivity of
species j here dened as:

v j F out
j

Sj 
in
n F C n Hm  F out
C n Hm
where vj is the number of C atoms in species j, F out
its outlet molar
j


in
out
ow, F Cn Hm  F Cn Hm is the difference between inlet and outlet molar ows of the specic feedstock CnHm containing n atoms of C.
2.3. Operating conditions
Catalytic tests were generally performed at the operating pressure of 1.5 bara; in a few tests this was set at 2 bara. According to
previous reports [5,915,2325] hydrogen was fed to the reactor as
a sacricial fuel in order to maximize the selectivity to olens. The
inlet temperature of the pre-mixed stream was set at 250270 C
with ethane and 200250 C with n-butane, i.e. the maximum temperatures allowed by our experimental apparatus. Nitrogen was
used as an internal standard: its content in the feed stream (ca.
8 vol.% with ethane and 5 vol.% with n-butane) was kept as low
as possible according to the minimum stable ow rate of N2
achievable with the specic mass ow controllers employed in
each set of experiments.
Oxygen conversion in the CPO reactor was almost complete;
since in industrial units the presence of oxygen in the efuent cannot be tolerated, we performed specic tests to ascertain that the
eventual presence of residual O2 was caused by some lateral bypass between the honeycomb catalyst and the reaction tube. This
is a common issue with bench-scale CPO reactors, which has a direct negative impact on feed conversion and selectivity to olens.
The experimental campaign with ethane was performed with a
single catalyst sample that was on stream for a total of 550 h. During the rst 220 h the operating conditions were widely changed in
order to optimize the yield in C2 + C3 olens while minimizing the
consumption of reactants:

 O2 =C ratio
 H2 =O2 ratio

0:200:25 v=v
1:003:10 v=v

 Ethane load 300600

Nl=h:

These values resulted in space velocities ranging from 230,000


to 510,000 Nl/kg/h (corresponding to a maximum GHSV of
400,000 h1 referred to the volume of the monolith catalyst).
A life test was then performed for 320 h at xed inlet conditions:
O2/C = 0.22 v/v, H2/O2 = 1.53 v/v and GHSV = 360,000 Nl/kg/h.
With n-butane the main operating parameters were varied in
the following range:

O2 =C

0:150:27 v=v

H2 =O2

02:50

n  Butane load 150300

v=v
Nl=h:

The corresponding space velocities were comprised between


100,000 and 450,000 Nl/kg/h.
3. Results and discussion
3.1. CPO of ethane
Initial experiments, performed with ethane feedstock, examined the effects of process parameters on the overall process
performance.

475

The effects on ethane conversion and ethylene selectivity


played by: (i) O2/C and H2/O2 feed ratios, (ii) space velocity and
(iii) preheating temperature were strongly related: a single parameter variation determined an opposite effect on conversion and
selectivity; i.e. the increase of O2/C ratio and preheating temperature enhanced conversion and reduced selectivity, while the increase of the H2/O2 ratio and space velocity reduced conversion
but enhanced the selectivity towards ethylene.
As shown in Fig. 2ac, in the conversion-selectivity diagram the
data obtained with the PtSn/LaMnO3 catalyst in the bench scale
rig under optimized process conditions follow a single line and
are highly reproducible as already reported in previous works on
the same catalyst at smaller scale [10,11]. A maximum total Catom yield per pass to ethylene and propylene up to 60% (equivalent to 55.7% on a mass basis) was achieved for several operation
conditions. All these runs were carried out with H2 added to the
feed as sacricial fuel and we always observed an overall net production of hydrogen from the catalytic reactor and a C-selectivity
to propylene of ca. 11.5%.
A direct comparison (Fig. 2ac) of the products distribution obtained in the present study with previous experimental data at lab
scale [10,11] shows that the main species follow the same qualitative trends as a function of ethane conversion (i.e. process severity), whose increment is always accompanied by an increase of
C-atom selectivity to CO, CH4, C2H2 and CO2 and, consequently, a
reduction in C2H4 selectivity. However, with the bench-scale reactor a limited and constant loss of selectivity to ethylene was measured at any ethane conversion level. This effect was mainly
counterbalanced by a slightly higher selectivity to methane and
CO and a larger formation of C4C6 hydrocarbons (C-sel. = 35%),
whose selectivity was estimated 62% in the lab-scale reactor. At
the same time, identical C-selectivity to acetylene, CO2 and to valuable propylene (not shown) were measured from the two reactors
for every ethane conversion level. The overall catalytic performance of the bench-scale reactor with PtSn/LaMnO3 catalyst is
closer to the results obtained with the undoped (without Sn) catalyst at lab scale (Fig. 2ac). However, it is pointed out that similar
ethane conversion levels were obtained with less oxygen in the
feed (13%) in the bench scale reactor, thanks to its higher degree of adiabaticity, related to the higher capacity and to a more
favorable surface/volume ratio limiting heat loss. Since pure oxygen is a signicant cost component of the feedstream, the decrease
in oxygen employed per unit of C2H4 produced translates directly
into economic savings [4,5,10]. Moreover, since oxygen reacts with
hydrogen, lower amounts of oxygen lead to a decrease in hydrogen
consumed. As a consequence, a surplus of H2 was found in the
products, which was not always the case with previous reports
on PtSn and PtSn/LaMnO3 based catalysts evaluated at smaller
scale [1013,23].
The analysis of the relationship between ethane cracking conversion (C-atom ratio of all hydrocarbon products over all hydrocarbon products including unconverted feed [9]) and cracking
selectivity (C-atom selectivity based on all hydrocarbon products,
excluding unconverted feed and COx [9]) is reported in Fig. 3.
Cracking selectivity data converge in single narrow curves for each
hydrocarbon product, insensitive to the wide range of experimental process conditions and largely following the trends predicted
for homogeneous isothermal steam cracking of the residual C2H6
feed not converted to COx [11]. However CH4 formation is underpredicted by the simplied isothermal steam cracking model, as
also reported by Lange et al. who compared data from several catalysts and attributed this effect to the temperature spike at the entrance of the catalytic reactor [9]. This analysis provides evidence
that most olens are produced through homogeneous dehydrogenation and pyrolysis reactions driven by the heat generated by heterogeneous oxidation reactions [911,14]. The comparison

476

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

(a)

100

60 %

50 %

80 %

70 %

Pt/Perovskite
Pt-Sn/Perovskite
Pt-Sn/Perov. ENI

C-atom selectivity (%)

90
40 %

80

C2H4
70

60

C2H2

5
0

50

60

70

80

90

100

C 2 H 6 conversion (%)

(c)

C-atom selectivity (%)

25
20
15

CO

10

CO2

5
0
50

60

70

80

90

100

C2H6 conversion (%)

C-atom selectivity (%)

(b)

15

CH4
10

C2H2
0
50

60

70

80

90

100

C2H6 conversion (%)

Fig. 2. Process selectivity to (a) C2H4 and C2H2, (b) CO and CO2, (c) CH4 and C2H2 as a function of ethane conversion during the CPO reaction over PtSn/LaMnO3 catalyst in the
bench scale reactor (eni) as compared to previous results at lab scale [10,11]. Dashed lines represent iso-yield curves.

ene just formed. This circumstance points out the importance of


the design of the post catalytic and cooling section to effectively
quench the valuable and highly reactive products, an easier task
to achieve with a small quartz reactor than with a larger ceramic
one [11].

Cracking selectivity (%)

100
90
80
70
60
20

Pt-Sn/Perov. ENI
Pt-Sn/Perov. labscale
model 1000C

10
0
50

60

70

80

90

100

C2H6 Cracking conversion (%)


Fig. 3. Effect of the reactor size (eni bench scale vs. lab scale) on C-atom cracking
selectivity to ethylene, methane and acetylene as a function of ethane cracking
conversion over PtSn/LaMnO3 catalyst after normalization to exclude COx
products. Lines represent predictions by a purely homogeneous steam cracking
model in an isothermal plug ow reactor operated at 1000 C and 1.5 bara with
C2H6:H2O = 1:1 [11].

between lab-scale and bench-scale reactivity features in terms of


cracking selectivity and conversion (Fig. 3) conrmed a reduction
in ethylene selectivity in the latter case, which is due to a (slightly)
larger formation of methane, and mainly to heavier hydrocarbons
formed by undesired consecutive reactions consuming the ethyl-

3.1.1. Life test


As reported in Fig. 4a and b the PtSn/LaMnO3 system displayed
a substantial stability of catalytic performance over a life test lasting for 330 h at xed inlet conditions, with several start-up and
shut-down of the bench-scale rig. In particular, ethylene yield
was stable at the average value of 56.1 vol.% from 221 to
380 h.o.s., then it started a slow decrease settling down to
54.5 vol.% after 541 h.o.s. (end of the test), almost exclusively for
a corresponding reduction in ethane conversion. In fact, all main
products were formed with substantially unchanged selectivities:
CO selectivity, a measure of how inefciently the catalyst utilizes
the oxygen in the feed, passed from 11.1 to 11.6 vol.% at the end
of the run; this is a remarkable result as opposed to the increase
of some points percent previously observed for CO on PtSn on
a-alumina foams during only 24 h on stream [13].
The slight reduction in ethane conversion was accompanied by
a progressive increase in the pressure drops measured across the
reactor and the transfer line which increased from 50 mbar up to
77 mbar at the end of the test (Fig. 4a). Visual inspection of the
reactor revealed the presence of carbonaceous deposits in the lower section of the catalytic monolith and in the post catalyst zone
(heat shields and thermocouple sheath).

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

C2H6 conversion %

100

(a)

80
100
60
75
40
50

20
0
100
100

200

300

400

500

25
600

(b)

80

Selectivity %

125

60
15
10
5
0
100

200

300

400

500

600

time on stream, h
Fig. 4. Ethane CPO life test on PtSn/LaMnO3 catalyst: (a) fuel conversion and
pressure drops across the reactor and (b) process selectivities to main products as
function of time on stream at xed inlet conditions.

3.1.2. Catalyst investigation


Microscopic inspection of the catalytic monolith after the life
test (Fig. 5a and b) showed a partial collapse of the honeycomb
support structure in its top face but not of the oxide overlayers:
since the melting temperature of cordierite is higher than
1200 C, this is a clear indication of strong hot spot formation in
the oxidation zone of the catalyst. This was never observed during
operation of the same monolith catalyst in the lab scale reactor for
a total of 100 h.
In fact, in two occasions the threshold value for the O2/C ratio
was exceeded (for reasons not dependent on the process) causing
the peaking of the inlet temperature for the time needed (10 s)
for the Emergency Shut Down (EDS) to trip the Unit. Apart from that,
due to the concentrated heat generation [25], the upper section of
the CPO catalyst is always exposed to very harsh conditions, which
are expected to be even more severe in the bench-scale rig (no dilution, higher ow rates, lower heat loss). Therefore, the robustness of
the catalyst is a key feature for the industrial application.
XRD analysis detected that, where the cordierite structure was
damaged, some Si had migrated and formed a mixed phase with
the oxide overlayers (Al/La/Mn + Si). Where the cordierite honey-

477

comb structure was preserved, the active washcoat layer appeared


partially consumed and thinner close to the entrance (Fig. 5c).
Looking at the back face of the honeycomb catalyst, roughly 30%
of the channels appeared partially blocked by carbonaceous deposits, that built-up forming thick overlayers above the catalytic
washcoat (Fig. 5d). These deposits contributed to the increase in
pressure drops across the reactor. However, some adjacent catalytic channels in Fig. 5d appeared free from carbon, suggesting that
carbon formation was strongly related to gas ow (mal-)distribution in the inlet section, due to the collapse of the cordierite structure that caused longer residence times at high temperatures at
specic locations. From a different point of view, since the partial
occlusion of the honeycomb catalyst did not cause a signicant decrease of catalytic performance, it can be argued that the catalytic
system was efciently operated at contact times signicantly lower than the nominal ones (i.e. higher GHSV).
SEM-BSE analysis revealed that the residual active phase close
to the inlet section contained Pt clusters but was almost completely depleted in Sn, whereas Pt and Sn were still alloyed further
downstream along the channels of the monolith. In particular,
accordingly to a progressive consumption of oxygen along the catalytic reactor, SnO2 was recognized in the central zone of the channels whereas metallic Sn was detected in the last part.
Apart from the obvious issue of catalyst failure due to melting of
the ceramic substrate, which could be coped with a more heat
resistant support, high temperatures and high oxygen partial pressures caused a depletion of active phase due to volatilization of tin
(and possibly Pt), thus explaining the curves of Fig. 2ac approaching the results obtained on the undoped Pt/LaMnO3 systems. Similar conclusions were drawn by Bodke et al. [13], who observed a
much faster decay in catalytic performance (in terms of both ethane conversion and ethylene selectivity) of their PtSn system during only 20 h.o.s., and veried that a continuous addition of SnCl2
aqueous solution to the feed was able to restore activity by replenishing lost Sn in the catalyst.
These ndings suggest that the main differences observed in the
bench scale reactor are due to the development of a different temperature prole along the catalytic monolith, most likely characterized by a steeper hot spot on the surface close to the entrance.
High temperatures (>1000 C) are known to favor the formation
of methane over alkenes, and this was already identied as a drawback of adiabatic CPO reactors [9]. In fact CH4 and C2H2 formation
in gas phase are both enhanced by high temperatures (and indeed
they are used as an indication of process severity in the steam
cracking); however methane can be formed at comparable rates
also on the Pt surface [24,26]. In our case, the higher selectivity
to methane associated to the same levels of C2H2 (Figs. 2c and 3)
may be due to the occurrence of a hot spot on the inlet section
of the catalyst and not in the gas phase due to heat transfer limitations [14,25] speeding up the heterogeneous formation of CH4
from ethane.

Fig. 5. Optical microscopy and SEM-BSE images of the PtSn/LaMnO3 monolith catalyst after the life test of ethane CPO. (a and b) Partially melted cordierite structure close to
the top inlet section and (c) loss/thinning of active washcoat. (d) Build-up of carbon overlayer above the active phase causing the partial occlusion of some channels in the
outlet section of the catalyst.

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

3.2. CPO of n-butane


Fig. 6 compares the results of two CPO tests performed with
ethane and n-butane under the same operating conditions. The
higher reactivity of n-butane with respect to ethane, as well as
the higher partial pressure of oxygen when operating with a heavier alkane at the same O2/C feed ratio, determined a signicantly
higher process severity with fuel conversion reaching 91.2% for nbutane vs. 76.3% for ethane. As already reported [4,15] the CPO of
n-butane produced less ethylene, which was partly compensated
by the formation of propylene and butenes. The relevance of the
cracking and condensation reactions producing methane and
C4 + hydrocarbons increased. On the other hand, total COx selectivity did not change signicantly, but CO/CO2 ratio decreased with nbutane feed. Since oxygenates represent a net loss of feedstock, low
CO/CO2 ratios are preferred because they imply a more advantageous use of the oxygen feed to produce the heat required to sustain the endothermic reactions leading to olens formation.
Fig. 7a summarizes in the conversion-selectivity plane the datasets collected in a wide range of conditions. Yields in light olens
(ethylene + propylene) can easily exceed 50% on a carbon atom basis, which is considered the minimum acceptable value for the economics of the process. The degree of feedstock conversion (i.e.
process severity) controlled the selectivity trends for all the products. Ethylene selectivity progressively increased with conversion
since C2H4 can be formed either via primary or secondary cracking
reactions, involving butenes and propylene. Accordingly, butenes
and propylene selectivities were adversely affected by the increase
in fuel conversion. For conversion values > 90%, propylene decrease
was no longer compensated by ethylene increase, hence the total
selectivity to C2 + C3 olens dropped. In fact methane and acetylene were formed at a higher extent following the same trend already observed with the CPO of ethane .
The maximum C-atom yield of C2 + C3 olens was 55.6% corresponding to 53.7% on a mass basis, which increases up to 55.1 wt.%

100%
90%
80%
70%
60%
50%
40%
30%
20%
10%
0%

C2H6

n-C4H10

HCn>4

C4H10

C4H8

C3H6

C3H8

C2H2

C2H6

C2H4

CO2

CH4

CO

FS Conv.

C2H6
n-C4H10

Fig. 6. Comparison of fuel conversion and C-atom selectivities in the SCTCPO of


C2H6 and nC4H10 over PtSn/LaMnO3 catalyst operated at comparable inlet
conditions. O2/C = 0.2 v/v, H2/O2 = 1.0 v/v, space velocity 330,000 Nl/kg/h and
preheating temperature 250 C.

80

50%
40%

C-Selectivity %

In order to reduce the extent of the hot spot formation, specic


strategies for the optimization of the thermal behavior and heat
management of the reactor have been proposed and could be considered [25 and ref. therein]. We mention the variation of channel
opening [22] or support morphology (foams, honeycombs, microchannels in stacked metal plates [20,27]), the enhancement of
the thermal conductivity of monolith support [20,28], and the
reduction of the catalytic bed aspect ratio L/D to increase back heat
dispersion [25].

60%

70%

(a)

C2 + C3 Olefins

60
30%

C2H4

40
20%

C3H6
20

10%

0
50
10

molar ratios

478

CH4
C4H8

C2H2
60

70

80

90

100

C4H10 Conversion %
(b)

C2H4 / C3H6

6
4
2
0
50

CH4 / C3H6
60

70

80

90

100

C4H10 Conversion %
Fig. 7. Results of n-butane CPO over PtSn/LaMnO3 monolith catalyst. (a) C-atom
selectivities to C3H6, C2H4, C2H2 and CH4 (dashed lines represent iso-yield curves).
(b) C2H4/C3H6 and CH4/C3H6 molar ratios as a function of C4H10 conversion.

including butenes: C-selectivity to ethylene and propylene were


respectively 46.3% and 14.5% at 91.3% conversion. A maximum Catom yield of propylene around 18% was achieved for n-butane
conversions between 70% and 80%.
At low process severity ethylene and propylene were formed in
a molar ratio around 2 (Fig. 7b) whereas this value rapidly increased for conversion values above 80%. A similar trend was observed for the methane to propylene ratio, which was initially
close to 1, conrming that the valuable propylene is overcracked to smaller molecules at higher n-butane conversions.
Fig. 8ad compares the effect of the O2/C ratio on conversion
and products selectivity at different values of H2/O2 ratio. Butane
conversion was directly controlled by the O2/C ratio (Fig. 8a), while
H2/O2 ratio showed almost no effect.
Ethylene selectivity was enhanced by increasing O2/C, while
propylene was adversely affected by this parameter (Fig. 8b). An
important difference with the case of ethane CPO is related to
the effect of the sacricial fuel: H2 addition had no direct impact
on propylene formation curve, whereas it enhanced ethylene selectivity up to H2/O2 = 1. Similar results were found during the CPO of
propane over the same catalyst [10].
In fact, H2 addition raised the operating temperature on the catalyst since hydrogen has a higher net heat of combustion per mole
of O2 [10] and it is preferentially oxidized instead of the hydrocarbon feed [14], as demonstrated by the reduction observed in CO
and CO2 selectivities (Fig. 8d). However propylene formation does
not appear to be limited by the selectivity of the catalytic oxidation
reactions but only by the low selectivity of the homogeneous
cracking reaction.
Methane selectivity (Fig. 8c) increased with both the oxygen
and hydrogen content in the feed due to higher operating temperatures and faster CAC cleavage in the presence of larger amounts of
H2. C4+ selectivity (Fig. 8c) was slightly affected by O2/C and rather
insensitive to hydrogen addition.

479

H2/O2=0.0
H2/O2=0.9

70

H2/O2=1.0

60

H2/O2=2.0
H2/O2=3.0

50

C-selectivity %

50
40

0.14

0.16

0.18

0.20

0.22

0.24

0.26

(b)

30

2
in

O2in

20

H2 /

10

Fig. 9. Moles of H2 in the outlet stream per mole of O2 fed vs. the feed H2/O2 ratio
during the CPO of ethane and n-butane over PtSn/LaMnO3 catalysts. Points below
the dashed line indicate a net consumption of H2 throughout the catalytic reactor.

0.14

0.16

0.18

20

C-selectivity %

O2/C

0
0.20

0.22

0.24

0.26

O2/C

(c)
15
10
5
0
0.14

C-selectivity %

3
in

80

/ O2

90

(a)

out

100

H2

n-C4H10 conv. %

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

14
12
10
8
6
4
2
0

0.16

0.18

0.22

0.24

0.26

0.22

0.24

0.26

O2/C

(d)

0.14

0.20

0.16

0.18

0.20

O2/C
Fig. 8. Effect of O2/C and H2/O2 feed ratios on fuel conversion and C-atom selectivity
to the main products during the CPO of n-butane over PtSn/LaMnO3 monolith
catalyst. Data at H2/O2 = 0.9 were obtained at P = 2 bara. Arrows indicate trends for
increasing H2/O2 values.

Regarding the effect of pressure on the catalytic features, moving from 1.5 to 2 bara (data set at H2/O2 = 0.9 in Fig. 8) we observed
a small reduction in the process selectivity to ethylene whereas
fuel conversion and selectivity to propylene were almost
unaffected.
A net hydrogen production was identied as a key condition to
achieve favorable overall economics of the process. As seen in
Fig. 9, for ethane to ethylene the process produces more H2 than
fed for all the operating condition explored with H2/O2 up to 3.
On the contrary, for n-butane to olens the CPO process produces
a H2 surplus only at low H2/O2 ratios, namely below 0.5 v/v. Similar
qualitative results were reported by Bodke et al. [29] but the crossover points depend strongly on the type of catalyst [21] as well as
on reactor design and process conditions such as the fuel to oxygen
ratio and preheat [29]. In fact, a further peculiar feature of the Pt/
LaMnO3 catalyst is the possibility to advantageously use CO or
mixtures of CO and H2 as sacricial fuels to avoid the eventual
H2 unbalance [11,23].

However, since in the case of n-butane CPO the positive effect of


sacricial H2 addition was modest and only limited to ethylene
selectivity, an upper limit to the H2/O2 feed ratio at 0.5 did not preclude the achievement of high olen yields.
Though the identication of the reaction mechanism was not
the primary scope of these tests, the data-sets collected in a wide
range of operating conditions have much improved our understanding on the behavior of the system.
With ethane as well as with n-butane, the sacricial fuel fed together with the feedstock is preferentially oxidized and produces
heat for following pyrolysis reactions. From this point of view,
the superior performance of the PtSn/LaMnO3 catalyst can be explained considering the presence of an additional active phase (the
perovskite LaMnO3) able to selectively promote the combustion of
the sacricial fuel without catalyzing undesired side reactions such
as steam reforming or hydrogenolysis consuming both reactant
and/or desired products [14]. In other words, an effective SCT
CPO catalyst for olens production must produce the maximum
heat release with the minimum oxygen consumption [9]. Though
we cannot exclude a partial contribution to olens production
via heterogeneous dehydrogenation or oxidative dehydrogenation
reactions, the strong dependence of all hydrocarbon products on
feedstock conversion (directly related to the operating temperature [15,17]) is a clear indication that homogeneous reactions plays
a major role. As a general rule, all the parameters that enhance
conversion induce an increase in ethylene, acetylene, methane,
C4+ hydrocarbons and a corresponding decrease to butenes and
propylene. Since C3 and C4 olens are destroyed through subsequent reactions (secondary cracking), it is likely that an optimized
plant design, with strong integration between reaction and cooling
zones, will reduce the incidence of this undesired effect.
4. Conclusions
The experimental campaigns performed so far have substantially conrmed, on a larger scale, that the SCTCPO technology
over PtSn/LaMnO3 monolith catalysts originally proposed for ethylene production from ethane, can be successfully applied to heavier alkanes feedstock.
Ethylene yields exceeding 55 wt.% were achieved from ethane
and, for the rst time, the reactivity performance were demonstrated for more than 500 h.o.s. With n-butane feedstock, ethylene + propylene yields approached 54 wt.% in a wide range of
experimental conditions.

480

L. Basini et al. / Chemical Engineering Journal 207208 (2012) 473480

Though the bench-scale unit was specically designed for


obtaining high feedstock conversion and high olens yield, wide
changes in the operating conditions conrmed that olens are
mainly produced via gas phase reactions; however a proper choice
of active phase and catalytic reactor design are of paramount
importance for achieving and maintaining high performance due
to the strong interplay between hetero-homogeneous and exoendothermic reactions.
The understanding of the prevailing mechanisms leading to olens formation indicated the way for maximizing light olens yield.
However, high yields of light olens could not be sufcient to
achieve favorable economics for the industrial application of the
technology. Other features are as much as important, such as the
net production of hydrogen, the minimization of oxygen consumption and of by-products formation. From this point on, it is of paramount importance that technical-economic evaluations support
and address the experimental work.
In conclusions, we deem that SCTCPO technology has a great
potential for the production of light olens; however, this novel
technology will be ready for the industrial application only after
completing an extensive experimental work in the range of operating conditions identied as the most promising by the technicaleconomic evaluations and after improving the reliability of the catalyst and the reactor design, particularly with respect to heat
management.
Those issues, which have been partly addressed in this work,
are currently under investigation.
Acknowledgments
The authors wish to thank Pierino Visioli and Luigi Romano, for
their patience and precision in performing the experimental work,
Eleonora Di Paola and Danila Ghisletti, for their extensive work
respectively on optical microscopy/SEM-BSE and XRD characterization of catalysts. Special thanks go to Giovanni Forlani, who
realized most of the technical devices utilized in this work, with
unique passion and devotion.
References
[1] C. Eng, M. Tallman, Consider new catalytic routes for olens production,
Hydrocarbon Process. 87 (4) (2008) 95101.
[2] D. Sanlippo, I. Miracca, Dehydrogenation of parafns: synergies between
catalyst design and reactor engineering, Catal. Today 111 (2006) 133139.
[3] F. Cavani, N. Ballarini, A. Cericola, Oxidative dehydrogenation of ethane and
propane: how far from commercial implementation?, Catal Today 127 (2007)
113131.
[4] Chemsystems Nexant, Production of Olens via Oxidative Dehydrogenation of
Light Parafns at Short Contact Times PERP, Report 03/04S2, 2004.
[5] S. Bharadwaj, J. Maj, J. Siddall, M. Bearden, C. Murchison, G. Lazaruk,
Autothermal process for the production of olens, Dow Global Technologies,
US patent 6,566,573, 2003.
[6] L. Basini, A. Guarinoni, D. Sanlippo, Catalytic system for the production of
olens, Saipem, US Patent 7,829,759, 2010.

[7] A. Bodke, D. Olschki, L. Schmidt, E. Ranzi, High selectivities to ethylene by


partial oxidation of ethane, Science 285 (1999) 712715.
[8] G. Groppi, A. Beretta, E. Tronconi, Monolithic catalysts for gas-phase syntheses
of chemicals, in: A. Cybulski, J. Moulijn (Eds.), Structured Catalysts and
Reactors, second ed., CRC Press Taylor & Francis Group, 2006, pp. 243310.
[9] J.-P. Lange, R. Schoonebeek, P. Mercera, F. van Breukelen, Oxycracking of
hydrocarbons: chemistry, technology and economic potential, Appl. Catal. A
283 (2005) 243253.
[10] S. Cimino, F. Dons, G. Russo, D. Sanlippo, Olens production by catalytic
partial oxidation of ethane and propane over Pt/LaMnO3 catalyst, Catal. Today
157 (2010) 310314.
[11] S. Cimino, F. Dons, G. Russo, D. Sanlippo, Optimization of ethylene
production via catalytic partial oxidation of ethane on PtLaMnO3 catalyst,
Catal. Lett. 122 (2008) 228237.
[12] S. Hakonsen, J. Walmsley, A. Holmen, Ethene production by oxidative
dehydrogenation of ethane at short contact times over PtSn coated
monoliths, Appl. Catal. A 378 (2010) 110.
[13] A. Bodke, D. Henning, L. Schmidt, S. Bharadwaj, J. Maj, J. Siddall, Oxidative
dehydrogenation of ethane at millisecond contact times: effect of H2 addition,
J. Catal. 191 (2000) 6274.
[14] B. Michael, D. Nare, L. Schmidt, Catalytic partial oxidation of ethane to
ethylene and syngas over Rh and Pt coated monoliths: Spatial proles of
temperature and composition, Chem. Eng. Sci. 65 (2010) 38933902.
[15] M. Huff, L.D. Schmidt, Production of olens by oxidative dehydrogenation of
propane and butane over monoliths at short contact times, J. Catal. 149 (1994)
127141.
[16] S. Pavlova, V. Sadykov, Yu. Frolova, N. Saputina, P. Vedenikin, I. Zolotarskii, V.
Kuzmin, The effect of the catalytic layer design on oxidative dehydrogenation
of propane over monoliths at short contact times, Chem. Eng. J. 91 (2003) 227
234.
[17] L. Liebmann, L. Schmidt, Oxidative dehydrogenation of isobutane at short
contact time, Appl. Catal. A 179 (1999) 93106.
[18] S. Mulla, O. Buyevskaya, M. Baerns, Autothermal oxidative dehydrogenation of
ethane to ethylene using SrxLa1.0Nd1.0Oy catalysts as ignitors, J. Catal. 197
(2001) 4348.
[19] O. Buyevskaya, M. Baerns, Oxidative functionalization of ethane and propane,
in: J.J. Spivey (Ed.), Catalysis, A Specialist Periodical Report, vol. 16, Royal
Society of Chemistry, Cambridge-UK, 2002, pp. 155197.
[20] B. Yang, T. Yuschak, T. Mazanec, A. Tonkovich, S. Perry, Multi-scale modeling of
microstructured reactors for the oxidative dehydrogenation of ethane to
ethylene, Chem. Eng. J. 135S (2008) S147S152.
[21] F. Dons, S. Cimino, R. Pirone, G. Russo, D. Sanlippo, Crossing the
breakthrough line of ethylene production by short contact time catalytic
partial oxidation, Catal. Today 106 (2005) 7276.
[22] F. Dons, S. Cimino, A. Di Benedetto, R. Pirone, G. Russo, The effect of support
morphology on the reaction of oxidative dehydrogenation of ethane to
ethylene at short contact times, Catal. Today 105 (2005) 551559.
[23] F. Dons, S. Cimino, R. Pirone, G. Russo, Autothermal oxidative
dehydrogenation of ethane on LaMnO3 and Pt-based monoliths: H2 and CO
addition, Ind. Eng. Chem. Res. 44 (2005) 285295.
[24] F. Dons, K. Williams, L. Schmidt, A multistep surface mechanism for ethane
oxidative dehydrogenation on Pt- and Pt/Sn-coated monoliths, Ind. Eng. Chem.
Res. 44 (2005) 34533470.
[25] D. Livio, A. Donazzi, A. Beretta, G. Groppi, P. Forzatti, Experimental and
modeling analysis of the thermal behavior of an autothermal C3H8 catalytic
partial oxidation reformer, Ind. Eng. Chem. Res. 51 (2012) 75737583.
[26] S. Cimino, L. Lisi, G. Russo, Effect of sulphur during the catalytic partial
oxidation of ethane over Rh and Pt honeycomb catalysts, Int. J. Hydrogen
Energy 37 (2012) 1068010689.
[27] R. Long, F. Daly, A. Glass, T. LaPlante, T. Yuschak, T. Mazanec, Superior Pt-alloy
catalysts for oxidative dehydrogenation of ethane to ethylene in microchannel
reactors, in: Proc. of 20th NAM, Houston, 2007.
[28] G. Landi, P.S. Barbato, S. Cimino, L. Lisi, G. Russo, Fuel-rich methane
combustion over Rh-LaMnO3 honeycomb catalysts, Catal. Today 155 (2010)
2734.
[29] A. Bodke, D. Henning, L.D. Schmidt, A comparison of H2 addition to 3 ms partial
oxidation reactions, Catal. Today 61 (2000) 6572.

Potrebbero piacerti anche