You are on page 1of 23

Applied Catalysis A: General 221 (2001) 397–419

Dehydrogenation and oxydehydrogenation of paraffins to olefins


M.M. Bhasin a,∗ , J.H. McCain a , B.V. Vora b , T. Imai b , P.R. Pujadó b
a Union Carbide & UOP LLC, Subsidiary of Dow Chemical Co., P.O. Box 8361, 25303 South Charleston, WV, USA
b UOP LLC, Subsidiary of Dow Chemical Co., P.O. Box 8361, 25303 South Charleston, WV, USA

Abstract
Catalytic paraffin dehydrogenation for the production of olefins has been in commercial use since the late 1930s, while
catalytic paraffin oxydehydrogenation for olefin production has not yet been commercialized. However, there are some
interesting recent developments worthy of further research and development.
During World War II, catalytic dehydrogenation of butanes over a chromia-alumina catalyst was practiced for the production
of butenes that were then dimerized to octenes and hydrogenated to octanes to yield high-octane aviation fuel. Dehydrogenation
employs chromia-alumina catalysts and, more recently, platinum or modified platinum catalysts. Important aspects in dehydro-
genation entail approaching equilibrium or near-equilibrium conversions while minimizing side reactions and coke formation.
Commercial processes for the catalytic dehydrogenation of propane and butanes attain per-pass conversions in the range of
30–60%, while the catalytic dehydrogenation of C10 –C14 paraffins typically operates at conversion levels of 10–20%. In the
year 2000, nearly 7 million metric tons of C3 –C4 olefins and 2 million metric tons of C10 –C14 range olefins were produced
via catalytic dehydrogenation.
Oxydehydrogenation employs catalysts containing vanadium and, more recently, platinum. Oxydehydrogenation at
∼1000 ◦ C and very short residence time over Pt and Pt-Sn catalysts can produce ethylene in higher yields than in steam
cracking. However, there are a number of issues related to safety and process upsets that need to be addressed. Important
objectives in oxydehydrogenation are attaining high selectivity to olefins with high conversion of paraffin and minimizing
potentially dangerous mixtures of paraffin and oxidant. More recently, the use of carbon dioxide as an oxidant for ethane
conversion to ethylene has been investigated as a potential way to reduce the negative impact of dangerous oxidant–paraffin
mixtures and to achieve higher selectivity.
While catalytic dehydrogenation reflects a relatively mature and well-established technology, oxydehydrogenation can in
many respects be characterized as still being in its infancy. Oxydehydrogenation, however, offers substantial thermodynamic
advantages and is an area of active research in many fronts. © 2001 Elsevier Science B.V. All rights reserved.
Keywords: Paraffin dehydrogenation; Olefin; Chromia-alumina catalyst; Paraffin oxydehydrogenation; Noble metal catalysts

1. Historical overview and chromia-alumina World War II, catalytic dehydrogenation of butanes
catalysts over a chromia-alumina catalyst was practiced for
the production of butenes, which were then dimer-
Paraffin dehydrogenation for the production of ized to octenes and hydrogenated to octanes to yield
olefins has been in use since the late 1930s. During high-octane aviation fuel.
Dehydrogenation of butanes over a chromia-alumina
∗ Corresponding author. Tel.: +1-304-747-4910; catalyst was first developed and commercialized at
fax: +1-304-747-5430. Leuna in Germany and was also independently de-
E-mail address: bhasinmm@dow.com (M.M. Bhasin). veloped by UOP (then Universal Oil Products) in the

0926-860X/01/$ – see front matter © 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 1 ) 0 0 8 1 6 - X
398 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

United States, together with ICI in England. The first process. A similar oxydehydrogenation approach for
UOP-designed plant came on stream in Billingham, the production of butadiene was also practiced by
England, in 1940 and was soon followed by two other Phillips Petroleum [3].
units in Heysham, England, in 1941 [1]. The primary Large quantities of butadiene have become available
purpose of this butane dehydrogenation was to pro- over the past 30 years, mostly as a by-product from the
duce butenes, which were then dimerized to octenes thermal cracking of naphtha and other heavy hydrocar-
using solid phosphoric acid catalysts discovered by bons. This market shift has resulted in the shutdown
Schaad and Ipatieff [2]. of all on-purpose catalytic dehydrogenation units for
Other companies soon followed these pioneering butadiene production in North America, western Eu-
efforts. For example, Phillips Petroleum built a multi- rope, and the far East.
tubular dehydrogenation reactor near Borger, TX, in In the late 1980s, the application of chromia-alumina
1943 [1]. However, the most significant development catalysts was extended by Houdry to the dehydrogena-
was made by Houdry using dehydrogenation at less tion of propane to propylene and isobutane to isobuty-
than atmospheric pressure for higher per-pass conver- lene. The new process application called CatofinTM
sions. This process, which came on stream toward the [4,5] operates on the same cyclic principle as in the
end of World War II, was also used for the production former Catadiene process. As of late 2000, a total of
of butenes. After the war, Houdry further developed eight Catofin units exist for the production of isobuty-
and commercialized the chromia-alumina dehydro- lene (including two converted older Catadiene units)
genation system and extended it to the production of with an aggregate capacity of about 2.8 million metric
butadiene in what became known as the CatadieneTM tons per annum (MTA) isobutylene. In addition, two
process [3]. Other companies, including Shell, Gulf, Catofin units were built for the production of propy-
and Dow, also practiced similar dehydrogenation lene, but it is understood that only one is operational
technologies. with a nameplate dehydrogenation capacity of about
In the dehydrogenation process using chromia- 250,000 MTA propylene, but usually operating only
alumina catalysts, the catalyst is contained in a fixed on a seasonal basis. Plans for another 450,000 MTA
shallow bed located inside a reactor that may be either Catofin propane dehydrogenation unit in Saudi Ara-
a sphere, a squat vertical cylinder, or a horizontal cylin- bia have also been announced. The Catofin process
der. The actual design reflects a compromise between technology is currently owned by Süd-Chemie and is
gas flow distribution across a large cross-sectional offered for license by ABB Lummus.
area and the need to maintain a low pressure drop. A In 1959, an alternative chromia-alumina catalytic
significant amount of coke is deposited on the catalyst dehydrogenation process was developed in the for-
during the dehydrogenation step, therefore, a number mer Soviet Union. This process avoided the use of the
of reactors are used in parallel—some for dehydro- cyclic operation by using a fluidized bed reactor con-
genation while the rest are being purged or regen- figuration similar to the fluidized catalytic cracking
erated. The dehydrogenation reactions are strongly (FCC) process used in refineries [6]. However, back-
endothermic, and the heat is provided, at least in mixing common to dense fluidized bed operations re-
part, by the sensible heat stored in the catalyst bed sults in poor selectivity and increases the formation
during regeneration (carbon burn); additional heat is of heavies, sometimes called “green oils”. Circulating
provided by direct fuel combustion and also by heat regenerated catalyst is used to provide the heat of re-
released in the chromium redox cycle. The length of action in the riser and spent catalyst is reheated by
the total reactor cycle is limited by the amount of heat carbon burn in the regenerator. During the 1990s, a
available, and can be as short as 10–20 min. large scale fluid bed isobutane dehydrogenation unit
The Houdry Catadiene process was used exten- for about 450,000 MTA isobutylene was commercial-
sively for the production of butadiene, either by ized by Snamprogetti in Saudi Arabia based on tech-
itself (n-butane to butadiene) or in conjunction with nology from Yarsintez in Russia [6], but it is under-
catalytic oxydehydrogenation of n-butene to butadi- stood that this unit has only operated at lower than de-
ene. The latter was commercialized by the Petro-Tex sign capacity. Recent literature articles report further
Chemical Corp. [3] and was called the Oxo-DTM improvements by Snamprogetti [7,8].
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 399

2. Noble metal dehydrogenation catalysts

A different approach to catalytic dehydrogenation


was first introduced in the mid-1960s for the sup-
ply of long-chain linear olefins for the production of
biodegradable detergents.
Synthetic detergents, based on the use of branched
alkylbenzene sulfonates derived from propylene
tetramer and benzene, had been introduced in the
1940s. By the early 1960s, however, it became appar-
ent that branched dodecylbenzene-based detergents,
though very active and offering excellent perfor-
mance characteristics, did not biodegrade readily and
were accumulating in the environment. The need for
biodegradable detergents prompted the development
of catalytic dehydrogenation of long-chain linear
paraffins to linear olefins.
The work on catalytic reforming with noble metal Fig. 1. Propane dehydrogenation equilibrium at 1.00 atm abs. pres-
sure.
(Pt) catalysts done in the 1940s by Haensel clearly
demonstrated that Pt-based catalysts had high activity
for the dehydrogenation of paraffins to the correspond- where xe is the equilibrium conversion, P the total ab-
ing olefins [9]. In the 1960s, Bloch [10] further ex- solute pressure and Kp is the equilibrium constant for
tended this thinking by developing Pt-based catalysts the dehydrogenation reaction. The equilibrium con-
that could selectively dehydrogenate long-chain linear stant can be easily calculated from Gibbs free energies
paraffins to the corresponding internal mono-olefins as tabulated in the API 44 report or in similar sources
with high activity and stability and with minimum of thermodynamic data. Figs. 1 and 2 illustrate the
cracking. This was the basis for the UOP PacolTM pro- equilibrium conversion levels that can be obtained for
cess for the production of linear olefins for the man- propane at 1 and 0.23 atm abs. (175 Torr), respectively.
ufacture of biodegradable detergents [11]. In 1999,
there were more than 30 commercial Pt-catalyzed de-
hydrogenation units in operation for the manufacture
of detergent alkylate.
Long-chain paraffins are both valuable and highly
prone to cracking. Therefore, in order to maintain high
selectivity and yield, it is necessary to operate at rela-
tively mild conditions, typically below 500 ◦ C, and at
relatively low per-pass conversions. While this is eco-
nomical for the production of heavy linear olefins, it
is not for the production of light olefins.
Paraffin dehydrogenation is an endothermic reaction
that is limited by chemical equilibrium and, accord-
ing to Le Chatelier’s principle, higher conversion will
require either higher temperatures or lower pressures.
In a somewhat abbreviated form for the production of
mono-olefins, this can be expressed as follows:

Kp
xe2 = Fig. 2. Propane dehydrogenation equilibrium at 0.23 atm abs. pres-
Kp + P sure.
400 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

temperatures, but side reactions, coke formation, and


catalyst deactivation are also accelerated. Thus, ex-
trapolation directly from heavy olefins to light olefins
cannot be done without taking other factors into
consideration.
Production of light olefins by the catalytic dehydro-
genation of light paraffins must be able to maintain
reasonable per-pass conversion levels and high olefin
selectivity. Very importantly, it must be able to pro-
duce olefins in high yields over long periods of time
without shutdowns.
In the early 1970s, UOP introduced continuous cata-
lyst regeneration (CCR) technology that enabled noble
metal catalysts to remain at their most desirable stable
Fig. 3. Equilibrium constants for n-paraffin dehydrogenation at activity for several years without having to shut down
500 ◦ C.
the reactor for catalyst regeneration. The combination
of noble metal catalysts operating at high severity in
The equilibrium constant for paraffin dehydrogena- conjunction with CCR technology made it possible to
tion increases significantly as the carbon number in- design, build, and economically operate large catalytic
creases. Fig. 3 shows the equilibrium constant for the dehydrogenation units that can produce light olefins,
dehydrogenation of n-paraffins ranging from ethane in particular, propylene and isobutylene, at high selec-
to pentadecane [12]. Fig. 4 shows the temperatures tivities while still operating at superatmospheric pres-
required to achieve 10–40% equilibrium conversion sures. This technology is known as the UOP OleflexTM
based on these equilibrium constants. Fig. 4 indicates process. As of late 2000, there were four propane de-
that the temperature required for the dehydrogenation hydrogenation units, five isobutane dehydrogenation
of light paraffins is much higher than for heavy paraf- units, and one combined propane/isobutane dehydro-
fins. For 40% conversion, for example, the dehydro- genation unit of this type in commercial operation,
genation of propane requires a temperature of at least with an aggregate operating capacity of 900,000 MTA
about 580 ◦ C, while dodecane can be theoretically polymer grade propylene and 2.3 million MTA
dehydrogenated to the same extent at only 450 ◦ C. isobutylene. In addition, another propane dehydro-
The equilibrium conversion increases at higher genation unit for 350,000 MTA polymer grade propy-
lene was under design and construction.
The world propylene production capacity, based on
the use of catalytic dehydrogenation of propane has
increased steadily over the past 10 years [13] and is ex-
pected to grow even further under the right economic
conditions relative to the availability of propane; on
the other hand, environmental concerns on the use of
MTBE are expected to adversely impact the future ex-
pansion of isobutane dehydrogenation applications.
Although production of ethylene via catalytic dehy-
drogenation over Pt catalysts is very selective (about
95%), extension of this dehydrogenation technology
to ethane has not taken place due to the need for even
more severe operating conditions; higher temperatures
and lower pressures. Such conditions cause excessive
Fig. 4. Temperatures required to achieve 10 and 40% conversion coking of the catalyst or require costlier operation
of C2 –C15 n-paraffins at 1 atm. under vacuum.
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 401

Fig. 5. Reactions by platinum and acid sites in light paraffin dehydrogenation with unmodified catalyst.

Practically, all existing catalytic dehydrogenation 3. Process chemistry


capacity based on Pt catalysts is based on the Oleflex
process with CCR; however, there are also two smaller The main reaction in catalytic dehydrogenation is
units for isobutane dehydrogenation for 118,000 and the formation of mono-olefins from the correspond-
13,000 MTA isobutylene, respectively, both based on ing feed paraffin. Other reactions include consecutive
the STAR technology developed by Phillips Petroleum and side reactions. The reaction pathways involved in
and derived from their earlier multitubular reactor de- heavy paraffin dehydrogenation (e.g. detergent-range
sign experience. This reactor design resembles a typ- C10 –C14 n-paraffins) are more complicated than those
ical steam reformer that is operated until the catalyst in light paraffin dehydrogenation (e.g. propane and
deactivates as a result of coke deposition; banks of isobutane). The main difference in reaction pathways
tubes are sequentially taken out of service for cata- is that a significant amount of cyclic compounds can
lyst regeneration. The STAR technology is currently form via dehydrocyclization from heavy paraffins;
owned and licensed by Krupp–Uhde. this is not the case for light paraffins. Figs. 5 and 6

Fig. 6. Reactions by platinum and acid sites in heavy paraffin dehydrogenation with unmodified catalyst.
402 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

illustrate possible reactions that take place on Pt and The reaction of olefins on platinum is faster than
acid (A) sites, respectively, in the dehydrogenation of that of paraffins, because olefins interact with platinum
light and heavy paraffins when the catalyst is not se- more strongly than do paraffins. The role of platinum
lective, e.g. unmodified platinum catalysts supported modifiers is to weaken the platinum–olefin interaction
on alumina. selectively without affecting the platinum–paraffin in-
The consecutive reactions, the dehydrogenation of teraction. Arsenic, tin, germanium, lead, bismuth are
mono-olefins to diolefins and triolefins, are catalyzed among metals reported as platinum activity modifiers.
on the same active sites as the dehydrogenation of The consecutive dehydrogenation rate of mono-olefins
paraffins to mono-olefins. The consecutive reactions and diolefins is decreased by this modification without
that form triolefins, aromatics, dimers, and polymers lowering the rate of paraffin dehydrogenation signifi-
must be suppressed kinetically or by catalyst modifi- cantly. The modifier also improves the stability against
cations. fouling by heavy carbonaceous materials.
Platinum is a highly active catalytic element and is
not required in large quantities to catalyze the reaction
4. Role of catalysts and supports when it is dispersed on a high surface-area support.
The high dispersion is also necessary to achieve high
The discussion in this section pertains to alumina- selectivity to dehydrogenation relative to undesirable
supported platinum catalysts. The work by Poole and side reactions, such as cracking.
coworkers [14,15] provides an extensive review of The typical high surface area alumina supports em-
chromia-alumina catalysts. ployed have acidic sites that accelerate skeletal isomer-
The key role of dehydrogenation catalysts is to ization, cracking, oligomerization, and polymerization
accelerate the main reaction while controlling other of olefinic materials, and enhance “coke” formation.
reactions. Unmodified alumina-supported platinum Alkali or alkaline earth metals assist in the control of
catalysts are highly active but are not selective to de- the acidity. Also, ␣-alumina supports that have essen-
hydrogenation. Various by-products, as indicated in tially no acidity can be utilized; however, the challenge
Figs. 5 and 6, can also form. In addition, the catalyst is to obtain high dispersion of platinum on such very
rapidly deactivates due to fouling by heavy carbona- low surface area supports. Therefore, acidity must be
ceous materials. Therefore, the properties of platinum eliminated by using suitable modifiers.
and the alumina support need to be modified to sup- Modified catalysts possess high activity and high
press the formation of by-products and to increase selectivity to mono-olefins. The major by-products
catalytic stability. are diolefins that can be controlled kinetically. Coke

Fig. 7. Paraffin dehydrogenation on modified Pt catalyst.


M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 403

formation is also suppressed and, therefore, stabil- ratio (R). Therefore, Eq. (3) can be rewritten as
ity is greatly improved. Over modified catalysts,
d(sx) f2 [ki (T ), Ki (T ), x, P , R]
the major reaction pathways for both light and = (4)
heavy paraffin dehydrogenation systems are simpler dx f1 [ki (T ), Ki (T ), x, P , R]
(Fig. 7).
As Eq. (4) is the ratio of two functions, the rate
Alumina has excellent thermal stability and me-
constants become relative values and can be expressed
chanical strength under processing, transport, and
as ki (T)/k0 (T), where k0 (T) is the rate constant for
catalyst regeneration conditions. However, the most
the forward reaction of paraffin dehydrogenation to
important reason alumina is used as support ma-
mono-olefins.
terial is its superior capability to maintain a high
Eq. (4) can be written in a functional form in F as
degree of platinum dispersion, which is essential
follows:
for achieving high dehydrogenation activity and
 
selectivity. d(sx) ki (T )
=F , Ki (T ), x, P , R (5)
The catalytic reaction rate is limited by the intra- dx k0 (T )
particle mass transfer rate. If the mass transfer rate is
relatively slow, both activity and selectivity are low- Eq. (5) indicates that selectivity is a function of con-
ered. As a result, the support must have a low pore version for the catalyst used (relative rate constants)
diffusional resistance (high effectiveness factor). For and the given reaction conditions (temperature, pres-
a given pore volume, the surface area and the strength sure, feed ratio).
of the support increase as the pore diameter decreases, Selectivity decreases as the conversion increases
and the pore diffusional resistance decreases as the because n-mono-olefins are consecutively converted
pore diameter increases. Thus, an appropriate pore into by-products. Selectivity decreases sharply as
structure must be determined for the support to achieve conversion approaches equilibrium because the main
optimal catalytic performance. dehydrogenation process is limited by equilibrium,
but other reactions continue to occur. Therefore,
if side reactions are controlled, the selectivity is
5. Dehydrogenation catalyst evaluation improved as the equilibrium conversion becomes
higher by increasing the temperature and by decreas-
In paraffin dehydrogenation, the rate of paraffin con- ing the pressure and the feed ratio of hydrogen to
version (x) and mono-olefin production (sx) are given paraffin.
by Eqs. (1) and (2) respectively: The relationship between selectivity and conver-
dx sion can be simulated according to Eq. (5), if rate
= f1 (ki , Ki pj ) (1) functions, relative rate constants, and equilibrium
dt
constants are known. Fig. 8 shows simulated selectiv-
d(sx) ities to n-heptene and n-heptadiene for the dehydro-
= f2 (ki , Ki , pj ) (2) genation of n-heptane. In this simulation, the relative
dt
rate constants used are unity, which represents that
where s is the selectivity to n-mono-olefins, t the con- the catalyst possesses perfect selectivity regarding
tact time, f the rate function, ki the rate constant for consecutive dehydrogenation; the dehydrogenation
reaction step i, Ki the equilibrium constant for reaction rate of paraffin is equal to that of mono-olefin and
step i, and pj is the partial pressure of the j compound. diolefin. Experimental selectivities obtained over a
The following relationship between selectivity and UOP dehydrogenation catalyst show good agreement
conversion can be derived from Eqs. (1) and (2): with the predicted values.
d(sx) f2 (ki , Ki , pj ) The rate of light paraffin conversion (Eq. (1)) over
= (3) a Pt catalyst (Oleflex type process) can be expressed
dx f1 (ki , Ki , pj )
as a modified first order equation according to a
The ki and Ki are a function of temperature and pj Langmuir–Hinshelwood mechanism. The rest of the
is a function of conversion, total pressure (P), and feed equations may be derived accordingly.
404 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

Snamprogetti’s dehydrogenation process consists


of a fluidized bed reactor and regeneration system.
Here too the coke build-up is very low and the
“regeneration” loop is actually a means of supplying
heat to the reactor.

7. Heat of reaction

The heat of reaction for paraffin dehydrogenation


is about 30 kcal/mol (125 kJ/mol). In a cyclic adi-
abatic operation (e.g. Houdry), heat is provided by
reheating the catalyst to a high temperature during
the regeneration step, so that the catalyst cools down
and conversion decreases during the reaction step; be-
cause several reactors are used in parallel, an average
Fig. 8. Simulation of selectivity for dehydrogenation of n-heptane. conversion is obtained. In an isothermal process (e.g.
STAR), the catalyst is loaded inside vertical tubes
inside a furnace and the heat is introduced through
6. Catalyst stability and regeneration
the tube walls. In a fluidized reactor, the temperature
profile can be maintained uniformly in the backmixed
The dehydrogenation of long-chain paraffins is per-
zone of the bed, while heat is provided by intro-
formed under relatively mild temperature conditions
ducing hot regenerated catalyst. In Oleflex adiabatic
of 400–500 ◦ C. Thus, the catalyst can maintain a long
reactors, a significant temperature drop occurs across
life even at high space velocity, and high catalyst pro-
the catalyst bed which lowers the equilibrium conver-
ductivity. Therefore, it is not economical to build fa-
sion level; a multistage reactor system with interstage
cilities for catalyst regeneration.
reheating is used for higher paraffin conversions.
Because of equilibrium limitations, the dehydro-
Fig. 9 illustrates conversion, equilibrium conver-
genation of light paraffins requires significantly higher
sion, and temperature along the catalyst bed in a
temperatures above 600 ◦ C to achieve economically
attractive conversions. The catalyst deactivation is
accelerated under high-temperature conditions, and
frequent catalyst regeneration is necessary for light
paraffin dehydrogenation. For the dehydrogenation of
light paraffins, a number of different types of reactor-
regeneration systems are commercially utilized.
Houdry’s Catofin and similar processes employ a
cyclic sequence of steps—process, purge, air regener-
ation, purge, hydrogen reduction, and back to process.
The Phillips STAR process also regenerates the cat-
alyst on a cyclic basis, but while the Houdry regener-
ation is actually a mechanism to provide the heat for
the reaction even when coke build-up is still very low,
the catalyst in the isothermal STAR process is only
regenerated after coke has accumulated to appreciable
levels that result in low catalyst activity.
UOP’s Oleflex process uses multi-stage adiabatic Fig. 9. Temperature profile and conversions of three-stage isobutane
reactors with CCR. dehydrogenation process.
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 405

(Fig. 10). Thus, thermal cracking and catalyst deacti-


vation, which are accelerated at higher temperatures,
can be controlled to low levels.

8. Process flow and reactor characteristics

8.1. Cyclical processes

As described earlier, the Houdry Catadiene process,


the Houdry Catofin process, and other similar cyclical
processes make use of parallel reactors that contain a
shallow bed of chromia-alumina catalyst. Fig. 11 il-
Fig. 10. Isobutane dehydrogenation. lustrates a schematic of such a process. This technol-
ogy has been used extensively for the production of
butadiene and, in more recent years, for the produc-
three-stage adiabatic reactor system for the dehydro- tion of isobutylene and propylene [16,17]. The feed is
genation of isobutane. For propane dehydrogenation, preheated through a fired heater before being passed
a four-stage reactor system becomes more economi- over the catalyst in the reactors. The hot reactor ef-
cal because higher average temperatures are needed. fluent is cooled, compressed, and sent to the product
A multi-stage reactor system also affords lower inlet fractionation and recovery section. The dehydrogena-
temperatures, relative to a single stage reactor system tion reactors are refractory-lined carbon steel vessels

Fig. 11. Catofin process flow diagram.


406 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

Fig. 12. Typical timing cycle for a five reactor system.

(i.e. cold wall design). In order to accommodate con- range up to 40 in. (1 m) in diameter, are designed
tinuous flow of the main streams (hydrocarbons and for high-temperature service, and are equipped with a
regeneration air), the reactors are operated on a timing pressurized inert seal in the bonnet to prevent leakage
cycle that satisfies the following requirement [1]: of air into the process gas when the valve is closed.
Overall, this mechanical design has proven to be very
on-stream time + regeneration time + purge time reliable over many years of operation.
= total cycle time. The regeneration is done with air that has been pre-
heated through a direct fired burner or, alternatively,
The number of reactors in each cycle is the prorated with the exhaust of a gas turbine. The regeneration
time fraction of the total cycle time. Thus, with five step is intended to preheat the catalyst to the on-stream
reactors, two reactors can be on stream simultaneously, temperature necessary to initiate the next process cycle
two on regeneration, and one on purge, evacuation, and and to remove coke deposits on the catalyst. Flue gas
valve changes. Fig. 12 provides a typical timing cycle sensible heat may be recovered in a waste heat boiler.
for a five-reactor unit [1], but as many as eight reactors The hydrogenation step prepares the catalyst for the
in parallel have been provided in some units. The total dehydrogenation phase and also contributes additional
cycle time is usually in the range of 15–30 min. heat from the reduction of Cr6+ –Cr3+ .
The on-stream period at sub-atmospheric pressure Another cyclical process is the Phillips STAR
is followed by a purge. Next comes regeneration at es- (steam active reforming) process [18]. This process
sentially atmospheric pressure, followed by purge, hy- uses a fixed-bed fired-tube reactor operating at a pos-
drogen reduction, and evacuation to reaction pressure, itive superatmospheric pressure. In many respects, it
after which the reactor is ready for another on-stream is similar in design to a steam reforming furnace with
period. Process streams enter and leave the reactors the heat of reaction provided by firing outside the
through fast-acting gate valves. The gate valves can tubes, thus operating at near isothermal conditions.
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 407
408 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

Steam is used as a diluent to lower the partial pres- an average diameter <100 ␮m and an apparent bulk
sure of the reactants and, thus, to achieve reasonable density <2000 kg/m3 [19]. The heat of reaction is
conversion levels of about 30–40% for propane and provided by circulating hot regenerated catalyst back
45–55% for butanes. Steam also helps slow down the to the reactor. In all concepts, the FBD process is
deposition of carbon (coke) on the catalyst, thereby, very similar to the FCC process units commonly used
extending cycle time from minutes to hours. in petroleum refineries. However, because backmix-
Periodic catalyst regeneration or carbon burnoff is ing has a negative effect on the yields, horizontal
required to maintain the activity of the catalyst. Typ- baffles with suitable openings are inserted within the
ical cycle time is reported to be at least 8 h, with 7 h fluidized bed to limit the back-flow of solids, such
of process time and 1 h of regeneration time. For con- that the fluidized bed is split into a series of stages,
tinuous operation, various furnace modules can be op- each comparable to a CSTR [19]. A typical process
erated such that, for example, seven operate in the scheme is shown in Fig. 14.
process mode while one is in the regeneration mode. Fresh feed is vaporized, mixed with the recycle of
Fig. 13 shows a schematic diagram of a STAR process unconverted paraffins, and fed to the fluidized reactor
unit [18]. through a distributor for optimal even distribution. En-
trained catalyst is removed from the product off-gas by
means of cyclones. Catalyst circulates continuously
9. Continuous processes from the reactor to the regenerator and vice-versa by
means of transfer lines. Coke deposited on the catalyst
Snamprogetti, an Italian company, has commer- is burnt off in the regenerator; however, because the
cialized fluidized bed dehydrogenation (FBD) for the amount of coke is relatively small, additional fuel must
catalytic dehydrogenation of light paraffins using a be burnt in the regenerator in order to satisfy the ther-
chromia-alumina catalyst with an alkaline promoter mal requirements of the endothermic dehydrogenation
[6–8], which is used primarily for the dehydrogena- reaction. However, while this approach is similar to
tion of isobutane to isobutylene during the manufac- that in the Houdry process, FBD does not have a cat-
ture of MTBE. The catalyst is microspheroidal with alyst reduction step with hydrogen before proceeding

Fig. 14. Snamprogetti’s FBD process scheme.


M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 409

Fig. 15. UOP Pacol dehydrogenation process.

to the dehydrogenation cycle; lack of this step is be- performances of the Pacol and Oleflex processes are
lieved to be detrimental to the overall performance of summarized in Table 1.
the process. Use of the Oleflex process for the dehydrogenation
UOP’s catalytic dehydrogenation processes typi- of ethane to ethylene has also been investigated but, to
cally make use of radial flow adiabatic fixed-bed (or date, the economics do not appear to be favorable be-
slowly moving bed) reactors with modified Pt-alumina cause of the low equilibrium conversion and the need
catalysts. to operate at a pressure lower than atmospheric if a rea-
The UOP Pacol process for selective long-chain sonable ethane conversion is to be expected. The cost
paraffin dehydrogenation to produce linear mono- of fractionating ethylene in an ethane–ethylene split-
olefins is shown in Fig. 15 in combination with the ter is otherwise too high. Dow Chemical has recently
UOP detergent alkylation process. The Pacol process been awarded a patent [21] for the dehydrogenation
consists of a radial-flow reactor and a product recov- of ethane over a metal-mordenite catalyst complex
ery section. Worldwide, more than 2 million MTA of at relatively low conversions in which the product
linear alkyl benzene (LAB) is produced employing
this process [20].
Table 1
The flow diagram of the UOP Oleflex process Performance of Pacol and Oleflex processes
is shown in Fig. 16. The process consists of a re-
Process Feed Conversion (%) Selectivity (%)
actor section and a product recovery section. The
reactor section consists of three or four stages of Oleflex Propane 40 90
radial-flow reactors, charge and interstage heaters, n-Butane 50 85
Isobutane 50 92
reactor feed-effluent exchangers, and the CCR sec-
tion (Fig. 17). As noted earlier, today more than 1 Pacol n-Heptane 20 90
million metric tons propylene and 2 million metric n-C10 –C13 13 90
n-C11 –C14 13 90
tons isobutylene are produced via this route [13]. The
410 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

Fig. 16. UOP Oleflex process.

Fig. 17. Oleflex regeneration section.


M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 411

Table 2
Characteristics of five reactor systemsa
Downflow Radial flow Tubular Fluidized bed

Low pressure drop – –


Plug flow – – –
Catalyst addition or removal – –
High heat transfer, near isothermal – –
Variable heat transfer coefficient
a The choice of the right reactor depends on the catalyst and the selection of operating conditions. The ‘dash’ represents beneficial

characteristics of each reactor type.

ethylene is selectively recovered from the dilute proposed by Benson [22] for the production of
ethylene–ethane stream by alkylating it with benzene. acetylene and ethylene. Other chlorination/dehydro-
chlorination cycles have been proposed for the
production of ethylene from ethane. Propane dehydro-
10. Reactor design options genation in the presence of iodine via a propyl iodide
intermediate has also been proposed [4,23]. Apart
The choice of reactor design plays a very important from the apparent corrosion problems associated with
role in the success of catalytic processes. The follow- the use of halogens, other difficulties readily come
ing types of reactor design are commercial today for to mind owing to the relatively high cost of chlorine,
endothermic catalytic dehydrogenation processes: and even more so of iodine, and the need to either
dispose of or recycle vast quantities of halogens.
• downflow adiabatic fixed-bed;
Oxydehydrogenation or oxidative dehydrogenation
• radial flow fixed-bed or moving bed adiabatic;
can be considered in at least two different ways.
• tubular isothermal and
Use of oxygen to oxidize the hydrogen coproduct
• fluidized bed
from dehydrogenation, and thus to displace the dehy-
Table 2 summarizes the main characteristics of the drogenation equilibrium to higher conversions. This
four reactor systems. approach has been used commercially in the catalytic
dehydrogenation of ethylbenzene to styrene as in
the UOP Styro-PlusTM process or in the ABB Lum-
11. Other dehydrogenation technologies mus/UOP SMARTTM process, but to date has not
succeeded in the dehydrogenation of light or heavy
The processes discussed above are for the direct cat- paraffins. This technology has been used in a styrene
alytic dehydrogenation of paraffins to the correspond- unit at Mitsubishi Chemicals, Kashima, Japan. Al-
ing olefins or of olefins to diolefins. Other approaches though a similar approach has been proposed for the
have also been considered, although none has reached dehydrogenation of paraffins [24–27], it has not been
the level of commercialization. Some of the most no- commercialised.
table are Direct use of oxygen as a means of dehydrogenat-
ing, for example, ethane to ethylene. Oxydehydro-
• halogen-assisted dehydrogenation and
genation has successful commercial applications in the
• oxydehydrogenation.
conversion of n-butenes to butadiene (e.g. as in the
Use of halogens for the dehydrogenation of paraf- Oxo-D process referred to earlier), but not yet for the
fins has been proposed in different ways. For example, production of ethylene or propylene. This subject is
heavy paraffins were first chlorinated and then dehy- analyzed in more detail in the following section.
drochlorinated to heavy olefins commercially in the Use of oxydehydrogenation relative to straight cat-
past both by Shell and by Hüls, among others. Pyrol- alytic dehydrogenation must be viewed both in terms
ysis of methane in the presence of chlorine has been of safety issues and in an economic context. On the
412 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

latter, even though oxydehydrogenation offers advan- than the olefins, and so catalysts must be found that
tages as a means of overcoming thermodynamic equi- can stop the reaction at the olefin rather than allowing
librium limitations, it also leads to the total or partial it to proceed on to the oxides. Also, to compete with
loss of byproduct hydrogen, which in some instances steam cracking, the selectivity to olefins must be quite
can have a very significant economic impact. high. Selectivity to ethylene from ethane in steam
Although less apparent, oxydehydrogenation also cracking is reported to be about 84% at 54% ethane
plays a role in the work done by BP Amoco, Asahi, conversion (800 ◦ C, 0.3 kg steam per kg feed, 0.79 s
and others to extend ammoxidation to the direct con- residence time, and 154 kPa hydrocarbon partial pres-
version of propane to acrylonitrile. It is believed that sure), and 78% at 69% ethane conversion (833 ◦ C,
the ammoxidation of propane proceeds through a tran- 0.3 kg steam per kg feed, 0.75 s residence time, and
sient propylene intermediate from which acrylonitrile 154 kPa hydrocarbon partial pressure) [29].
is derived through a conventional ammoxidation path-
way [28]. 11.1.1. Ethane oxydehydrogenation
Ethane oxydehydrogenation as a route to ethylene
11.1. Oxydehydrogenation of ethane and propane has been examined at temperatures in the range of
300–500 ◦ C with reducible metal oxide catalysts, and
Oxydehydrogenation of ethane and propane as a at higher temperatures, above about 600 ◦ C, with
route to ethylene and propylene, respectively, has largely non-reducible and reducible metal oxide cata-
the attractive feature of removing the equilibrium lysts. The latter have evolved primarily out of investi-
conversion restriction of dehydrogenation. Table 3 gations of methane oxidative coupling [80–82]. In the
contrasts the calculated equilibrium conversions of oxidative coupling of methane, ethane and ethylene
oxydehydrogenation of ethane and propane with the are major products, and ethylene has been shown to
calculated equilibrium conversions for dehydrogena- be derived from ethane [80–82]. The dehydrogena-
tion. The formation of water rather than hydrogen tion or oxydehydrogenation of ethane to ethylene,
in oxydehydrogenation effectively removes the equi- must occur over these catalysts, and since the overall
librium constraint on conversion at all temperatures selectivity is high, the selectivity to ethylene must
of interest. At the same time, oxydehydrogenation be good. In addition to these two lines of research,
is not without its own set of challenges that, in the recent work at high temperatures using Pt and Pt-Sn
case of ethane and propane, have kept oxydehydro- on a monolith support has been reported.
genation from being practiced on a commercial scale.
Challenges in oxydehydrogenation include handling
mixtures of paraffins and oxygen, which can be ex- 12. Lower temperature ethane
plosive at certain compositions, suitable conversion of oxydehydrogenation
paraffins, which often is limited by maintaining a safe
paraffin–oxygen composition, and suitable selectivity Ethane oxydehydrogenation at temperatures in the
to olefins. The carbon oxides—carbon monoxide and range of 300–400 ◦ C is conducted with reducible
carbon dioxide—are thermodynamically more stable metal oxide catalysts usually containing vanadium.

Table 3
Percentage conversions at equilibrium for dehydrogenation and oxydehydrogenation of ethane and propane at atmospheric pressurea
Temperature (K) Ethane Propane

Dehydrogenation Oxydehadrogenation Dehydrogenation Oxydehydrogenation

400 <1 100 <1 100


600 <1 100 1 100
800 7 100 25 100
1000 51 100 87 100
a Stoichiometric mixture of alkane and oxygen for oxydehydrogenation.
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 413

Early work in this area was conducted by Thorstein- of materials, most not containing the elements
son et al. using a vanadium molybdenum niobium molybdenum and vanadium, which are common
oxide catalyst [30,31], and a number of papers and in lower-temperature catalysts. As opposed to the
patents based on that work have been issued by Union lower-temperature catalysts which tend to produce
Carbide Corp. and other laboratories [32–39]. Ethane coproducts, the higher temperature processes tend to
is thought to react with molybdenum or vanadium produce only ethylene as C2 product, though some do
in the catalyst to form surface ethoxide, which can produce methane from cracking reactions. Oxygenated
then undergo a beta-elimination process to form ethy- products other than carbon oxides largely are absent,
lene. The surface ethoxide can be oxidized further likely because of their instability at high temperature,
to make surface acetate, which leads to acetic acid especially in contact with the catalyst and oxygen.
on hydrolysis. Acetic acid in varying amounts is co- Li and coworkers at the Lanzhou Institute
produced with ethylene at pressures greater than one of Chemical Physics, China, have shown that
atmosphere. Later work has shown that catalysts can a Na2 WO4 -Mn/SiO2 catalyst (a high-selectivity
be made selective for either ethylene or acetic acid methane coupling catalyst) at 700 ◦ C is capable of
by modifications in the elemental makeup of the cat- giving greater than 70% selectivity to ethylene (and
alyst and suitable adjustments of reaction conditions 10% selectivity to methane) at >70% conversion of
[38,39]. Selectivity to ethylene with these catalysts is ethane [42]. Wang and coworkers at the National In-
in the range of approximately 70% at approximately stitute of Materials and Chemical Research, Japan,
70% conversion of ethane. found that lithium chloride on sulfated zirconia at
Ethane oxydehydrogenation at temperatures in 650 ◦ C yielded 70% selectivity to ethylene (and 2%
the range of ≥500 ◦ C may be conducted with phos- to methane) at 98% conversion of ethane [43,44]. Al-
phorous/molybdenum/antimony oxide catalysts [40] though the catalyst did show some deactivation with
or iron-containing solid solution catalysts stabilized time, it can still maintain an ethylene yield as high
with one or more metal oxides [41]. Phosphorous/ as 50% after 24 h. Lin et al. at the Tokyo Institute
molybdenum/antimony oxide catalysts, suitably modi- of Technology, Japan, and Chonnam National Uni-
fied with other elements give a selectivity to ethylene versity, Korea, found that a SrBi3 O4 Cl3 catalyst at
in the range of 78% at 20% conversion of ethane. Iron- 640 ◦ C is capable of giving 90% selectivity to ethy-
containing solid solution catalysts stabilized with lene at a 25% conversion of ethane [45]. Co-feeding
metal oxide are optimally used with a flow of hydro- HCl did not change the conversion or the selectivity,
gen chloride and water in addition to ethane and wa- but did slow the activity decrease along with the loss
ter. At a contact time of 12 s at 550 ◦ C over an iron in of chlorine observed in its absence.
␣-alumina solid solution catalyst stabilized with lant- Dang et al. at the Lanzhou Institute of Chemical
hanum oxide, ethane conversion is in the range of 90%. Physics, China, have examined a vanadium oxide
Products are ethylene (80–93% selectivity) and vinyl on barium carbonate catalyst at 650 ◦ C that gave
chloride (1–4% selectivity). As a route to ethylene- good activity [46]. At 34% conversion of ethane,
only, this route seems unattractive because of the pro- they obtained 76% selectivity for ethylene. Au and
duction of vinyl chloride, which must be separated coworkers at Hong Kong Baptist University examined
from ethylene. For vinyl chloride producers, however, SrFeO3−δ Clσ and Ho2 O3 with BaCl2 [47,48]. Of the
where the stream may be able to be used without two, SrFeO3−δ Clσ gave the higher yield of ethylene.
separation, such a high selectivity/conversion process At 680 ◦ C, SrFeO3−δ Clσ yielded 90% ethane con-
to ethylene may possibly be economically attractive. version and 70% selectivity. Over 40 h, there was no
drop-off in yield or in chlorine content of the cata-
lyst. Longer-duration tests were not conducted. The
13. Higher temperature ethane Ho2 O3 with BaCl2 catalyst at 640 ◦ C yielded 57%
oxydehydrogenation conversion and 68% selectivity to ethylene.
Choudhary et al. at the National Chemical Labo-
Ethane oxydehydrogenation at temperatures in the ratory, India, have examined strontium and other rare
range of 650–800 ◦ C is conducted with a variety earth oxides deposited on sintered low surface area
414 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

supports precoated with lanthanum and other rare earth ethane of 67% [56,57]. The authors’ consideration
oxides [49,50]. A Sr-Nd2 O3 catalyst showed the high- on mechanism follows: “Qualitatively, this process
est activity and selectivity for ethane. At 800 ◦ C, the must involve first the oxidation of H2 to H2 O, which
Sr-Nd2 O3 catalyst was capable of giving 60% con- generates heat and removes O2 , followed by dehy-
version of ethane and greater than 80% selectivity to drogenation of C2 H6 to produce C2 H4 and H2 , and
ethylene. all of these reactions occur within ∼10−3 s. Possi-
Workers at Phillips Petroleum Company have ex- ble mechanisms to explain the results are (i) purely
amined catalysts comprising lithium, titanium, and catalytic reactions on the Pt–Sn surface, (ii) purely
manganese, and catalysts comprising cobalt, phos- homogeneous reaction and (iii) catalytic H2 oxidation
phorous, and at least one promoter selected from a list followed by homogeneous ethane decomposition. We
of elements [51,52]. At 650 ◦ C, the lithium/titanium will consider each of these mechanisms and show
catalyst with manganese gave an ethane conversion of that, while each of these gives partial interpretation of
47% and a selectivity to ethylene of 75%. At 670 ◦ C, results, none appears to be totally satisfactory” [56].
a catalyst containing cobalt, phosphorous potassium, Hydrogen is produced in the process in greater
and zirconium where the cobalt was introduced as amount than fed, so that hydrogen required for the
cobalt sulfide, and calcining of the catalyst was ac- process can be recycled from downstream equipment.
complished in the absence of oxygen at 670 ◦ C, gave Such high selectivity to ethylene coupled with low
an 85% conversion of ethane and a 86% selectivity to contact time and high conversion of ethane makes this
ethylene (25 vol.% ethane/75 vol.% air at 3 psig and process an attractive possibility as an alternative to
2400 GHSV). Addition of a halogen-containing com- steam cracking of hydrocarbons as a commercial route
pound, such as methyl chloride, increases ethylene to ethylene. While the possibility is bright, a number
yield. There is an indication that the catalyst loses of questions need to be resolved. Among these are the
activity with use over time for oxydehydrogenation of questions of long-term catalyst activity, long-term se-
hydrocarbons. If catalyst activity and selectivity can lectivity, and whether such mixtures of ethane, hydro-
be maintained over long periods of time, this catalyst gen, and oxygen in the ratios needed, reported to be
seems like a good potential candidate for use in an 2:2:1 at 950 ◦ C, can be handled safely on a commercial
economical process for ethane oxydehydrogenation scale. In addition, issues related to process start-ups,
to ethylene. shutdowns, and re-starts have to be addressed.

14. Ethane oxydehydrogenation over 15. Ethane oxydehydrogenation with


monolith catalysts carbon dioxide

Ethane oxydehydrogenation over Pt and Pt-Sn Another approach that is being pursued in sev-
coated monolith catalysts at high temperatures and eral laboratories is to use CO2 as a mild oxidant for
extremely short contact times has been the subject of oxydehydrogenation of hydrocarbons. [42,62–68]. In
work by Schmidt and coworkers at the University of addition to ethane, oxidative dehydrogenation of ethyl-
Minnesota for the past several years [53–58]. Work benzene [67] and of propane has also been demon-
along similar lines has also been reported to be un- strated [65,66,68]. Extensive work has been done on
der way in Russia. Similar results are reported over oxidative dehydrogenation of ethane with CO2 by
platinum gauze and platinum coated pellets [59–61]. Xu et al. [62] of the Dalian Institute of Chemical
Conversions of ethane on the order of 70% and se- Physics. Thus, reaction temperature was significantly
lectivities on the order of 65% are obtained over a lowered from that used in steam cracking, though it
Pt-coated ceramic foam monolith at approximately is still quite high. High selectivity was obtained at
1000 ◦ C at contact times on the order of 1 ms. With ∼800 ◦ C, 1000 h−1 GHSV and 0.1 MPa (∼1 atm) us-
the addition of tin to the catalyst and with the addi- ing several catalysts containing oxides of Cr, Cr–Mn,
tion of hydrogen to the feed, selectivity to ethylene Cr–Mn–Ni, Cr–Mn–La onto a Silicalite-2 (Si-2) ze-
could be increased to above 85% at a conversion of olite. Highest conversion/selectivity observed were
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 415

69.8%/90.8% at 800 ◦ C and 800 h−1 and 72.2%/90.4% Thus, using a tail-gas containing 18.8% C2 H6 and
at 820 ◦ C/1000 h−1 over a Cr/Si-2 catalyst. These con- 19.2% C2 H4 , the ethylene content was increased to
versions approached equilibrium for these two condi- 25.4–27.2% depending on the composition of the
tions; which were calculated to be 72.8 and 77.6%. catalyst. This reaction was carried out at 1073 K,
Oxydehydrogenation using CO2 according to the 0.1 MPa, 1000 h−1 GHSV and CO2 :C2 H6 mole ratio
authors [62] has the following special features. of 1:1. The Cr/Si-2 catalyst gave the lowest ethylene
enhancement (25.4%) while Cr-Mn-La/Si-2 catalyst
• CO2 acts as a mild oxidant for the oxidative dehy-
gave the highest ethylene enhancement to 27.2%.
drogenation of C2 H6 to yield C2 H4 .
The corresponding ethane conversions/selectivity are
• A significant lowering of the reaction temperature
60.6%/79.6% and 63.6%/85.8%, respectively.
as compared to the steam cracking process, resulting
Some of the other early work in this area was
in lower coke formation on the catalyst.
done using highly selective catalysts for the oxidative
• C2 H6 conversion and C2 H4 selectivity are higher
coupling of methane by Liu et al. of the Lanzhou
in this process than in all other processes of C2 H4
Institute of Chemical Physics [42]. Their work,
production from C2 H6 . Furthermore, no C3 + and
employing a selective methane coupling catalyst
C2 H2 products are formed.
(Na2 WO4 -Mn/SiO2 ), showed that >70% conversion
The dehydrogenation of ethane is facilitated by the and selectivity could be achieved at 700–750 ◦ C and
reaction of CO2 with H2 (reverse water-gas shift) to space velocities of >30,000 h−1 , employing O2 as the
make CO + H2 O and also with ethane and methane as oxidant. This is in contrast with 53% conversion/97%
shown as follows: selectivity at 800 ◦ C and 69.5% conversion/90.5%
selectivity at 850 ◦ C employing C2 H6 :CO2 = 1:1 and
C2 H6 = C2 H4 + H2 3600 h−1 space velocity. These catalysts were sta-
CO2 + H2 = CO + H2 O ble for 100 h of operation. The authors propose that
surface lattice oxygen is responsible for selective oxy-
C2 H6 + 2CO2 = 4CO + 3H2 dehydrogenation while the bulk lattice oxygen is re-
sponsible for deep (non-selective) oxidation of ethane.
CH4 + CO2 = 2CO + 2H2
Nakagawa et al. [63] studied a series of metal ox-
In addition, ethane hydrogenolysis to methane also ides for ethane dehydrogenation in the presence of
provides favorable free energy of reaction. CO2 at 650 ◦ C, C2 H6 :CO2 = 5:25 ml/min and SV =
900 h−1 ml (g-cat)−1 . The order of activity was as
C2 H6 + H2 = 2CH4 follows:
The authors suggest that ethane oxydehydrogena-
tion with carbon dioxide takes place according to the Ga2 O3 > Cr 2 O3 > V2 O5 > TiO2
following overall stoichiometry: > Mn3 O4 > In2 O3 > ZnO > La2 O3 .

16C2 H6 + 9CO2 The ethylene selectivity was generally >85% in the


= 14C2 H4 + 12CO + 6H2 O + 12H2 + CH4 presence of CO2 . Ethylene yields for Ga2 O3 in the
presence of CO2 was approximately twice than in its
It can be seen from this reaction scheme that in or- absence. The ethylene yield enhancement with Cr2 O3
der to increase the selectivity to ethylene, the forma- was only a modest 1.2%, while in the case of V2 O5 ,
tion of methane must be suppressed and that the key there was a 2.7% detrimental effect. The authors pro-
problem is to develop a catalyst that can suppress the pose involvement of acid sites in the dehydrogenation
thermodynamically favorable side reactions. reaction.
The work of Xu et al. [62] also demonstrated Wang et al. [64] of Japan studied the effect of sul-
that the inclusion of steam in the feed leads to a fate and Na-treatment silica on the dehydrogenation
three-fold reduction in coke formation at 800 ◦ C. of ethane with CO2 at 650 ◦ C and 1 atm. Sulfating
These authors also studied the C2 H6 + CO2 reac- silica (348 m2 /g) with (NH4 )2 SO4 (and calcining
tion in FCC tail-gas to increase the ethylene content. at 700 ◦ C/3 h to give ∼2 wt.% SO4 2− ), and adding
416 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

5 wt.% Cr gave the best catalytic performance. • at higher temperatures (>950 K), alkanes may react
A conversion/selectivity/yield of 53.9/87.6/55.2% directly with oxygen left from CO2 .
was observed at 650 ◦ C/1 atm. With a 10:50:40 =
The authors also explain the reactivity of catalysts
ethane:CO2 :N2 mixture. The corresponding 5% Cr
in terms of acidic and basic nature of supports and
catalyst (3.34 m2 /g on silica gave 56.1/92.9/52.1%
the reactivity of CO2 with such supports. Typically,
conversion/selectivity/yield. However, Na-Cr/SiO2
good active catalysts appear to approach thermody-
gave very poor performance of 4.9/98.2/4.8%. Further-
namic equilibrium. However, temperatures required
more, the surface area dropped to 3.1 m2 /g. In com-
to achieve higher conversions and yields also lead to
parison, silica alone of 348 m2 /g gave 2.2/97.2/0.2%
higher levels of coking and by-product formation. No
conversion/selectivity/yield. Dehydrogenation is pro-
economic assessment of the CO2 -based dehydrogena-
posed to occur by abstraction of H by oxygen species
tion processes is available. The authors are not aware
which in turn are formed from surface carbonates
of any commercialization efforts using this reaction.
decomposition. At low temperatures, decomposition
of carbonates absorbed on strong basic sites restricts
formation and mobility of oxygen resulting in lower
activity of Cr/Na-SiO2 catalyst. However, sulfation 16. Propane oxydehydrogenation
of silica favors reaction with hydrocarbon and hence,
higher activity (along with formation of CO + H2 ). Propane oxydehydrogenation combining high
The Cr/SiO2 and Cr/SO4 2− -SiO2 were shown to be propane conversion and high propylene selectivity has
stable for 5–6 h of operation. proved an elusive goal, and propane oxydehydrogena-
Macho and coworkers [65,66] studied oxy- tion as a route to propylene appears to be far from
dehydrogenation of propane and butane over realizing its commercialization potential. One diffi-
Mn-Cr-K/Y-Al2 O3 to give propene and butenes, re- culty is that propylene is more easily oxidized than
spectively, according to the following reaction: is propane, so that selectivity tends to decline rapidly
with conversion. Another is that at temperatures above
C3 H8 + CO2 ⇔ C3 H6 + CO + H2 O
about 700 ◦ C, propane cracking becomes significant,
C4 H10 + CO2 ⇔ C4 H8 + CO + H2 O and a variety of products other than propylene are
produced. Efforts to increase propylene selectivity
Splitting of the C–C bond proceeds simultaneously
have centered in the areas of more selective catalysts
leading to the formation of both higher and lower hy-
at temperatures below 700 ◦ C, more selective cata-
drocarbons as shown in the following equation. These
lysts and conditions at high temperatures, membrane
reactions are preferred at relatively higher tempera-
reactors, and cyclic-operation reactors.
tures while the relatively lower temperatures prefer
A variety of catalysts for propane oxydehydrogena-
straight dehydrogenation with CO2 . In addition, coke
tion have been examined recently at temperatures be-
formation is more prevalent at relatively higher tem-
low 700 ◦ C [69–75]. Particularly, notable for its high
peratures.
selectivity to propylene is a vanadia-silica-zirconia
2C3 H6 ⇔ C4 H8 + C2 H4 catalyst of Rulkens and Tilley at the University of
California, Berkley, CA [75]. These workers using
4C3 H8 + 4CO2 ⇔ 3C4 H8 + 4CO + 4H2 O
a molecular precursor route, produced an 18/36/46
C3 H8 + 3CO2 ⇔ C2 H4 + 4CO + 2H2 O V2 O5 -SiO2 -ZrO2 catalyst which gave 81.5% selec-
tivity to propylene at 8% conversion of propane at
C3 H8 ⇔ C3 H6 + H2
550 ◦ C. The presence of zirconium is important in re-
C2 H4 + CO2 ⇔ CH4 + 2CO taining high selectivity, likely by stabilizing the dis-
persion of vanadia. These catalysts also are among the
The authors propose that:
most active for propane oxydehydrogenation.
• at middle temperatures (850–1000 K) dehydrogena- Ranzi and coworkers at the Polytechnic University
tion of alkanes proceeds first and hydrogen gener- of Milan [76] and Choudhary et al. at the National
ated reacts with CO2 giving CO + H2 O; Chemical Laboratory of India [77] have examined
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 417

oxydehydrogenation of propane at high temperatures intermediate. This certainly appears to be the case with
with and without catalysts. Selectivities to olefins, the majority, if not all, of the catalysts that are active
comprising primarily ethylene and propylene, can be for oxydehydrogenation that operate at 500–800 ◦ C.
obtained in selectivities of 50–80%. Conversion of However, under these conditions, it is very difficult
propane is on the order of 90% under these conditions. to avoid undesirable partial/total oxidation to CO and
Alfonso, Julbe, Farrusseng, Menéndez, and Santa- CO2 via gas phase and surface reactions. Therefore,
marı́a at the University of Zaragoza, Spain, and the the challenge for oxydehydrogenation catalysis is
Laboratoire des Matériaux et Procédés Membranaires, to develop highly active and selective catalysts for
France, examined the oxidative dehydrogenation of totally selective oxidation of alkanes to alkenes.
propane over V-Al2 O3 catalytic membranes which al-
low separation of propane and oxygen feeds [78]. Se-
References
lectivities of 51% to propylene were obtained at 8%
conversion of propane at 550 ◦ C, which is higher than
[1] G.F. Hornaday, F.M. Ferrell, G.A. Mills, Manufacture
the 44% selectivity obtained at the same temperature of mono- and diolefins from paraffins by catalytic
and conversion with premixed feeds. dehydrogenation, in: Advances in Petroleum Chemistry and
Creaser et al. at Lulea University of Technol- Refining, Vol. 4, Interscience, Pans, 1961.
ogy, Sweden, University of Waterloo, Canada, and [2] V.N. Ipatieff, R.E. Schaad, Mixed polymerization of butenes
by phosphoric acid catalyst, Ind. Eng. Chem. 30 (1938) 596–
Chalmers University of Technology, Sweden, exam- 599.
ined cyclic operation of the oxidative dehydrogena- [3] A.L. Waddams, Chemicals from Petroleum, 4th Edition, Gulf
tion of propane over a V-Mg-O catalyst [79]. In a Publishing Company, Houston, 1978, 1980.
cyclic reactor where oxygen and propane were alter- [4] A.H. Weiss, The manufacture of propylene (158th Meeting of
the American Chemical Society, 10–12 September 1969), in:
nately passed over the catalyst, propylene selectivity
Refining Petroleum for Chemicals, Advances in Chemistry
was considerably increased at 510 ◦ C compared to Series, Vol. 97, American Chemical Society, Washington, DC,
steady state operation with mixed feeds. At a 1:1 1970.
propane:oxygen ratio, selectivity to propylene was [5] R.G. Craig, D.C. Spence, Catalytic dehydrogenation of
78% at 4.3% conversion of propane with cyclic op- liquefied petroleum gas by the Houdry Catofin and Catadiene
processes, in: R.A. Meyers (Ed.), Handbook of Petroleum
eration compared to 55% selectivity to propylene at
Refining Processes, McGraw-Hill, New York, 1986 (Chapter
5.5% conversion of propane with mixed feeds. These 4.1).
conversions/selectivities are still far from being com- [6] D. Sanfilippo, F. Buonomo, G. Fusco, I. Miracca, Paraffins
mercially attractive. Activation through Fluidized Bed Dehydrogenation: the
Answer to Light Olefins Demand Increase, Elsevier,
Amsterdam, Stud. Surf. Sci. Catal. 119 (1998) 919–924.
[7] R. Iezzi, A. Bartolini, U.S. Patent 5,633,421 (27 May 1997),
17. Summary assigned to Snamprogetti.
[8] E.C. Luckenbach, F.A. Zenz, Giovanni Papa, A. Bertolini,
It is interesting to note that though oxydehydro- U.S. Patent 5,656,243 (12 August 1997), assigned to
Snamprogetti.
genation is thermodynamically favored at even am-
[9] V. Haensel, U.S. Patent 2,602,772 (8 July 1952), and other
bient conditions, no catalyst has been discovered that patents on catalytic reforming with Pt catalysts assigned to
can provide such a true oxidative conversion at tem- UOP.
peratures below ∼300 ◦ C. Catalysts discovered by [10] H.S. Bloch, UOP discloses new way to make linear
Union Carbide that operate at 300–400 ◦ C [80–82], alkylbenzene, Oil Gas J. 79–81 (1967); U.S. Patent
3,448,165 (3 June 1969), and other patents on the catalytic
are the best known catalysts for producing ethylene dehydrogenation of paraffinic hydrocarbons assigned to
(with co-production of some acetic acid in varying Herman Bloch and UOP.
amounts). Thus, a true oxydehydrogenation catalyst [11] R.C. Berg, B.V. Vora, Detergent alkylate, in: Encyclopedia
is yet to be discovered. Another explanation may of Chemical Processing and Design, Vol. 15, Marcel Dekker,
very well be that all of the currently known cata- New York, 1982, pp. 266–284.
[12] T. Imai, B.V. Vora, J.C. Bricker, Development of
lysts proceed through a dehydrogenation step (C–H dehydrogenation catalysts and processes, Petroleum and
activation and bond breakage) followed by oxida- Petrochemical College, Chulalongkorn University, Bangkok,
tion of the evolved H2 or a surface H-containing December 1989.
418 M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419

[13] J.H. Gregor, R. Antonelli, T.D. Foley, E.C. Arnold, Increased [29] Ullmann’s Encyclopedia of Industrial Chemistry, 5th Edition,
opportunities for propane dehydrogenation, DeWitt World Vol. A10, VCH Verlagsgesellschaft, GmbH, Germany, 1987,
Petrochemical Review, Houston, TX, 23–25 March 1999. p. 54.
[14] C.P. Poole, D.S. MacIver, The physical–chemical properties [30] E.M. Thorsteinson, T.P. Wilson, F.G. Young, P.H. Kasai, J.
of chromia-alumina catalysts, in: Advances in Catalysis, Vol. Catal. 52 (1978) 116.
17, Academic Press, New York, 1967, pp. 223–314. [31] E.M. Thorsteinson, T.P. Wilson, F.G. Young, P.H.U.S. Kasai,
[15] C.P. Poole, W.L. Kehl, D.S. MacIver, Physical properties of U.S. Patent 4,250,346 (1981), assigned to Union Carbide.
coprecipitated chromia-alumina catalysts, J. Catal. 1 (1962) [32] J.H. McCain, U.S. Patent 4,524,236 (18 June 1985), assigned
407–415. to Union Carbide.
[16] E.L. Tucci, J.M. Dufallo, R.J. Feldman, Commercial [33] R.M. Manyik, J.L. Brockwell, J.E. Kendall, U.S. Patent
performance of the Houdry Catofin process for isobutylene 4,899,003 (6 February 1990), assigned to Union Carbide.
production for MTBE, in: Proceedings of the Workshop on [34] K. Ruth, R. Kieffer, R. Burch, J. Catal. 175 (1998) 16.
Catalysts and Catalytic Processes, Research Institute, King [35] K. Ruth, R. Burch, R. Kieffer, J. Catal. 175 (1998) 27.
Fahd University of Petroleum and Minerals, Dhaharan, Saudi [36] M. Kitson, U.S. Patent 5,260,250 (9 November 1993),
Arabia, 6 November 1991. assigned to BP Chemicals.
[17] S. Gussow, R. Whitehead, Isobutane dehydrogenation by [37] W. Ueda, N.F. Chen, K. Oshihara, Kinet. Catal. 40 (1999)
Catofin as feed for motor fuel ether, in: Proceedings of the 401 (English).
NPRA Annual Meeting, San Antonio, TX, 17–19 March [38] W. Ueda, N.F. Chen, K. Oshihara, Chem. Commun. 517
1991. (1999).
[18] R.O. Dunn, F.M. Brinkmeyer, G.F. Schuette, The Phillips [39] C. Hailett, European Patent Application, Publication
STAR process for the dehydrogenation of C3 , C4 , and C5 0,480,594,A2 (15 April 1992), assigned to B.P. Chemicals.
paraffins, in: Proceedings of the NPRA Annual Meeting, New [40] M. Kutyrev, F. Cavani, F. Trifiro, European Patent
Orleans, LA, 22–24 March 1992. Application, Publication 0,544,372,A1 (2 June 1993),
[19] F. Buonomo, G. Fusco, I. Miracca, G. Papa, D. Sanfilippo, assigned to Eniricherche S.p.A.
Fluid bed dehydrogenation of light paraffins: the FBD [41] J. Magistro, U.S. Patent 5,087,791 (11 February 1992),
technology, in: Proceedings of Petrotech ’96, Bahrain, 10–12 assigned to B.F. Goodrich.
June 1996. [42] Y. Liu, J. Xue, X. Liu, R. Hou, S. Li, in: A. Parmaliana, et
[20] R.J. Lawson, T.R. Fritsch, Developing a clean, efficient al. (Eds.), Natural Gas Conversion V, Elsevier, Amsterdam,
process, Asia-Pacific Chem. 99 (1999) 8–9. Stud. Surf. Sci. Catal. 119 (1998) 593,
[21] A.Q. Campbell, J.M. Garcés, T.M. May, R.F. Pogue, [43] S. Wang, K. Murata, T. Hayakawa, S. Hamakawa, K. Suzuki,
Preparation of ethylbenzene and substituted derivatives by Catal. Lett. 59 (1999) 173.
alkylation using unpurified reaction products of ethylene [44] S. Wang, K. Murata, T. Hayakawa, S. Hamakawa, K. Suzuki,
prepared by dehydrogenation of ethane, World Patent WO Chem. Commun. 103 (1999).
96/34843 (7 November 1996), assigned to Dow Chemical Co. [45] S.W. Lin, Y.C. Kim, W. Ueda, Bull. Chem. Soc. Jpn. 71
[22] S. Benson, Conversion of methane using chlorine, U.S. Patent (1998) 1089.
4,199,533 (22 April 1996), assigned to the University of [46] Z. Dang, J. Gu, J. Lin, D. Yang, Catal. Lett. 54 (1998) 129.
Southern California. [47] H.X. Dai, C.F. Ng, C.T. Au, Catal. Lett. 57 (1999) 115.
[23] A.H. Weiss, Which propylene process is best? Hydrocarbon [48] C.T. Au, K.D. Chen, H.X. Dai, Y.W. Liu, C.F. Ng, Appl.
Process. 47 (1968) 123–127. Catal. A 177 (1999) 185.
[24] T. Laegreid, M. Rönnekleiv, A. Sobbaken, in: Proceedings of [49] V.R. Choudhary, S.A.R. Mulla, V.H. Rane, J. Chem. Technol.
the DGMK Conference, 11–12 November 1993, p. 147. Biotechnol. 71 (1998) 167.
[25] R. Lodeng, P. Soaker, Reactor for catalytic dehydrogenation [50] V.R. Choudhary, B.S. Uphade, S.A.R. Mulla, U.S. Patent
of hydrocarbons with selective oxidation of hydrogen, U.S. 5,712,217 (27 January 1998), assigned to Council of Scientific
Patent 5,997,826 (7 December 1999), assigned to Den Norske and Industrial Research, India.
Stats Oljeselskap A.S. (Statoil). [51] A.D. Eastman, J.B. Kimble, U.S. Patent 4,450,313 (22 May
[26] J.G. Tsikoyiannis, D.L. Stern, R.K. Grasselli, J. Catal. 184 1984), assigned to Phillips Petroleum.
(1999) 77. [52] A.D. Eastman, J.P. Guillory, C.F. Cook, J.B. Kimble, U.S.
[27] P.A. Agaskar, R.K. Grasselli, J.N. Michaels, P.T. Reischman, Patent 4,497,971 (5 February 1985), assigned to Phillips
D.L. Stern, J.G. Tsikoyiannis, Process for the catalytic Petroleum.
dehydrogenation of alkanes to alkenes with simultaneous [53] L.D. Schmidt, M. Huff, World Patent 96/13475 (9 May 1996),
combustion of hydrogen, U.S. Patent 5,430,209 (4 July 1995), assigned to Regents of the University of Minnesota.
assigned to Mobil Oil Corp. [54] S.S. Bharadwaj, L.D. Schmidt, World Patent 96/33149 (24
[28] J.F. Brazdil, A.P. Cavalcanti, J.P. Padolewski, Method for October 1996), assigned to Regents of the University of
preparing vanadium antimony oxide-based oxidation and Minnesota.
ammoxidation catalysts, U.S. Patent 5,693,587 (2 December [55] C. Yokoyama, S.S. Bhardadwaj, L.D. Schmidt, World Patent
1997), assigned to The Standard Oil Company of Ohio 97/26987 (31 July 1997), assigned to Regents of the
(Sohio/BP). University of Minnesota.
M.M. Bhasin et al. / Applied Catalysis A: General 221 (2001) 397–419 419

[56] A.S. Bodke, D.A. Olschki, L.D. Schmidt, E. Ranzi, Science [71] S. Wang, K. Murata, T. Hayakawa, S. Hamakawa, K. Suzuki,
285 (1999) 712. Chem. Lett. 25 (1999).
[57] M. Huff, L.D. Schmidt, in: M.M. Bhasin, D.W. Slocum (Eds.), [72] P. Viparelli, P. Ciambelli, L. Lisi, G. Rupooplo, G. Russo,
Methane and Alkane Conversion Chemistry, Vol. 227, Plenum J.C. Volta, Appl. Catal. A 184 (1999) 291.
Press, New York, 1995. [73] D. Creaser, B. Andersson, R.R. Hudgins, P.L. Silverston,
[58] A.S. Bodke, L.D. Schmidt, Catal. Lett. 63 (1999) 113. Chem. Eng. Sci. 54 (1999) 4365.
[59] S.S. Bharadwaj, L.D. Schmidt, J. Catal. 155 (1995) 403. [74] Z.M. Fang, Q. Hong, Z.H. Zhou, S.J. Dai, W.Z. Weng, H.L.
[60] D.W. Flick, M.C. Huff, J. Catal. 178 (1998) 315. Wan, Catal. Lett. 61 (1999) 39.
[61] D.I. Iordanoglou, A.S. Bodke, L.D. Schmidt, J. Catal. 187 [75] R. Rulkens, T.D. Tilley, J. Am. Chem. Soc. 120 (1998)
(1999) 400. 9959.
[62] L. Xu, L. Lin, Q. Wang, L. Yan, D. Wang, W. Liu, in: A. [76] A. Beretta, P. Forzatti, E. Ranzi, J. Catal. 184 (1999) 469.
Parmaliana, et al. (Eds.), Natural Gas Conversion V, Elsevier, [77] V.R. Choudhary, V.H. Rane, A.M. Rajput, AIChE J. 44 (1998)
Amsterdam, Stud. Surf. Sci. Catal. 119 (1998) 605. 2293.
[63] K. Nakagawa, M. Okamura, N. Ikenaga, T. Suzuki, S. Wang, [78] M.J. Alfonso, A. Julbe, D. Farrusseng, M. Menéndez, J.
Chem. Commun. 1025 (1998). Santamarı́a, Chem. Eng. Sci. 54 (1999) 1265.
[64] S. Wang, K. Murata, T. Hayakawa, S. Hamakawa, K. Suzuki, [79] D. Creaser, B. Andersson, R.R. Hudgins, P.L. Silverston,
Chem. Lett. 569 (1999). Chem. Eng. Sci. 54 (1999) 4437.
[65] M. Kralik, V. Macho, Petrol. Coal 40 (1998) 115. [80] G.E. Keller, M.M. Bhasin, Synthesis of ethylene via oxidative
[66] V. Macho, M. Kralik, E. Jurecekova, L. Jurecek, Petrol. Coal coupling of methane. 1. Determination of active catalysts, J.
39 (1997) 46. Catal. 73 (1999) 9.
[67] M. Sugino, H. Shimada, T. Turuda, H. Miura, N. Ikenaga, T. [81] M.M. Bhasin, Feasibility of ethylene synthesis via catalytic
Suzuki, Appl. Catal. A 121 (1995) 125. oxidative coupling of methane, in: paper presented at the
[68] Takahara, M. Saito, Chem. Lett. 973 (1966). Methane Conversion Symposium, Auckland, New Zealand,
[69] H. Jachow, World Patent 99/42404 (26 August 1999), assigned 27 April to 1 May 1987.
to BASF. [82] M.M. Bhasin, Feasibility of Ethylene Synthesis via Oxidative
[70] G. Descat, World Patent 98/24742 (11 June 1998), assigned Coupling of Methane, Elsevier, Amsterdam, Stud. Surf. Sci.
to ELF Atochem. Catal. 36 (1988) 343.

You might also like