You are on page 1of 13

Applied Catalysis A: General 221 (2001) 253–265

Recent advances in processes and catalysts


for the production of acetic acid
Noriyuki Yoneda a , Satoru Kusano a,∗ , Makoto Yasui b , Peter Pujado c , Steve Wilcher c
a Chiyoda Corporation, 3-13 Moriya-cho, Kanagawa-ku, Yokohama 221-0022, Japan
b Chiyoda Corporation, 2-12-1 Tsurumichuo, Tsurumi-ku, Yokohama 230-8601, Japan
c UOP LLC, 25 East Algonquin Road, Des Plaines, IL 60017-5017, USA

Abstract
Novel acetic acid processes and catalysts have been introduced, commercialized, and improved continuously since the
1950s. The objective of the development of new acetic acid processes has been to reduce raw material consumption, energy
requirements, and investment costs. At present, industrial processes for the production of acetic acid are dominated by
methanol carbonylation and the oxidation of hydrocarbons such as acetaldehyde, ethylene, n-butane, and naphtha. This paper
discusses advances in acetic acid processes and catalysts according to the following routes: (1) methanol carbonylation; (2)
methyl formate isomerization; (3) synthesis gas to acetic acid; (4) vapor phase oxidation of ethylene, and (5) other novel
technologies. © 2001 Published by Elsevier Science B.V.
Keywords: Acetic acid; Methanol carbonylation; Hydrocarbon oxidation; Reaction mechanisms

1. Introduction is potentially a great demand for acetic acid in this


market.
Acetic acid is an important commodity chemical The total world capacity of acetic acid has reached
used in a broad range of applications. As shown in approximately 7.8 million t in 1998 with BP-Amoco
Fig. 1, acetic acid is used primarily as a raw ma- and Celanese accounting for more than 50% of
terial for vinyl acetate monomer (VAM) and acetic the world’s capacity [1]. BP-Amoco and Celanese
anhydride synthesis, and as a solvent for purified have installed capacities of 1.5 million t (19%), and
terephthalic acid (PTA) production. The demand for 2.0 million t (26%), respectively.
acetic acid has increased, especially in southeast Asia,
where several new PTA plants have been built. With
the increased demand and installed capacity for PTA
2. Processing routes to acetic acid
in southeast Asia, the region has become a major
producer of polyester (PET) fiber, film, and resin.
Originally, acetic acid was produced by aerobic fer-
Although the economic crisis in Asia momentarily
mentation of ethanol, which is still the major process
suppressed the demand for acetic acid to less than
for the production of vinegar. The first major com-
expected levels, in the medium and long terms there
mercial process for the synthetic production of acetic
acid was based on the oxidation of acetaldehyde. In
∗ Corresponding author. Tel.: +81-45-441-9151; an early process for the conversion of acetylene to ac-
fax: +81-45-441-1281. etaldehyde introduced in 1916 in Germany and used

0926-860X/01/$ – see front matter © 2001 Published by Elsevier Science B.V.


PII: S 0 9 2 6 - 8 6 0 X ( 0 1 ) 0 0 8 0 0 - 6
254 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

process were 90, and 70% based on methanol and CO


consumption, respectively, [2]. In 1970, Monsanto
commercialized an improved homogeneous methanol
carbonylation process using a methyl-iodide-promoted
Rh catalyst [3–6]. Compared to other acetic acid syn-
thesis routes (ethanol fermentation, and acetaldehyde,
n-butane, or naphtha oxidation), homogeneous Rh
catalyzed methanol carbonylation is an efficient route
that exhibits high productivity and yields. The pro-
cess operated at much milder conditions (180–220 ◦ C,
30–40 atm) than the BASF process and exhibited su-
perior performance: acetic acid yields were 99 and
Fig. 1. Use of acetic acid.
85% based on methanol and CO consumption, re-
spectively, [7]. Celanese and Daicel further improved
in China until recently, an organo-mercury compound the Monsanto process during the 1980s by adding
was used as the catalyst. The toxicity of the mercury a lithium or sodium iodide promoter to enable the
catalyst resulted in significant environmental pollu- operation in a reduced water environment [8–15]. At
tion, and as a result, has essentially been phased out. lower water concentrations, by-product formation via
As the petrochemical industry developed in the 1950s, the water gas shift reaction is reduced, thus improving
the raw material for the production of acetaldehyde raw materials consumption and reducing downstream
shifted to ethylene. Other processes for the production separation costs.
of acetic acid introduced in the 1950s and 1960s were Homogeneous metal catalysts less costly than Rh
based on the oxidation of n-butane or naphtha. The (for example, Ni [16,17,75,76] and Ir [3,18–24] with
major producers of acetic acid via direct oxidation of other metal additives) have also been investigated.
hydrocarbons were Celanese (via n-butane) and BP The Ir-based process allows operation at reactor water
(via naphtha). However, these reactions also produce levels comparable to those of the improved Celanese
significant amounts of oxidation by-products, as sum- process and was commercialized by BP Chemicals in
marized in Table 1, and their separation and recovery 1996.
can be very complex and expensive. Until recently, virtually all new acetic acid ca-
The homogeneous methanol carbonylation route to pacity has made use of the homogeneous methanol
acetic acid that used a homogeneous Ni catalyst was carbonylation technology developed by Monsanto
first commercialized by BASF in 1955. An improved and practiced commercially by all major acetic acid
process was later disclosed by BASF in 1960. The manufacturers, including BP-Amoco, Celanese, and
process used an iodide-promoted CO catalyst and others. As a result, more than 60% of the world acetic
operated at elevated temperature (230 ◦ C) and pres- acid production employs the methanol carbonylation
sure (600 atm). The product yields exhibited by this methods, as shown in Fig. 2.

Table 1
Acetic acid process
Catalyst Reaction condition Yield By-product
(◦ C, atm)
Methanol carbonylation Rhodium complex 180–220, 30–40 MeOH: 99%, CO: 85% None
Acetaldehyde oxidation Manganese acetate 50–60, atmospheric CH3 CHO: 95% None
or cobalt acetate pressure
Ethylene direct oxidation Palladium/heteropolyacid/ 150–160, 80 Ethylene: 87% Acetaldehyde CO2
metal
Hydrocarbon oxidation Cobalt acetate or 150–230, 50–60 n-Butane: 50%, Formic acid,
(n-butane, naphtha) manganese acetate naphtha: 40% propionic acid, etc.
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 255

acid process based on butylene feedstock. This process


also employs direct oxidation. Its key features are the
use of a relatively cheap raffinate-2 feedstock and com-
petitive economics in medium size plants. Recently,
Poulenc and others have disclosed the direct produc-
tion of acetic acid from ethane; there are no indica-
tions of impending commercialization for this route.
Generally, the production cost of commodity chem-
icals such as acetic acid is dominated by the raw mate-
rial costs, and methanol carbonylation is still regarded
as the preferred route to produce acetic acid. Table 1
Fig. 2. Acetic acid process routes. summarizes reaction conditions, catalysts and yields
for the major processes used to produce acetic acid. A
number of reviews on production of acetic acid have
Inherent to the homogeneous system, however, are been published and are referred [7,29,73,74].
drawbacks relating to catalyst solubility limitations
and the loss of expensive Rh metal due to precipi-
tation in the separation sections. Accordingly, immo- 3. Methanol carbonylation
bilization of the Rh complex on a support has been
the subject of considerable investigation. Chiyoda and 3.1. Rhodium catalyzed methanol carbonylation
UOP have jointly developed an improved methanol
carbonylation process for the production of acetic acid The methanol carbonylation process, “Mon-
based using a heterogeneous Rh catalyst system [25]. santo process”, is operated under mild conditions
A direct oxidation process for the production of (180–220 ◦ C, 30–40 atm) and exhibits high selectivity
acetic acid starting from ethylene was commercialized to acetic acid based on methanol (99%) and carbon
by Denko in 1997. While the raw material, ethylene, monoxide (85%) [7]. While the reaction, as shown
is more expensive than in the methanol carbonylation below, can be carried out in a variety of rhodium
route, the investment cost is reported to be lower and (I) or rhodium (III) complexes [6,18], under reaction
competitive for small or medium-size capacity plants. conditions they are almost invariably converted to the
Wacker-Chemie plans to commercialize a new acetic active catalyst [RhI2 (CO)2 ]−1 . As shown in Fig. 3,

Fig. 3. Catalytic cycle for rhodium carbonylation.


256 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

methyl iodide is provided by the reaction of feed routes are active since more propionic acid is observed
methanol with hydrogen iodide: than can be accounted for by only this mechanism.
The rhodium catalyst system can generate acetalde-
Rh complex
CH3 OH + CO → CH3 COOH hyde, and it is proposed that this acetaldehyde is
reduced by hydrogen in the system to give ethanol
Methyl iodide is oxidatively added to the rhodium–
which subsequently yields propionic acid. One possi-
dicarbonyl–diiodide complex [RhI2 (CO)2 ]−1 (A)
ble precursor for the generation of acetaldehyde is the
to generate a rhodium–methyl complex (B). This
rhodium–acetyl species, as shown in the following
rhodium–methyl complex can rapidly undergo a
mechanism [28]:
methyl migration to a neighboring carbonyl group in
the acetyl form (CH3 CO) and react with CO (C) to [RhI3 (CO)(COCH3 )]− + HI
generate the rhodium–acetyl complex (D). Reductive
→ [RhI4 (CO)]− + CH3 CHO
elimination of acetyl iodide (CH3 COI) can then lib-
erate the original rhodium complex (A). Hydration of [RhI4 (CO)]− → RhI3 + I− + CO
acetyl iodide is very rapid in the presence of excess
Reaction of this species with hydrogen iodide would
water and will result in the formation of acetic acid
yield acetaldehyde and [RhI4 (CO)]−1 . The latter
and hydrogen iodide to complete the cycle.
species is well known in this system and is postulated
The reaction rate is independent of methanol
as the principal cause of catalyst loss by precipitation
concentration and carbon monoxide pressure. The
of inactive rhodium tri-iodide [28].
rate-determining step is believed to be the oxidative
Acetaldehyde undergoes self-condensation or aldol
addition of methyl iodide to the rhodium center of the
condensation and yields butenal and higher aldehy-
rhodium complex (A), and the reaction rate is essen-
des. These can undergo further reactions to alcohols
tially of first order in both catalyst and methyl iodide
and carboxylic acids as summarized in the network of
concentrations under normal reaction conditions:
Fig. 5 [28]. It would be expected that the homologa-
reaction rate ∝ [catalyst][CH3 I] tion observed would result in unsaturates and iodides
having an even number of carbon atoms, and long
A substantial quantity of water (14–15 wt.%) is re- chain carboxylic acids with an odd number of carbon
quired to achieve high catalyst activity and also to atoms. Particular problems are encountered with the
maintain good catalyst stability [8,9,12–14]. However, C6 species present. The boiling points of the unsatu-
as rhodium also catalyzes the water gas shift reaction rated compounds, including hexanal and some of its
(Fig. 4), the side reaction forming CO2 and H2 from isomers, are very similar to that of acetic acid. Further-
CO is significantly affected by water and hydrogen more, hexyl iodide is observed to form a constant boil-
iodide concentration in the reaction liquid [26,27]. ing azeotropic mixture with acetic acid. The presence
Propionic acid is observed as the major liquid of the unsaturates, even at low parts per million con-
by-product in this process. This is produced by the centrations can cause problems with product stability.
carbonylation of ethanol that is often present as a Separation of pure acetic acid product from the re-
minor impurity in the methanol feed; however, other action medium presents few problems. In this process,
however, the expensive Rh metal can be lost due to its
precipitation and vaporization in the flash column. A
schematic of a conventional methanol carbonylation
plant configuration is shown in Fig. 6 [28]. Rhodium
catalyst is separated from the product acetic acid by
conducting a simple flash; the catalyst remains in the
liquor and can be recycled to the reactor. The sepa-
ration of light compounds, such as methyl iodide and
methyl acetate, may be carried out in the first distil-
lation column. This column is followed by a drying
Fig. 4. Mechanism for water gas shift. column and then a column for the removal of heavy
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 257

Fig. 5. Network of liquid by-products.

by-products. Energy usage in this fractionation train Group I metal iodides, especially lithium iodide in
can be high, depending on the concentration of water combination with methyl iodide, were identified early
and impurities such as propionic acid, heavy unsatu- as a good agent for enhancing the stability of the
rates, and hexyl iodide present. rhodium catalyst at low reactor water concentrations
In the Monsanto process, because of the high water (4–5 wt.%), and also for decreasing liquid by-product
concentration in the reactor (14–15 wt.%), the separa- formation [12–14]. Further work in this area revealed
tion of water from the acetic acid product is a major that the addition of a substantial quantity (16–20 wt.%)
energy consumer and can limit the unit capacity. In of group I metal iodides also enhanced the reactor
addition, excess water causes carbon monoxide yield productivity even at quite low water concentrations
loss due to the water gas shift reaction, and increases (2 wt.%) [8–11]. These features reportedly allow exist-
the formation of by-products such as propionic acid, ing plants to expand their capacity for little incremen-
thus lowering the acetic acid quality. Considerable sav- tal capital cost. The improved methanol carbonylation
ings in operating costs can be realized by operating process, “low water process”, effected by adding
at low water concentration if a way can be found to group I metal iodides to the Monsanto process was first
compensate for the consequent decrease in the reac- commercialized in the 1980s by Celanese and Daicel.
tion rate and catalyst stability [29]. As a result, the In this process, it is proposed that the addition of
rhodium complex stability at low water concentrations a significant quantity of group I metal causes the Rh
has been extensively investigated. complex to be more coordinated by CH3 COO− and

Fig. 6. Schematic of a acetic acid plant configuration.


258 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

with alkali-iodide promoters results in a higher iodide


environment, and higher residual iodide in the final
product. High iodide concentration in acetic acid leads
to catalyst poisoning problems in some downstream
applications, such as in the manufacture of VAM. To
overcome the problems associated with high iodide
concentration in the final product, treatment by active
carbon [30], hydrogenation [31,32], and extra distilla-
tion [33,34] have been proposed. Celanese disclosed
the “silver-guard process” for the removal of very low
levels of iodide impurities from acetic acid in their
Fig. 7. Reaction acceleration mechanism by iodide salt.
patent [35]. The use of silver metal on an ion exchange
resin such as Amberlyst-15 reduces the iodide level
to below 1 ppb, as opposed to 20 ppb more normally
increases the rate of oxidative insertion of methyl io-
achieved by conventional methods. One particular
dide (the rate determining step), thus promoting the
advantage of this system is the ability to effectively
primary carbonylation reaction (Fig. 7). As Figs. 7 and
remove the halide impurity in a single step, thus avoid-
8 shows the effect of the addition of lithium iodide
ing the need for additional distillation and recovery.
on the reaction rate, the overall carbonylation rate in-
crease is presumably due in part to the formation of
a strong nucleophilic five-coordinate dianionic inter- 3.2. Nickel catalyzed methanol carbonylation
mediate [Rh(CO2 )2 I2 L]2− (L− = I− , OAc− ) which
Recent studies have shown that nickel catalysts
is more active toward oxidative addition of methyl io-
can operate under mild conditions (190 ◦ C, 70 atm)
dide [8–11,29].
with the addition of methyl iodide as a co-promoter
The main advantages of the low water process rela-
[16]. The activity of nickel catalyst systems can be
tive to the conventional Monsanto process are reduced
increased and the volatility of nickel carbonyl com-
raw materials consumption, increased productivity,
pounds lowered by the introduction of stabilizers such
lower utility requirements, and lower capital costs
as phosphines, alkali metals, tin, and molybdenum
per unit of product. However, low water operation
[16,17,25,75,76]. The active catalysts are thought to
be Ni(0) complexes. For phosphine-promoted cata-
lyst, Ni(PR3 )2 is considered an active form of catalyst
and, in addition, Ni(CO)4 was observed in all cases,
and its concentration was reduced by strongly coordi-
nating ligands and enhanced by weakly coordinating
ligands [76]. Recent work on nickel catalyst systems
shows that reaction rates and selectivities can ap-
proach those achieved in the rhodium catalyst system.
Although nickel catalysts have the advantage of being
much cheaper than rhodium, and are easy to stabi-
lize at low reactor water concentrations, [Ni(CO)4 ] is
known to be a very toxic and volatile compound. To
date, commercialization has not proceeded.

3.3. Iridium catalyzed methanol carbonylation

Fig. 8. Effect of Li salt addition. Reaction condition: The potential use of iridium instead of rhodium was
[CH3 ] = 1.0 M, [MeOAc] = 0.3 M, [H2 O] = 1.0 M, temperature identified as part of the early work done by Monsanto
= 190 ◦ C, total pressure = 400 psig. [3,18,25], however, the reaction rate exhibited by the
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 259

rhodium catalyst system was superior to that of irid- that the oxidative addition of methyl iodide to the
ium. Recently, it was disclosed that an improved irid- iridium center is of the order of 150 times faster
ium catalyst, in combination with a promoter metal than the equivalent reaction with rhodium [36]. This
such as ruthenium, has higher activity and results in represents a possible improvement in the available
lower product impurity levels than reported in pre- reaction rates, as methyl iodide addition is not the
vious iridium systems [19]. The production of acetic rate determining step. The slowest step in this cycle is
acid using the iridium catalyst system has been com- the insertion of carbon monoxide to form the iridium
mercialized by BP-Amoco in two world scale plants to acetyl species, that involves the elimination of ionic
date, and has received wide publicity as the “Cativa” iodides and the coordination of an additional carbon
process. Although much iridium is required to achieve monoxide ligand. This would suggest the following
an activity comparable to the rhodium catalyst-based expression. The dependence on ionic iodide:
processes, the catalyst system is able to operate at re-
[CO]
duced water levels (less than 8 wt.% for the Cativa reaction rate ∝ [catalyst]
process versus 14–15 wt.% for the conventional Mon- [I− ]
santo process). Thus, lower by-product formation and suggests that high reaction rates should be achiev-
improved carbon monoxide efficiency are achieved, able by operating at low iodide concentrations. It
and steam consumption is decreased. Until the early also suggests that the inclusion of species capable of
1990s, the difference in the prices of rhodium (US$ assisting in the abstraction of iodide should promote
500/oz) and iridium (US$ 60/oz) was the driving force the rate-limiting step. The patent would suggest that
for replacing rhodium with iridium. However, current ruthenium, or rhenium are the preferred promoters
price increases for iridium (US$ 450/oz) negate the [20,21]. In effect, a proprietary blend of promoters
advantage in catalyst price. has been found to increase reaction rate. The above
The unique differences between the rhodium cat- expression does not imply any effect from the water
alytic cycle and that of iridium in methanol car- present in the matrix, but water is found to have a
bonylation have been investigated [36]. The anionic significant effect on rate [22].
iridium cycle shown in Fig. 9, is similar to that shown In the improved iridium system, low water concen-
earlier for rhodium. Model studies have demonstrated tration in the reactor results in the formation of fewer

Fig. 9. Catalytic cycle for iridium carbonylation.


260 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

by-products such as propionic acid than in the original development of a commercial heterogeneous Rh cata-
Monsanto rhodium system, and no addition of lithium lyst system by Chiyoda, no successful demonstration
iodide is required. Consequently, the iridium catalyst of such a catalyst had been known [7].
system is also characterized by the formation of fewer The heterogeneous catalyst commercialized for the
higher alkyl-iodide species than in the conventional acetica process consists of Rh complexed on a novel
low water process. poly-vinyl pyridine resin [50], which is tolerant of
elevated temperatures and pressures. Under reaction
3.4. Heterogeneous rhodium catalyzed conditions, the Rh is converted to its catalytically
methanol carbonylation active anion form [Rh(CO)2 I2 ]−1 . Furthermore, the
nitrogen atoms of the resin pyridine groups become
In order to overcome the limitations of the ho- positively charged after quaternization with methyl
mogeneous catalyst system (e.g. Rh precipitation iodide. Thus, the strong ionic association between the
and catalyst solubility limitations), the immobiliza- pyridine nitrogen groups and the Rh complex causes
tion of the Rh complex on a support has been the the immobilization (Fig. 10). The concentration of Rh
subject of considerable investigation. Active carbon on the solid phase is determined by the ion exchange
was investigated as a possible support and proposed equilibrium. Because equilibrium strongly favors the
for vapor-phase operation [7,37,38]. However, the solid phase, virtually all the Rh in the reaction mixture
reaction rate was 1/1000–1/10 that of Monsanto’s ho- is immobilized.
mogeneous process and selectivity was also poorer. In the acetica process, the methanol carbonyla-
Inorganic oxides and zeolites were also investigated tion reaction is conducted at moderate temperature
for use in vapor-phase operation [39,40]. For exam- (160–200 ◦ C) and pressure (30–60 atm) and at low
ple, attaching the Rh–phosphine ligand complex to water concentration without any additives present.
alumina by silylation was attempted [41,42]. The Catalyst stability has been demonstrated in both once-
resultant reaction rates for these catalysts were also through and continuous-recycle pilot plant testing at
found to be poor relative to those observed for the process conditions, low water content, and no Rh or
homogeneous system. To increase catalyst activity resin makeup. The catalyst exhibited no deactivation
for operation in the liquid phase, ion exchange resins after continuous operation for more than 7000 h [50].
based on cross-linked polystyrene and incorporating With homogeneous methanol carbonylation routes,
pendant phosphines, or vinyl pyridine copolymers acetic acid productivity is directly proportional to
have been evaluated [43–45]. Although the activity of catalyst concentration in the reaction liquid, and as
these catalysts in the liquid phase was comparable to a result, acetic acid production is restricted by the
Monsanto’s homogeneous catalyst, there were prob- solubility of the active metal. Limited success has
lems with rhodium metal leaching from the resins been achieved in improving catalyst solubility in
and the decomposition of the resins during opera- these systems by increasing the reaction-mixture wa-
tion at elevated temperature. Vinyl pyridine resin was ter concentration or by adding iodide salt stabilizers
known to be more robust and more tolerant of oper- [8,9,12–14]. Both additives, however, result in in-
ation at elevated temperature relative to polystyrene creased recycle and separation costs, higher corrosion
resins. It was disclosed that catalysts using pyridine rates, and difficulty in product purification.
resins exhibited high tolerance to operation at ele- With the heterogeneous catalyst system, catalyst
vated temperature and pressure, and higher reaction solubility limitations no longer govern reactor capac-
rate than Monsanto’s rhodium system [46]. Further- ity since catalyst concentrations several times greater
more, Chiyoda introduced novel pyridine resins and than those achievable in the homogeneous systems
catalysts that exhibited high activity, long catalyst are possible. Immobilization also significantly re-
life, and no significant rhodium loss [47–49]. Based duces the loss of expensive Rh metal because the
on this heterogeneous Rh catalyst, Chiyoda and catalyst is confined to the reactor rather than circulat-
UOP have jointly developed an improved methanol ing downstream, where reduced pressures may cause
carbonylation process, called “the acetica process”, precipitation of rhodium and vaporization losses of
for the production of acetic acid. Until the recent metal carbonyl compounds. The lower water content
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 261

Fig. 10. Rhodium immobilization.

of 3–7 wt.% typical of the acetica process results seals and was designed to maximize the performance
in reduced production of CO2 , and hydrogenated of its unique heterogeneous catalyst system without
by-products via the water gas shift reaction. Also, any rotating equipment (Fig. 11). Methanol and CO
because of the lower water content, less hydrogen feeds are introduced at the reactor bottom, where the
iodide is present in the system, and consequently the compressed CO gas is distributed through a sparger.
process environment is less corrosive. Both of these feeds, along with the recycle liquid
While the continuously stirred tank reactors (CSTR) and catalyst, flow up the reactor riser, where the
used in the conventional homogeneous processes CO is consumed in the reaction. The process flow,
can be limited by gas solution rates to liquid and which is similar to that of a conventional homoge-
are often prone to mechanical problems, the bubble neous process is shown in Fig. 11. In cases where the
column, or gas lift reactor employed with the hetero- acetic acid product will be used for VAM production,
geneous catalyst process does not suffer from such novel iodide removal technology is available to re-
problems and limitations. The acetica three-phase duce the iodide in the acetic acid product to less than
gas lift reactor has no moving parts or mechanical 3 ppb [51].

Fig. 11. Bubble column reactor and acetica process flow.


262 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

4. Methyl formate isomerization etc.). The reaction is carried out at approximately


220 ◦ C and 40 atm:
It has been proposed that acetic acid can be pro-
CO + 2H2 → CH3 OH
duced by isomerization of methyl formate in the pres-
ence of a homogeneous rhodium catalyst together with 2CH3 OH → CH3 OCH3 + H2 O
other metal additives [52,53]. Heterogeneous rhodium
catalysts supported on poly-vinyl pyridine resin have H2 O + CO → CO2 + H2
also been proposed for this application [54]. This cata- In the acetic acid synthesis step, carbonylation of
lyst has the same chemical morphology as a methanol DME and methanol to acetic acid is carried out by
carbonylation catalyst. Methyl formate is produced by the rhodium carbonyl complex catalyst with carbon
dehydrogenation of methanol [55] or by methanol car- monoxide being supplied from the synthesis gas
bonylation under high pressure in the presence of: process unit:
HCOOCH3 → CH3 COOH CH3 OH + CO → CH3 COOH

copper oxide and alkali catalyst. It is noted that CH3 OCH3 + 2CO + H2 O → 2CH3 COOH
acetic acid production via methanol dehydrogena- Carbonylation reaction conditions of 170–250 ◦ C and
tion followed by methyl formate isomerization re- 25–50 atm, can be used to obtain acceptable reaction
quires only methanol and no carbon monoxide rates in the liquid phase.
plant:CH3 OH→HCHO→HCOOCH3 →CH3 COOH-
Acetic acid can be produced from only methanol us-
ing a Ru-Sn catalyst according to the following steps 6. Vapor phase oxidation of ethylene
[56,57]. Ru-Sn bimetallic complexes are proposed to
be the active species. The two-step oxidation process for the production
of acetic acid, starting from ethylene through acetalde-
hyde, was first commercialized in 1960:
5. Synthesis gas route to acetic acid
CH2 =CH2 + 21 O2 → CH3 CHO
A nearby synthesis gas plant to produce CO is nor- CH3 CHO + 21 O2 → CH3 COOH
mally required to provide feed to an acetic acid plant.
On the contrary, an efficient integrated synthesis This route involves the liquid phase oxidation of ac-
gas and methanol synthesis plant and acetic acid plant etaldehyde using air and typically a manganese ac-
are available by combination of current technology at etate catalyst operating at 50–60 ◦ C. The reaction is
the natural gas source. This integrated process could based on a free radical mechanism. Although this pro-
achieve a significant capital cost reduction relative to cess features high yield (approximately 90%) and a
the conventional flow scheme. relatively low capital investment cost, it suffers from
Applying this concept, Haldor Topsoe proposed high acetaldehyde feedstock cost and a very corrosive
an integrated process that includes the synthesis of catalyst system. Many plants utilizing this technology
methanol and dimethyl ether (DME) in a first catalytic have been shut down over the last 20 years.
reaction stage and the subsequent carbonylation of There is also an older process that entails liquid
methanol and DME into acetic acid [58,59]. Although phase free radical oxidation of n-butane or naphtha
the reaction pressure required for methanol synthesis in the C4 –C8 range. These reactions produce a wide
is higher than the pressure used in acetic acid syn- spectrum of oxidation by-products such as formic acid
thesis, the combination of methanol synthesis with and propionic acid:
dimethyl ether synthesis can reduce the pressure of
CH3 CH2 CH2 CH3 +O2 →CH3 COOH + by-products
the first reaction step. The catalyst consists of a mix-
ture of the catalyst for methanol synthesis (Cu-Zn-Al The direct production of acetic acid from ethylene via
oxide, etc.) and a dehydration catalyst (H-ZSM-5, an acetaldehyde intermediate is a desirable synthesis
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 263

enhance the activity and selectivity of the process for


the production of acetic acid. The selectivity to acetic
acid is approximately 86%, since it appears that ac-
etaldehyde and carbon dioxide are necessarily formed
in this type of process.
Scheme 1. Hydration route.

7. Other proposed technologies for the


route that has yet to be developed. Much work has been production of acetic acid
undertaken to develop a simpler, single stage process
for producing acetic acid directly from ethylene: 7.1. Ethane oxidation
catalyst
CH2 =CH2 + O2 → CH3 COOH In the 1980s, an acetic acid route from ethane was
Various groups have carried out extensive research and introduced. Two reaction mechanisms based on:
development in the area of direct vapor phase oxida- catalyst
tion of ethylene to acetic acid. Catalyst systems con- CH3 CH3 + O2 → CH3 COOH + by-product
sisting of palladium chloride and V2 O5 supported on
different catalyst systems were proposed: (1) partial
Al2 O3 [60], and combinations of Pd (2%) and H3 PO4
oxidation of the methyl group, and (2) ethane oxi-
(25%) on SiO2 , Pd-V2 O5 -Sb2 O3 on Al2 O3 [61], or
dation to ethylene followed by ethylene hydration to
Pd (1%) on V2 O5 [62] have been proposed. These
ethanol, or ethylene to acetaldehyde.
catalysts have acetic acid selectivities in the range of
A patent refers to the production of acetic acid
60–90% based on ethylene. The routes have been pro-
by reacting ethane, ethylene, or mixtures of ethane
posed according to the different catalyst systems in
and ethylene with oxygen over a catalyst containing
Schemes 1 and 2.
molybdenum, vanadium, and one other metal (Z) in
Denko has developed a direct oxidation process for
the general formula Mox Vy Zz [66]. In one example,
the production of acetic acid based on the hydration
the patent describes the gas phase oxidation of a 1/10
route [63–65] and has commercialized this technol-
mixture of ethane and ethylene at 255 ◦ C over a vana-
ogy in late 1997. The catalyst consists of either two
dium catalyst containing lesser amounts of molybde-
or three components. The first component is palla-
num, niobium, antimony, and calcium supported on
dium supported on a carrier, preferably in the range of
an LZ-105 molecular sieve to yield 63% selectivity
0.1–2% range. The second component is a heteropoly
to acetic acid, and 14% selectivity to ethylene at 3%
acid and their salts, preferably phosphotungstic acid
ethane conversion. In the combined ethane/ethylene
salts of lithium, sodium, and copper. The third com-
feed case, the hydration catalyst further catalyzes the
ponent is copper, silver, tin, lead, antimony, bismuth,
hydration of ethylene to ethanol, which is then con-
selenium, or tellurium.
verted to acetic acid (Scheme 1). The oxidation cat-
The reaction takes place in a fixed bed reactor at
alyst catalyzes the reaction of ethylene to acetic acid
operating temperatures and pressures of 150–160 ◦ C,
and other oxidation products that are converted to
and up to 8 atm, respectively. The gases fed to the re-
acetic acid (Scheme 2).
actor are ethylene, oxygen, steam, and nitrogen that is
In another catalyst system, rhenium or a combi-
used as a diluent. The presence of steam is required to
nation of rhenium and tungsten are introduced to
replace the molybdenum in the dehydrogenation cat-
alyst [67]. Tests showed that complete substitution
of molybdenum by rhenium (Rex Vy Zz ) is beneficial
in the reaction of ethane to ethylene, whereas partial
substitution can increase the selectivity to acetic acid.
Tests were not performed on ethylene feed, but tests
Scheme 2. Partial oxidation route. on ethane (21% ethane, 3.8% oxygen, and 75.2%
264 N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265

nitrogen) resulted in acetic acid selectivity as high as acid in the presence of rhodium trichloride dissolved
78% at an ethane conversion of 14.3%. in water [72]:
More recently in 1998, another oxidative process
RhCl3
and catalyst for the production of acetic acid from CH4 + CO + 21 O2 → CH3 COOH
ethane, or ethylene was disclosed [68,69]. A new
molybdenum vanadate catalyst system promoted with This reaction proceeds in an aqueous medium at a
Nb, Sb, Ca, and Pd allows the gas phase oxidation of temperature of approximately 100 ◦ C and gives a high
ethane and/or ethylene to acetic acid, with high yield yield of acetic acid. The reaction rates are reported
and higher selectivity under milder operating condi- to be too slow for an economically viable industrial
tions than previously achieved. The patent discloses process, but this novel process route has the potential
the production of acetic acid with 86% selectivity and to reduce the cost of acetic acid production.
11% ethane conversion per pass, at a temperature and
pressure of 250–280 ◦ C and 15 atm, respectively.
In 1999, a catalyst for the co-production of ethylene 8. Conclusions
and acetic acid from ethane was disclosed [70]. It con-
sists of phosphorus-modified molybdenum-niobium Acetic acid represents a commodity chemical
vanadate of formula Mo2.5 V1.0 Nb0.32 Px in which the growing at 3.5–4.5% per year from a significant
optimum range for the phosphorus (x) is 0.01–0.06: and large base capacity. Significant developments in
both process and catalyst technology have supported
CH3 CH3 + O2 the growth in this market since the 1950s when the
catalyst first commercial synthetic process was introduced.
→ CH2 ==CH2 + CH3 COOH + H2 O Methanol carbonylation has emerged as the domi-
nant route to this product and currently over 60% of
Ethane and air (15:85 (v/v)) at 260 ◦ C and 200 psig the world acetic acid is produced using this route.
(1100/h GHSV) reacted over the above catalyst system However, significant catalyst innovation has occurred
(x = 0.042) to produce acetic acid and ethylene with even within this production route resulting in greatly
selectivities of 49.9, and 10.5%, respectively, at 53.3% improved yield, and selectivity at milder operating
conversion. At phosphorus levels greater than 0.06%, conditions and lower cost of production. The lucra-
there is a marked increase in ethylene production with tive nature of this market and the need for the major
a corresponding decline in acetic acid. producers to continually protect their market position
Recently, many attempts have been disclosed re- and investments is expected to drive further inno-
garding the use of ethane as feedstock. Although vation within methanol carbonylation and the other
ethane is a relatively inexpensive and attractive raw promising technology options looming on the horizon
material for producing acetic acid, the oxidation pro- that have been discussed in this paper.
cesses produce a variety of co-products, the disposi-
tion of which needs to be considered in any business
plan. References

7.2. Methane carbonylation [1] Asian Chemical News, 16 November 1998.


[2] N. Kutenpov, W. Himmele, H. Hohenshutz, Hydrocarbon
Proc. 45 (1966) 141.
Novel methods for producing acetic acid directly [3] F.E. Paulik, J.R. Roth, J. Am. Chem. Soc., Chem. Commun.
from methane under relatively mild conditions have (1968) 1578.
been reported. It was first disclosed that acetic acid [4] H.D. Grove, Hydrocarbon Proc. 51 (1972) 76.
can be produced from methane and carbon monox- [5] US Patent 3769329 (1973) to Monsanto.
[6] R.T. Eby, T.C. Singleton, Appl. Indian Catal. 1 (1983) 275.
ide in the presence of: Pd(OCOCH3 )2 /Cu(OCOCH3 )2 /
[7] M.J. Howard, M.D. Jones, M.S. Roberts, S.A. Taylor, Catal.
K2 S2 O8 /CF3 COOH [71]. Today 18 (1993) 325.
Secondly, it was reported that the mixture of [8] M.A. Murphy, B.L. Smith, G.P. Torrence, A. Agulio, J.
methane, carbon monoxide and oxygen formed acetic Organomet. Chem. 303 (1986) 257.
N. Yoneda et al. / Applied Catalysis A: General 221 (2001) 253–265 265

[9] B.L. Smith, G.P. Torrence, M.A. Murphy, A. Agulio, J. Mol. [45] J. Hjortkjaer, Y. Chen, B. Heinrich, Appl. Catal. 67 (1991)
Catal. 39 (1987) 115. 269.
[10] European Patent Application 0161874 (1985). [46] European Patent 277824 (1988) to Reilly.
[11] Japanese Patent Koukai 60-239434 (1985) to Celanese. [47] US Patent 5334755 (1994) to Chiyoda.
[12] Japanese Patent Koukoku 04-69136 (1992) to Daicel. [48] US Patent 5364963 (1994) to Chiyoda.
[13] Japanese Patent Koukoku 03-38256 (1991) to Daicel. [49] US Patent 5576458 (1996) to Chiyoda.
[14] Japanese Patent Koukoku 07-23337 (1995) to Daicel. [50] N. Yoneda, T. Minami, J. Weiszmann, B. Spehlmann,
[15] R.F. Heck, J. Am. Chem. Soc. 85 (1963) 2013. Science and technology in catalysis, in: Proceedings of the
[16] N. Ritzkalla, Industrial chemicals via C1 processes, Am. Third Tokyo Conference on Advanced Catalytic Science and
Chem. Soc. (1987) 61. Technology, 1998, p. 93.
[17] US Patent 4133963 (1979) to Eastman. [51] US Patent 5962735 (1999) to UOP.
[18] D. Foster, Adv. Organomet. Chem. 17 (1979) 255. [52] D.J. Schreck, D.C. Busby, R.W. Wegman, J. Mol. Catal. 47
[19] Japanese Patent Koukai 06-340573 (1994) to Poulenc. (1988) 117.
[20] European Patent Publication 643034 (1994) to BP. [53] G. Bud, H.U. Hog, J. Mol. Catal. A: Chem. 95 (1995) 1134.
[21] European Patent Publication 728726 (1994) to BP. [54] Japanese Patent Koukai 7-60130 (1995) to Chiyoda.
[22] European Patent Publication 752406 (1995) to BP. [55] M. Komatsu, M. Yoneoka, Kagaku Kougyo 43 (1988) 1134.
[23] Kagakukougyou Nippou, 12 July 1996. [56] S. Shinoda, T. Yamakawa, J. Chem. Soc., Chem. Comun.
[24] T.W. Dekleva, D. Forster, Adv. Catal. 34 (1986) 81. (1990) 1151.
[25] European Chemical News, 26 May–1 June 1997. [57] T. Yamakawa, P. Tsai, S. Shinoda, Appl. Catal. A L1 (1992)
[26] D. Forster, T.C. Singleton, J. Mol. Catal. 17 (1982) 299. 92.
[27] D. Forster, T.W. Dekleva, J. Chem. Edu. 63 (3) (1986) 204. [58] European Patent Application 801050 (1997).
[28] D.J. Watson, Catalysis of Organic reactions, Marcel Dekker, [59] US Patent 5728871 (1998) to Topsoe.
New York, 1998, p. 369. [60] British Patent 1020068 (1965) to Halcon.
[29] M. Gauss, A. Seidel, P. Torrence, P. Heymanns, Applied [61] Japanese Patent Koukai 47-13221 (1971) to National
Homogeneous: Catalysis with Organometallic Compounds, Distillers.
Vol. 1, VHC, New York, 1996, p. 104. [62] FR 1568742 (1967) IFP.
[30] Japanese Patent Koukai 7-503489 (1995) to Bohan Iron. [63] Japanese Patent Koukai 7-89896 (1995) to Denko.
[31] Japanese Patent Koukoku 5-28212 (1993) to Hoechst. [64] Japanese Patent Koukai 9-67298 (1995) to Denko.
[32] Japanese Patent Koukai 62-72636 (1987) to Hoechst. [65] K. Sano, H. Uchida, Shokubai 41 (4) (1999) 290.
[33] Japanese Patent Koukai 46-5367 (1971) to Monsanto. [66] European Patent Application 294845 (1988) to UCC.
[34] Japanese Patent Koukoku 60-54294 (1985) to Monsanto. [67] European Patent Application 407091 (1991) to BP.
[35] US Patent 4615806 (1986) to Celanese. [68] European Chemical News, 29 March–4 April 1999, p. 28.
[36] P.M. Maitlis, A. Haynes, G.J. Sunley, M.J. Howard, J. Chem. [69] World Patent 9805619 (1998) to Hoechst.
Soc., Dalton Trans. (1996) 2187. [70] US Patent 5907056 (1999) to Sabic.
[37] R.G. Schultz, P.D. Montgromerry, J. Catal. 13 (1969) 105. [71] T. Nishiguch, et al., Chem. Lett. 20 (1992) 1141.
[38] US Patent 3717670 (1973) to Monsanto. [72] Chem. Week 20 (1994).
[39] A. Krzywicki, M. Marczewski, J. Mol. Catal. 6 (1979) [73] V.H. Agreda, J.R. Zoeller, Acetic Acid and Its Derivatives,
431. Marcel Dekker, New York, 1993, Chapters 1–6, p. 3.
[40] M.S. Scurrell, R.F. Howe, J. Mol. Catal. 7 (1980) 535. [74] F.S.Wagner, Kirk-Othmer Encyclopedia of Chemical
[41] US Patent 4657884 (1987) to Hoechst. Technology, 4th Edition, Vol. 1, 1991, p. 121.
[42] US Patent 4776987 (1988) to Hoechst. [75] W.R. Moser, B.J. Marshik-Guerts, S.J. Okrasinski, J. Mol.
[43] R.S. Drago, E.D. Nyberg, A.E. Amma, A. Zombeck, Inorg. Catal. A: Chem. 143 (1999) 57.
Chem. 20 (1981) 641. [76] W.R. Moser, B.J. Marshik-Guerts, S.J. Okrasinski, J. Mol.
[44] M.S. Jarrel, B.C. Gates, J. Catal. 40 (1975) 255. Catal. A: Chem. 143 (1999) 71.

You might also like