You are on page 1of 3

Catalysis of the Carbonylation of Alcohols to Carboxylic

Acids Including Acetic Acid Synthesis from Methanol


Denis Forster and Thomas W. Dekleva
Central Research Laboratories, Monsanto Company, 800 North Lindbergh Boulevard, St. Louis, MO 63167

Acetic acid has a long history as an industrially important Table 1. Comparison of Cobalt- and Rhodium-Catalyzed
chemical; however, over the years there have been many Methanol Carbonylation Reactions
changes in the manufacturing process. During the 19th cen-
Cobalt process Rhodium process
tury, ethanol fermentation was the principal process. The
advent of coal-derived acetylene in the early decades of this
Metal concentration ~10~1 M ~10~3 M
century led to the development of the acetaldehyde route to Reaction temperature —230°C ~180°C
acetic acid based on the mercury-catalyzed acetylene hydra- Reaction pressure 500-700 atm 30-40 atm
tion reaction. Subsequently, the availability of low-cost eth-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Selectivity (on methanol) 90% >99%


ylene as a by-product of oil refining led to its exploitation via Hydrogen effect ch4, ch3cho, C2HsOH no effect
the Wacker process to give acetaldehyde. formed as by-products
Downloaded via UNIV GADJAH MADA on March 14, 2023 at 12:34:58 (UTC).

CH2=CH2 + PdCljj + HoO —


CHiCHO + Pd(0) + 2HC1 (1)

Culth.Oa steady-state conditions the predominant rhodium species


observed is [Rh(CO)2I2]~. The catalytic cycle can be repre-
sented as indicated in Figure 1. Critical to the catalysis is the
With the development of natural gas as a major fuel fol- formation of methyl iodide from methanol since the oxida-
lowing the Second World War, we have seen the evolution of tive addition of methyl iodide to the rhodium(I) center is the
large scale chemical processes based on synthesis gas derived rate-determining step. Because of this, iodide sources inca-
from natural gas reforming. pable of generating this organic intermediate in significant
quantities (e.g., alkali metal iodides) are ineffectual promot-
CH4 + H20 -
CO + 3H2 (2) ers for this reaction.
In particular, methanol is one of the largest-volume organ- Another key feature of the metal complex cycle in which
ic chemicals produced in the world today, and it is made via the iodide acts most effectively is the nature of the active
the heterogeneously catalyzed reaction of carbon monoxide catalyst itself. The oxidative addition step is considered to
and hydrogen. be nucleophilic in nature, based on activation parameters
and relative rate data (4), and the presence of a negative
CO + 2H2 —
CH3OH (3) charge on the metal center appears to significantly enhance
The availability of methanol in large volumes and its rela- the nucleophilicity (and hence reactivity towards methyl
tively low price make it an attractive raw material for other iodide) of this metal relative to neutral rhodium(I) species.
chemicals. In particular it is becoming the feedstock of
Another significant feature of the catalytic phenomenon
choice for acetic acid manufacture by utilizing carbonylation involving rhodium is the ease of methyl migration to form an
technology. acetyl-rhodium species (3). This occurs so rapidly that the
alkyl-rhodium species in Figure 1 has never been detected.
CH3OH + CO-+ CH3C02H (4) The short life of this species probably contributes to the high
The technology that was initially commercialized to conduct selectivity in the system since a long-lived alkyl-metal spe-
this reaction involved a homogeneous cobalt catalyst and
rather severe reaction conditions (~230°C and ~600 atm (1).
The severe reaction conditions have limited attempts to "Rh"
achieve an understanding of this reaction, and, while there
has been considerable speculation about mechanistic path- CH3C—I CH;,I
[Rhl2(CO)2J-
ways, it is probably fair to say that at this time we have no o
definitive information about the basic outline of the reac-
tion. [CH3CRhI3(CO)2]- o [CH;jRh(CO)2I3
In 1970, Monsanto unveiled a commercial plant in which VV II
V|CH3C—RhI3(CO)r
yj
the methanol carbonylation was conducted using a homoge- CO
neous rhodium catalyst in conjunction with an iodide pro-
moter (2). The reaction conditions are much milder than
those for the cobalt-catalyzed reaction (Table 1). This reac-
tion is much more amenable to mechanistic study than the
cobalt system, and a combination of in situ spectroscopic
studies and kinetic measurements has led to a good under-
standing of the reaction at the molecular level (3).
The reaction is characterized by kinetics that are indepen-
dent of CO pressure and are first-order in both rhodium and
methyl iodide. A variety of rhodium-containing complexes
can be used as the catalyst precursor, but in most cases the
same reaction is achieved since a common active species is
formed. In situ spectroscopic studies show that under Figure 1. Proposed mechanism for the Monsanto acetic acid synthesis.

204 Journal of Chemical Education


cies ina catalytic cycle could allow its diversion into nonpro- Table 2. Relative Rate Data for the Iodide-Promoted, Rhodium-
ductive pathways. Catalyzed Carbonylation of a Variety of Alcohols
Another reaction can be catalyzed with the reactants
present under the acetic acid process conditions, namely the Sn2 Displacement Rate
Alcohol Relative Rate for Organic Flalides (1 f)
water gas shift reaction.
CO + H20 Methanol 21 30

C02 + H2 (5)
Ethanol 1.0 1.0
The role of rhodium in catalyzing this reaction has been 1-Propanol 0.47 0.4
studied by workers at Monsanto (5) and Rochester Universi- 2-Propanol 1.2-3.0 0.02
ty (6). The basic steps in the cycle are illustrated in Figure 2
and it can be seen that the [Rh(CO)2I2]_ ion also plays a key
role in this catalytic cycle. The facile reaction between
then proceeds to regenerate the RhI2(CO)2_ and n-butyric
[Rh(CO)2l2]~ and HI in which H2 is liberated and the rho-
dium(I) species is oxidized to rhodium(III) is of critical im- acid, by a route identical to that proposed for the methanol
portance in prolonging the life of the rhodium catalyst for system (vide supra). Alternatively, the alkyl dicarbonyl rho-
acetic acid synthesis. Thus, under conditions when a defi- dium(III) intermediate can dissociate a CO ligand to form a
more stable, but formally coordinatively unsaturated, mono-
ciency of CH3I exists, HI will oxidize the rhodium and initi-
ate a water gas shift cycle. By contrast, under neutral or carbonyl derivative. Facile /3-hydride elimination generates
an intermediate hydrido-olefin complex, which can then re-
basic conditions involving rhodium halide and carbon mon-
oxide, reduction to rhodium cluster carbonyls and free metal
insert to form either the same n-propyl or new isopropyl
occurs (7), and the catalyst would no longer be active. moiety. Further reaction of the isopropyl rhodium(III) spe-
cies gives rise to the isobutyric acid. The retention of isotopic
Carbonylation of Other Primary Alcohols integrity discussed earlier also indicated that the intermedi-
The general reaction scheme used to describe the rhodi- ate hydride was not rapidly exchanged with the solvent.
Since the reaction responsible for isomerization is the 0-
um-catalyzed methanol carbonylation has been extended to
include the carbonylations of benzyl (8), ethyl (4, 9) and n- hydride elimination to form the hydrido-olefin complex, it is
propyl (4) alcohol. likely that the same sort of process occurs during the carbon-
The relative rate data (Table 2) obtained for these alco- ylation of all linear alcohols higher than methanol.
hols suggest very convincingly that the oxidative addition of
the corresponding alkyl halide to the rhodium(I) center is
nucleophilic in nature. The kinetic profiles for these systems Carbonylation of Secondary Alcohols
again indicated that the reactions are first-order in both
rhodium and alkyl iodide and independent of CO pressure. To date, mechanistic studies into the carbonylation of
Labelling studies (4) indicated that there was no significant secondary alcohols with the same type of rhodium/RI cata-
kinetic isotope effect when the substrate (EtOH system) or lyst have used isopropanol as a model substrate. The carbon-
protic solvent (rc-PrOH system) was replaced by deuterium- ylation of isopropanol gives a mixture of n- and isobutyric
substituted species, again consistent with the SN2-type reac- acids. The most recent study (10) shows that the Sn2 path-
tivity. The study with n-propyl alcohol made it necessary to way prevalent with the primary alcohols is not a major con-
expand the original scheme to account for the production of
isomeric products. Over the pressure range examined
(~20-130 atm), the carbonylation of n-propyl alcohol gener-
ated mixtures of n- and isobutyric acids. Increasing the CO
pressure resulted in mixtures that contained decreasing
amounts of isobutyric acid. Again, since the reaction rate
exhibited rate parameters consistent with S^-type reactiv- [RWa(CO)(
r.......
>!" |Rhl3(CO)( >!"
ity, it was concluded that the product selectivity was deter-
mined after this single rate-determining step. This was also
+co
borne out by labelling studies, in which the products, n- and
isobutyric acids, both had the same isotopic composition as
the recovered n-Prl, indicating a similar history. The data [RhMCOM-'-s^r [RhI3(CO)2(—< )]-
are most consistent with the reactions shown in Figure 3. In
this model, nucleophilic oxidative addition is rate-determin-
ing and gives rise to a short-lived alkyl dicarbonyl rhodiu-
m(III) species. By all accounts, cis-dicarbonyl rhodium(III)
species are very unstable. This instability is relieved, at least
in this case, by two possible “decomposition” routes. The
first involves the familiar migratory insertion reaction to
generate an acyl monocarbonyl rhodium(III) center, which

|Rh(CO)2I

|Rh(CO)2L
(trans)

Figure 2. Proposed mechanism for 1he rhodium-catalyzed water gas shift Figure 3. Mechanism for the iodide-promoted, rhodium-catalyzed carbonyla-
reaction. tion of n-propanol to account for the formation of isomeric butyric acids.

Volume 63 Number 3 March 1986 205


tributor with the secondary alcohol. In fact, there appear to Summary
be two significant pathways with the major one being hydro- The rhodium-catalyzed reaction of alcohols with carbon
carboxylation (eq 5). monoxide has led to a new general route to carboxylic acids
with one major commercial application (acetic acid from
methanol) being in widespread use today. One of the major
Ft—CH=CHa + CO + H,0 —>-
limitations of the approach is that the necessity for the
r_CH,—CH2—C02H or R—CH—CFR (6) generation of alkyl halide from primary alcohols or olefin
from secondary alcohols requires that the medium be of
CO.H relatively high acidity.
Literature Cited
The propylene is formed via acid-catalyzed dehydration of (1) Hohenschutz, H.; Von Kutepow, N.; Himmele, W. Hydrocarbon Processing 1966,
the isopropyl alcohol and is present at significant concentra- 45(11), 141.
(2) Roth, J. F.; Craddock, J. H.; Hershman, A.; Paulik, F. E. Chem. Tech. 1971,600-
tioins during the reaction. The active catalyst for the hydro- (3) Forster, D. J. Amer. Chem. Soc. 1976,98,846.
carboxylation reaction is almost certainly a rhodium(III) (4) Dekleva, T, W.; Forster, D. J. Amer. Chem. Soc. 1985, 107, 3565.
(5) Singleton, T. C.; Park, L. J.; Price, J. L.; Forster, D. Prep. Diu. Pet. Chem., Amer.
hydride. The most likely candidate for this species is Chem. Soc. 1979, 24, 329.
[HRhI3(CO)]~ which arises by addition of HI to (6) Baker, E. C.; Hendrickson, D- E.; Eisenberg, R. J. Amer. Chem. Soc. 1980,102,1020
[Rhl2(CO)2]_. Reaction of this hydride species with propyl- (1980).
(7) Hughes, R. P. Comprehensive Organometal. Chem. 1982,5, 277.
ene will lead to the same rhodium(III) species discussed (8) Masuda, A.; Mitani, H.; Oku, K.; Yamazaki, Y. Nippon Kagaku Kaishi 1982,2, 249.
above with respect to the n-propyl alcohol carbonylation (9) Hjortkjaer, J.; Jorgensen, J. C J. Mol. Catalysis 1978, 4,199.
(10) Dekleva, T. W.; Forster, D. J. Amer. Chem. Soc, 1985,107, 3568.
and hence to a mixture of primary and secondary propyl- (11) Hendrickson,J. B.; Cram, D. J.; Hammond, G. S. "Organic Chemistry”; McGraw-Hill,
rhodium species. New York, 1970; p 393.

206 Journal of Chemical Education

You might also like