You are on page 1of 45

National Technical University of Ukraine 1

‘Igor Sikorsky Kiev Polytechnic Institute’

Organometallic Chemistry
(Structure, Reactivity, and Mechanisms)

Carbon Monoxide as a Chemical


Feedstock:
Carbonylation Catalysis
Andrey O. Kushko, Ph.D.,
Senior Lecturer

Kiev-KPI-2023
2
Carbonylation: Chemistry Organic
Synthesis and Technology
3
Introduction to Carbon Monoxide (CO): Preparation
• Burning elemental carbon in restricted supply of oxygen gas
heat
C + O2 CO

• Reduction of carbon dioxide with coke


800oC
CO2 + coke CO
• Dehydration of formic acid (small scale for laboratory)
HCO2H + H2SO4 CO + H2O
• Water-gas shift reaction (preparation of synthesis gas)
Ni
C or CH4 + H2O CO + H2
o
900 C
FexOy
Housecroft and Sharpe.
CO + H2O CO2 + H2 Inorganic Chemistry. Prentice
o
400 C Hall: England 2001.
4
Introduction to Carbon Monoxide (CO): Properties

• Colorless and odorless gas


• Highly flammable and toxic
• Bond length: 112.8 pm
• Bond energy: 257 kcal/mol
• Dipole moment: 0.112 D
• Insoluble in water (26 mg/L)
• HOMO is lone pair on C (σ3)

Housecroft and Sharpe. Inorganic Chemistry. Prentice Hall: England 2001.


5
Introduction to Carbon Monoxide (CO): Properties

CO acts as both a -donor (via the lone pair of


electrons on carbon) and a -acceptor ligand in
transition metal complexes, the metal carbonyls.
The reactivity of CO in a metal carbonyl is
considerably modified: it is for example attacked
by nucleophiles (at the carbon), and undergoes
migration reactions in which metal bonded alkyls
Housecroft and Sharpe. Inorganic Chemistry. Prentice Hall:
transfer to the CO, giving metal-acyls M-CO-R, and
England 2001. similar species.
6
Landmarks of the Use of CO for Carbonylation

Chem 5, 526–552, March 14, 2019 Elsevier Inc..


7

Uses of CO in some Industrial Processes

• Fischer-Tropsch Synthesis;
• Hydroformylation;
• Production of Acetic Acid

Housecroft and Sharpe. Inorganic Chemistry. Prentice Hall: England 2001.


8
Fischer-Tropsch Synthesis

Catalyst
H2 + CO CnHm + H2O

• Discovered in 1922
• Commercialized in 1928
• Heterogeneous catalysis
Fe and Co catalysts
• 200-300 oC and 10-60 bar
• Highly exothermic
• 6.5 Mt / yr by 1944 Prof. Franz Fischer Dr. Hans Tropsch
• Used as synthetic lubricants and
synthetic diesel/jet fuels
Housecroft and Sharpe. Inorganic Chemistry. Prentice Hall: England 2001. 5
9

Proposed Fischer-Tropsch Mechanism

Housecroft and Sharpe. Inorganic Chemistry. Prentice Hall: England 2001. 5


10

Methanol Carbonylation
(Production of Acetic Acid)
11

Industrial Processes for the Production of


Acetic Acid are dominated by Methanol Carbonylation

CH3OH + CO CH3COOH

R.T. Eby, T.C. Singleton, in: B.E. Leach (Ed.), Applied Industrial Catalysts, vol. 1, Academic Press, 1983, p. 275;
N. von Kutepow, W. Himmele, H. Hohenschutz, Chem. Ing. Technol. 37 (1965) 297.
12
The Cobalt-based BASF Process
The first methanol-to-acetic acid carbonylation process was
commercialized in 1960 by BASF

2CoI2 + 2H2O + 10CO Co2(CO)8 + 4HI + 2CO2


Co2(CO)8 + H2O + CO 2HCo(CO)4 + CO2

It used an iodide-promoted cobalt catalyst and required very high pressures (600 atm) as well as high
temperatures (230oC), but gave acetic acid in ca. 90% selectivity. By-products are methane,
acetaldehyde, ethanol and ethers

R.T. Eby, T.C. Singleton, in: B.E. Leach (Ed.), Applied Industrial Catalysts, vol. 1, Academic Press, 1983, p. 275;
N. von Kutepow, W. Himmele, H. Hohenschutz, Chem. Ing. Technol. 37 (1965) 297.
13
Catalytic Cycle of the Cobalt-catalyzed Methanol
Carbonylation (BASF process)

The presence of iodide is necessary in order to convert


methanol into methyl iodide prior to carbonylation. Thus,
D. Forster, M. Singleton, J. Mol. Catal. 17 (1982) 299. the actual substrate of carbonylation is methyl iodide
14
Nickel Catalysts
Nickel carbonyl, as well as a variety of nickel compounds, is also catalytically
active for the carbonylation of methanol in the presence of iodine

NiI2 + H2O + 5CO Ni(CO)4 + 2HI + CO2

Although nickel catalysts have the


advantage of being much cheaper than
rhodium and are easy to stabilize at low
water concentrations, no commercialization
has been achieved to date, since Ni(CO)4 is
a very toxic and volatile compound.

F.E. Paulik, A. Hershman, W.R. Knox, J.F. Roth, Chem. Abstr. 72 (1970)
110807y; J.S. Kanel, S.J. Okrasinski, US Patent to Eastman (1999)
5900504. R.F. Heck, J. Am. Chem. Soc. 85 (1963) 2013.
15
The Rhodium-based Monsanto Process

The production of acetic acid by the Monsanto process is


based on a rhodium catalyst and operates at a pressure of 30-
60 bar and at temperatures of 150-200 oC. The process gives
selectivity of over 99% based on methanol

CH3I, RhI3-nH2O (10-3 M)


CH3OH + CO CH3COOH (99%)
o
180 C, 30 atm.

Iodine sources: I2, HI, CH3I

R.T. Eby, T.C. Singleton, in: B.E. Leach (Ed.), Applied Industrial Catalysts, vol. 1, Academic
Press, 1983, p. 275; N. von Kutepow, W. Himmele, H. Hohenschutz, Chem. Ing. Technol. 37
(1965) 297.
16
Catalytic Cycle of the Rhodium-catalyzed Methanol
Carbonylation (Monsanto process)

D. Forster, J. Am. Chem. Soc. 98 (1976) 846.


B.C. Gates, Catalytic Chemistry, Wiley, New
York, 1992. M. Cheong, R. Schmid, T. Ziegler,
Organometallics 19 (2000) 1973. T. Kinnunen, K.
Laasonen, J. Mol. Struct. (Theochem) 542 (2001)
273. T. Kinnunen, K. Laasonen, J. Mol. Struct.
(Theochem) 540, 2001, 91. T. Kinnunen, K.
Laasonen, J. Organomet. Chem. 628 (2001) 222.
E.A. Ivanova, P. Gisdakis, V.A. Nasluzov, A.I.
Rubailo, N. Rosch, Organometallics, 20 (2001)
1161. T.R. Griffin, D.B. Cook, A. Haynes, J.M.
Pearson, D. Monti, G.E. Morris, J. Am. Chem.
Soc. 118 (1996) 3029.
Catalytic Cycle of the Water-Gas shift reaction as side- 17

reaction in the Rhodium-catalyzed Methanol


Carbonylation
A substantial amount of water (14-15 wt.%) is required to
achieve high catalyst activity and also to maintain good
catalyst stability. In fact, if the water content is less than
14-15 wt.%, the rate-determining step becomes the reductive
elimination of the acetyl species, from catalyst species.
However, as rhodium also catalyzes the water-gas shift
reaction, the side reaction leading to CO2 and H2 is
significantly affected by water and hydrogen iodide
concentration in the reaction liquid
Propionic acid is observed as the major liquid by-
product in this process. This is produced by the
carbonylation of ethanol which is often present as a
minor impurity in the methanol feed

D. Forster, T.W. Dekleva, J. Chem. Edu. 63 (1986) 204.


D.J. Watson, Catalysis of Organic reactions, Marcel Dekker, New York,
1998, p. 369.
18
Catalyst Deactivation in the Monsanto Process
A problem in all catalyzed processes is that the catalyst slowly loses its efficacy – it
deactivates. In the rhodium catalyzed methanol carbonylation this occurs at high acid levels by
oxidation by HI to a Rh(III) tetraiodide anion, [Rh(CO)2I4]-. This complex has a tendency to lose
CO leading, in several steps, to precipitation of RhI3 which appears as an insoluble and
intractable sludge.

[Rh(CO)2I4] [Rh(CO)I4] [RhI4] [RhI3]n

This occurs in parts of the plant where the CO pressure is lower, such as in the flash tank
and recycle loop. Water promotes reduction of the Rh(III) complex into the active Rh(I) form
via the water-gas shift reaction, and helps to minimize the proportion of the catalyst existing
as the problematic [Rh(CO)2I4]- , thus reducing precipitation.

R.T. Eby, T.C. Singleton, in: B.E. Leach (Ed.), Applied Industrial Catalysts, vol. 1, Academic Press, 1983, p. 275;
19
The iridium-based Cativa process
The production of acetic acid using the iridium catalyst system has been
commercialized in 1996 by BP-Amoco as the ‘Cativa TM ’ process

[Ir(CO)2I2] + 2HI [Ir(CO)2I4] + H2

[Ir(CO)2I4] + CO [Ir(CO)3I3] + I
One of the major advantages of the iridium-based process was
the high stability of the iridium catalyst species

Although much more iridium is required to achieve an activity comparable to the rhodium
catalyst-based processes, the catalyst system is able to operate at reduced water levels (less
than 8 wt.% for the Cativa process versus 14-15 wt.% for the conventional Monsanto process)

D. Forster, Adv. Organometal. Chem. 17 (1979) 255. G.J. Sunley, D.J. Watson, Catal. Today 58 (2000) 293; J.H. Jones, Platinum Metal
Rev. 44 (2000) 94. K.E. Clode, D.J. Watson, C.J.E. Vercauteren, European Patent to BP Chemicals (1994) 616997.
20
Catalytic Cycle of the Iridium-catalyzed Methanol
Carbonylation (Cativa process)

D. Forster, J. Chem. Soc.


Dalton Trans. (1979) 1639. P.M.
Maitlis, A. Haynes, G.J. Sunley,
M.J. Howard, J. Chem. Soc.
Dalton Trans. (1996) 2187.
Anionic and Neutral Cycles Proposed by Forster for 21

Iridium-Catalyzed Methanol Carbonylation

Forster, D. J. Chem. Soc., Dalton Trans. 1979, 1639.


22

Ligand Design for Rhodium-based Catalysts

General considerations The migratory insertion reaction of CO into metal-alkyl


bonds is a fundamental step in the metal iodide-catalyzed carbonylation of methanol
to acetic acid. The conditions used industrially (30-60 bar pressure and 150-200oC)
with the original [Rh(CO)2I2]- catalyst have spurred the search for new catalysts,
which could work in milder conditions. The rate-determining step of the rhodium-
based catalytic cycle is the oxidative addition of CH3I, so that catalyst design focused
on the improvement of this reaction. The basic idea was that ligands which increase
the electron density at the metal should promote oxidative addition, and
consequently increase the overall rate of the reaction. For this purpose, other rhodium
complexes have been synthesized in the last years, and they have been shown to be
active catalysts of comparable or better performance as compared to the Monsanto
Catalyst.

J. Rankin, A.D. Poole, A.C. Benyei, D.J. Cole-Hamilton, Chem. Commun. 19 (1997) 1835. J. Rankin, A.C.
Benyei, A.D. Poole, D.J. Cole-Hamilton, J. Chem. Soc. Dalton Trans. (1999) 3771.
23
Phosphine Ligands
The most important class of these rhodium complexes are those containing
simple phosphine ligands such as PEt3

CH3 C(O)CH3
OC PEt3 OC PEt3 OC PEt3
Rh Rh Rh
I PEt3 I PEt3 I PEt3
I I
In 1997, Ranking and coworkers reported a PEt3-rhodium-modified system, and when
employing this basic phosphine, the rate of oxidative addition of MeI to [RhI(CO)(PEt3)2]
is around 47 times higher than that reported for Monsanto catalyst at 25oC.

J. Rankin, A.D. Poole, A.C. Benyei, D.J. Cole-Hamilton, Chem. Commun. 19 (1997) 1835. J. Rankin, A.C.
Benyei, A.D. Poole, D.J. Cole-Hamilton, J. Chem. Soc. Dalton Trans. (1999) 3771.
24

Catalytic Cycle of the Methanol Carbonylation catalyzed by


the Neutral Complex Rh(PEt3)2(CO)I

J. Rankin, A.D. Poole, A.C. Benyei, D.J.


Cole-Hamilton, Chem. Commun. 19
(1997) 1835. J. Rankin, A.C. Benyei, A.D.
Poole, D.J. Cole-Hamilton, J. Chem. Soc.
Dalton Trans. (1999) 3771.
25

Chelating Symmetrical and Unsymmetrical Ligands

Identified intermediates for catalyst


based on dppp

Moloy, K. G.; Petersen, J. L. Organometallics, 1995, 14, 2931.


26
Diphosphine Ligands
In each case the conversion of methanol was greater than 98%, and the selectivity for acetic
acid was greater than 99%; however, the carbonylation rates are lower for these diphosphine
complexes than for the [Rh(CO)2I2]-catalyst

Synthesis of the neutral


rhodium complexes

C. M. Thomas, G. Suss-Fink, Coord. Chem. Rev. 243, 2003, 125-142


27
Trans-Diphosphines in Methanol
Carbonylation – Dinuclear Systems?
These ligands form stable complexes due to the chelate effect which
resemble the ones obtained with PEt3 (trans-[RhI(CO)(PEt3)2]) with the two
phosphorus atoms in a trans disposition
The proposed catalytic cycle is analogous to that for the Monsanto
process, but it involves neutral species with a trans-chelating
diphosphine

S. Burger, B. Therrien, G. Süss-Fink, Helv. Chim. Acta, 2005, 88, 478–486. C.M. Thomas, G. Süss-Fink, Coord.
Chem. Rev., 2003, 243, 125–142. C.M. Thomas, R. Mafia, B. Therrien, E. Rusanov, H. Stœckli-Evans, G. Süss-
Fink, Chem. Eur. J., 2002, 8, 3343–3352.
Comparison of Process Parameters for Methanol 28

Carbonylation Process
BASF (Co) Monsanto (Rh) Cativa (Ir)
1960 1970 1995
Metal concentration, 10-1 10-3 10-4
mole/liter of metal
Temperature, oC 250 150-200 180
Pressure (atm) 500-700 30-60 30-40
Selectivity (%) based on
Methanol
CO 90 99 99,5
70 90 94
Byproducts CH4, CO2, EtOH, MeCHO, CH4, CO2, H2, EtCOOH traces
EtCOOH
29
Synthesis of Acetic Anhydride
In many applications acetic acid is used as the anhydride and the synthesis of the latter is
therefore equally important. In the 1970’s Halcon (now Eastman) and Hoechst (now Celanese)
developed a process for the conversion of methylacetate and carbon monoxide to acetic
anhydride. The process has been on stream since 1983 and with an annual production of several
100,000 tons, together with some 10–20% acetic acid. The reaction is carried out under similar
conditions as the Monsanto process, and also uses methyl iodide as the "activator" for the methyl
group

CH3CO2CH3 + HI CH3CO2H + CH3I

CH3COI + CH3CO2H CH3CO3CCH3 + HI


30
Eastman Carbonylation of Methyl Acetate

C. M. Thomas, G. Suss-Fink, Coord. Chem. Rev.


243, 2003, 125-142. P. Kalck, C. Le Berre, P.
Serp, Coord. Chem. Rev. 402, 2020, 21, 3078.
31

Carbonylation of Higher Alcohols: Higher Carboxylic Acids

Dekleva, T. W.; Forster, D. J. Am. Chem. Soc.. 1985, 107, 3565. Forster, D.; Dekleva, T. W. J. Chem. Educ., 1986, 63,
204.
32

The Pauson–Khand reaction


33
The Pauson–Khand Reaction

The Pauson–Khand reaction (PKR), a [2+2+1] cycloaddition


of an olefin, alkyne, and carbon monoxide, is among the most
common methods to construct cyclopentenones

O
R5 R6
CO (gas) R1 C C
R1 C C R2 + R3 C C R
C R6 Co2(CO)8 catalyst 5
C C R
R4 4
R2 R3

This implies the formation of three new bonds and one or two cycles in the
intermolecular or intramolecular versions, respectively

I. U. Khand, G. R. Knox, P. L. Pauson, W. E. Watts, J. Chem. Soc. D 1971, 36.


34
The Pauson–Khand Reaction

The variants of the Pauson–Khand reaction, e.g., the synthesis of


cyclopentenones employing other transition metal complexes such as Ti, Zr,
Fe, Mo, W, Ni, Rh, Ru, Ir and Pd complexes
I. U. Khand, G. R. Knox, P. L. Pauson, W. E. Watts, J. Chem. Soc. D 1971, 36.
35
Regiochemistry of the Pauson–Khand Reaction

Formation of 2-substituted cyclopentenones is favored when terminal alkynes are utilized

W. J. Kerr, M. McLaughlin, P. L. Pauson and S. M. Robertson, J. Organomet. Chem., 2001, 630, 104.
36
One of the first examples of the CPKR employing a
continuous supply of ethyne (Pauson and co-workers)

I. U. Khand, G. R. Knox, P. L. Pauson, W. E. Watts, J. Chem. Soc. D, 1971, 36. I. U. Khand, G. R. Knox, P. L.
Pauson, W. E. Watts, M. I. Foreman, J. Chem. Soc. Perkin Trans. 1, 1973, 977.
Tuning the Reaction Conditions 37

CPKR under high pressures of carbon monoxide end ethene

V. Rautenstrauch, P. Megard, J. Conesa, W. Kuster, Angew. Chem. 1990, 102, 1441; Angew. Chem. Int. Ed. Engl. 1990, 29, 1413.

Thermal CPKR employing just one atmosphere pressure of carbon monoxide

D. B. Belanger, D. J. R. O. 'Mahony, T. Livinghouse, Tetrahedron Lett. 1998, 39, 7637. B. L. Pagenkopf, D. B. Belanger, D. J. R. O’Mahony, T.
Livinghouse, Synthesis, 2000, 7, 1009.
38
Generation of Cobalt(0) In Situ

Problems associated with the very labile [Co2(CO)8] have led to the
development of catalytic variants of the reaction that generate the active
catalyst in situ from Co(I) or Co(II) precursors

B. Y. Lee, Y. K. Chung, N. Jeong, Y. Lee, S. H. Hwang, J. Am. Chem. Soc., 1994, 116, 8793.
More Stable Sources of 39

Cobalt(0) – Carbonyl Complexes


The cobalt species [Co2(CO)8] is not very easy to handle: it is highly toxic and it ignites
spontaneously uponcontact with air

Examples of more stable sources of cobalt(0)–carbonyl complexes

D. B. Belanger, T. Livinghouse, Tetrahedron Lett. 1998, 39, 7641. M. E. Krafft, C. Hirosawa, L. V. R. Bonaga,
Tetrahedron Lett. 1999, 40, 9177. T. Sugihara, M. Yamaguchi, J. Am. Chem. Soc. 1998, 120, 10782; T. Sugihara, M.
Yamaguchi, M. Nishizawa, Chem. Eur. J., 2001, 7, 1589.
Mechanism of the Pauson–Khand Reaction 40

Energy of the relevant steps of the PKR


M. A. Pericàs, J. Balsells, J. Castro, I. Marchueta, A. Moyano, A. Riera, J.
Vázquez and X. Verdaguer, Pure Appl. Chem., 2002, 74, 167.

P. Magnus and L.-M. Príncipe, Tetrahedron Lett., 1985, 26, 4851.


41
Towards Different Metals
Titanium
PKR catalyzed by titanocene complexes (Buchwald and co-workers)

S. C. Berk, R. B. Grossman, S. L.
Buchwald, J. Am. Chem. Soc. 1993, 115,
4912; S. C. Berk, R. B. Grossman, S. L.
Buchwald, J. Am. Chem. Soc. 1994, 116,
8593. F. A. Hicks, S. C. Berk, S. L.
Buchwald, J. Org. Chem. 1996, 61, 2713.
F. A. Hicks, N. M. Kablaoui, S. L.
Buchwald, J. Am. Chem. Soc. 1996, 118,
9450; F. A. Hicks, N. M. Kablaoui, S. L.
Buchwald, J. Am. Chem. Soc. 1999, 121,
5881.
Towards Different Metals 42

Rhutenium
The [Ru3(CO)12]-catalyzed PKR (Murai and co-workers and Mitsudo
and co-workers)

T. Morimoto, N. Chatani, Y. Fukumoto,


S. Murai, J. Org. Chem. 1997, 62, 3762.
T. Kondo, N. Suzuki, T. Okada, T.-A.
Mitsudo, J. Am. Chem. Soc. 1997, 119,
6187.
Towards Different Metals 43

Rhodium
Rhodium-catalyzed PKR (Narasaka and co-workers). Note that this
example employs an electron-deficient alkyne

[{RhCl(CO)2}2]-catalyzed cyclization of the allene

T. Kobayashi, Y. Koga, K.
Narasaka, J. Organomet. Chem.
2001, 624, 73
44
Substrate Compatibility
1. The PKR of enynes with terminal alkyne groups is more favorable under Co or Rh
catalysis than under Ti catalysis. Enynes with disubstituted alkyne groups are
annelated better by Ti and Rh systems than by Co, Ru, and Ir systems.
2. Enynes with a disubstituted alkyne group and a substituted alkene function react
well with Ti-based catalysts, but give poorer results with Co, Rh, and Ru catalysts
3. Oxygen-containing enynes cyclize well in the presence of Ti-, Rh-, and Ru-based
systems, but not with Co catalysts
4. Nitrogen-containing enynes react well under Rh and Ru catalysis, but not under Ti
and Ir catalysis
5. As a general trend, polar functionalities are tolerated better by late-transition-metal
complexes (cobalt and rhodium) than by titanium complexes.
6. Enynes containing electron-withdrawing groups, either on the alkyne or on the
alkene moiety, have been successfully cyclized only by Rh complexes under a
particular set of conditions.
7. Intermolecular substrates have only really been examined with Co systems; in
these cases, they show greater limitations than intramolecular substrates.
45

You might also like