You are on page 1of 109

Fuel Processing Technology 48 (I 996) 189-297

Reactions of synthesis gas


Irving Wender
Chemicd und Prtrolrum Engineering Depurtment, University oj’Pitt.shurgh. Pittsburgh, PA 15261. USA

Abstract

The use of synthesis gas (syngas) offers the opportunity to furnish a broad range of
environmentally clean fuels and chemicals. There has been steady growth in the traditional uses of
syngas. Almost all hydrogen gas is manufactured from syngas and there has been a tremendous
spurt in the demand for this basic chemical; indeed, the chief use of syngas is in the manufacture
of hydrogen for a growing number of purposes. Methanol not only remains the second largest
consumer of syngas but has shown remarkable growth as part of the methyl ethers used as octane
enhancers in automotive fuels. The Fischer-Tropsch synthesis remains the third largest consumer
of syngas, mostly for transportation fuels but also as a growing feedstock source for the
manufacture of chemicals, including polymers. Future growth in Fischer-Tropsch synthesis may
take place outside the continental United States. The hydroformylation of olefins (the 0x0
reaction), a completely chemical use of syngas, is the fourth largest use of carbon monoxide and
hydrogen mixtures; research and industrial application in this field continue to grow steadily. A
direct application of syngas as fuel (and eventually also for chemicals) that promises to increase is
its use for Integrated Gasification Combined Cycle (IGCC) units for the generation of electricity
(and also chemicals) from coal, petroleum coke or heavy residuals. In the period 2005-2015, the
amount of syngas employed in this manner may approach that used for all other specific purposes.
Syngas is the principal source of carbon monoxide, which is used in an expanding list of so-called
carbonylation reactions.

0. Overview

Synthesis gas (syngas), a mixture of hydrogen and carbon monoxide, can be


manufactured from natural gas, coal, petroleum, biomass and even from organic wastes,
so that sources of syngas are ubiquitous in nature. The availability and flexibility of the
resource base are keys to the present and future uses of syngas and of its separate
components, hydrogen and carbon monoxide. Syngas is a present and increasing source
of environmentally clean fuels and chemicals and is also a potentially major fuel for the

037%3820/96/$15.00 Copyright 0 1996 Elsevier Science B.V. All rights reserved.


PII SO378-3820(96)01048-X
190 1. Wetuler/Fuel Processing Technology 48 (1996) 189-297

production of essentially pollution-free electricity, as sulfur and nitrogen in parts per


million can be removed from syngas relatively easily.
This report will list the commercial uses of syngas in the order of their importance,
compare them with uses of a decade ago and outline what promises to be a bright future
for this mixture of gases.
There are literally thousands of published papers and proceedings of various confer-
ences that deal with syngas reactions. A very large number of these publications deals
with catalytic reactions of syngas in an exploratory way aimed often at the elucidation of
mechanistic pathways. These researches provide an understanding of various aspects of
syngas reactions and point the way to possible future commercial applications and to
new uses of syngas; they will be referred to when they are of particular pertinence.
References to some of the large number of patents that deal with the chemistry and uses
of syngas will be referred to occasionally but are not a major source of material for this
review.
The birth of syngas chemistry occurred in the early part of the 20th century. Methane
was produced by the hydrogenation of carbon monoxide in 1902, followed by the
discovery of the synthesis of ammonia (NH,) in 1910. The Fischer-Tropsch synthesis
was developed in the following decade and then came the manufacture of methanol and
higher alcohols, Syngas is now used in a host of different ways.
The uses and reactions of syngas in the light of changes that have taken place from
1984 to 1994, with an update, constitute the body of this report. The chemistry of the
separate components of syngas, H, and CO, will be discussed in some detail. It is
remarkable that, although these are apparently simple diatomic molecules, they are
readily adsorbed in diverse ways on the surface of catalysts. Together with carbon
dioxide, CO,, they form an unexpectedly large number of complex species with both
heterogeneous and homogeneous catalysts.
The principal fuel uses of syngas in 1994 are given in Fig. 1. The major commercial,
near commercial and potentially commercial chemical uses of syngas in 1994 are
outlined in Fig. 2. Wender and Seshadri (1984) summarized the fuel and chemical uses
of syngas in 1984 (Figs. 3 and 4). Although it appears from a comparison of the

Gasoline
FISCHER-TROPSCH
- Diesel
Chemicals

isobulylene
METHANOL - MTBE
@ME)

Gasoline
Medium BTU GAS
(Turbine fuel. ICCC)

Fig. 1. Principal commercial fuel uses of synthesis gas (1994).


I. Wen&r/Fuel Processing Technology 48 (1996) 189-297 191

DIMETHYL ACETIC
CARBONATE ANHYDRIDE
FORMIC ACID
\ ETHYLENE

\
CHEMICALS
WAXES METHYL ACETATE

T
\ FORMALDEHYDE

CHjOll

FC
AMMONIAC + ACETIC ACID

co .I’
,’
CHLORCO
I IMETHYL 02
METHANE5 i AMINES
Rh ,.’ I
b’ ” I Ethylene
(hydroformylation) ZSM-5
(Zeolites)
Acelaldeh de
Ethsno r Rh I Co I I

i J VINYL
4
ACETATE
ALDEHYDES OldillS
Aromatics
ALCOHOLS
(Z-ethylhexanol)

___ Commercial . -. Near commercial, perhaps availnble for license .. .. .. .. . .. .. Polenlial (next decade)
-

Fig. 2. Commercial, near commercial and potential chemicals from synthesis gas (I 994).

corresponding figures that there have been no significant new uses of syngas in the past
decade, there have been great increases in the “traditional” uses of syngas as new
outlets have appeared.

GASOLINE ~_ZBO!l!?r.F,SCHER_TROPSCH~ GASOLINE


DIESEL OIL / DIESEL OIL
1 ,*’ OTHER PRODUCTS
FUEL
C,-C6 ALCOHOLS MTEIE CELLS

FUEL
METHANE H:, ISOSYNTHESIS GASOLINE (neat)
DIESEL OIL
AROMATICS (BTX)

____ commercial processes


-.-.-.- processes commercial III 1984-85
-~~-~~~~~~-~-~~ processes possibly for llle nsxl decade

Fig. 3. Commercial or near commercial processes for the production of liquid fuels from synthesis gas
(Wender and Seshadri, 1984).
192 I. Weruler/ Fuel Processing Technology 48 (1996) 189-297

FORMIC ACID

t
METHYL ETHYLENE ACETIC
AMMONIA FORMATE GLYCOL ANHYDRIDE

\H2 \ 1 cD~_/~~;~A~~~E

i \
ETHYLENE ETHANOL SINGLE-CELL
GLYCOL PROTEIN

IXYLENES 1

commercial
-.-.-.-- near commcrciol
._-......_.....-. polenlinl (ncnl decade)

Fig. 4. Commercial, near commercial and potential chemicals from synthesis gas (Wender and Seshadri,
1984).

Although hydrogen as such is not found in the list of top chemicals produced in the
United States, the principal use of syngas is for the manufacture of hydrogen, huge
amounts of which are consumed in the synthesis of ammonia. The demand for hydrogen
continues to grow. Environmental concerns constitute a driving force for greater growth
in the use of hydrogen to produce clean fuels and chemicals.
In addition, there has been a spectacular increase in the manufacture of methanol for
the synthesis of methyl r-butyl ether (MTBE), which is used in reformulated automobile
fuels. Indeed, methanol used for fuel and chemicals has achieved the status of a
commodity chemical. The Fischer-Tropsch synthesis remains the third largest consumer
of syngas. Sasol, using iron catalysts, continues to expand in the synthesis of fuels and
chemicals from coal based syngas. In 1993, the Shell Middle Distillate Synthesis came
onstream, utilizing natural gas to produce mainly gasoline and diesel fuel in fixed
tubular reactors. Plans for further growth in these newer syngas conversions are
contemplated. There is great interest in the utilization of remote natural gas reserves.
Exxon has built and operated a large Fischer-Tropsch (FI’) demonstration plant using
a multiphase slurry reactor. Both the Shell and the Exxon processes are based on the use
of cobalt catalysts. Both produce high molecular weight paraffins (waxes) which are
hydroisomerized and cracked to liquid products suitable for refinery or chemical plant
feedstocks.
Parallel with the development of the Exxon multiphase process, Sasol has developed
the so-called Sasol SPD (slurry-phase distillate) process. Until now it is the only
commercial experience that has demonstrated successful IT slurry operation. It is
directed towards the production of paraffinic waxes; however, the product also consti-
tutes a valuable feedstock for high-quality diesel fuel.
I. Wender/ Fuel Processing Technology 48 (19961 189-297 193

The hydroformylation of olefins to aldehydes and alcohols, which is completely a


chemical use of syngas, remains the fourth largest use of carbon monoxide and hydrogen
mixtures (it was inadvertently omitted from Fig. 4). In this versatile reaction, termed the
hydroformylation or 0x0 synthesis, a large variety of olefins react with syngas, in the
presence of cobalt or rhodium catalysts, to form aldehydes and alcohols with one carbon
more than the starting olefin. The wide applicability of this synthesis has resulted in its
steady growth in the synthesis of plasticizers, pharmaceuticals and hundreds of other
chemicals. As this is a chemical use, with chemicals sold by the pound, the amount of
syngas is considerably below that utilized in the production of hydrogen, methanol or
Fischer-Tropsch products.
A direct use of syngas as a fuel that has potential for rapid growth is as a clean fuel
for integrated gasification combined cycle (IGCC) units for the generation of electricity.
The amount of syngas employed in this manner may exceed that used in Fischer-Tropsch
reactions in a few years. Methanol may play a part as a peaking fuel in IGCC. If excess
syngas is available from IGCC units, its conversion to methanol may be the best road to
take, for methanol is a convenient storage fuel for peak use. In addition, methanol
synthesis plants can be converted to the production of ammonia.
A 260 ton per day liquid-phase methanol facility is currently being built in Kingsport,
Tennessee. It is the result of a joint venture of the Eastman Chemical Company and Air
Products and Chemicals. The process is based on the use of a multiphase slurry reactor
for the production of methanol from synthesis gas. Copper based catalysts are being
used. The process was developed by Air Products with the sponsorship of the US
Department of Energy.
It is important to remember that the preparation and clean-up of syngas from any
source approaches 60-70% of the cost of the final product. There is continuing steady
progress by industrial firms in lowering costs involved in syngas preparation.
Mobil’s methanol-to-gasoline (MTG) process is in commercial operation in New
Zealand, producing both methanol and high octane gasoline. The plant has been sold to
a private company which is a large producer of methanol. The plant has met the strategic
goal of reducing New Zealand’s dependence on foreign oil, but methanol could be sold
as such rather than being converted to high octane fuel. Mobil has viable processes for
the conversion of methanol to gasoline in a slurry reactor and also processes for the
conversion of methanol to diesel fuel. Mobil’s ZSM-5 technology has been optimized
and the future of these processes would brighten should oil prices rise.
A comparison of the data in Fig. 1 with those in Fig. 3 reveals that the commercial
syngas processes appear to be the same (solid arrows) but that essentially all of the
potential syngas uses (broken arrows) have not come to fruition, except for the
conversion of methanol to gasoline via the Mobil MTG process. Although operated in
Italy, no plant now exists for the synthesis of higher alcohols directly from syngas;
methyl r-butyl and other ethers have taken over. The Isosynthesis deserves further
research but conditions are too severe and selectivity is low. Neat methanol use as an
automobile fuel is not promising; it picks up water easily and is corrosive to certain
automobile parts, among other problems. Ethanol, made by fermentation, is used as an
octane extender and as part of ethyl r-butyl ether in reformulated gasoline.
The present syntheses of major chemicals from syngas are shown in Fig. 2, which
194 1. Wender/ Fuel Processing Technology 48 (1996) 189-297

should be compared with the diagram shown in Fig. 4. Not too much has changed in
terms of outlets for syngas. Ethylene glycol is still synthesized commercially via
ethylene oxide; other indirect routes to this glycol are under study. The direct conversion
of syngas to ethanol has been studied by many; a two-step synthesis of ethanol via
methanol and syngas (the homologation of methanol) is more promising and may
become commercial in the next decade. The conversion of methanol to olefins and
benzene, toluene and zylene (BTX) via Mobil technology may become of industrial
interest in a decade.
As a chemical, ammonia still consumes the largest amount of syngas via its
conversion to hydrogen. All new plants for the synthesis of acetic acid (derived solely
from syngas) will use the rhodium-catalyzed methanol plus carbon monoxide process,
which shows steady growth in terms of syngas use. The homogeneous carbonylations of
methanol to acetic acid and of methyl acetate to acetic anhydride are, at present, the only
examples of syngas use which have displaced petroleum based routes to commodity
chemicals. The Eastman Chemical coal based plant for converting methanol to acetic
anhydride via acetic acid is a great success and has undergone recent expansion. There is
much promise for growth in the production of other oxygenated chemicals from syngas
as this plant grows and diversifies.
Formaldehyde, long the chief chemical derived from methanol, has been overtaken by
the use of methanol for the synthesis of methyl ethers such as MTBE.
Mechanistic aspects of most syngas reactions are discussed, but it is not the purpose
of the present report to go deeply into this aspect. It is recognized that there remains
considerable need for further research in almost every area for the elucidation of
reaction intermediates and pathways.

INDLRJXTSYNTHESIS
SYNTHE8IS (VIA METHANOLS

Hydrogen Formaldehyde Olefins H&O Aldehydes


Methanol Acetic Acid Co/Rh - Alcohols
Ammonia Methyl Acetate
Carbon Monoxide Acetic Anhydride Isobutyle-ne Q&Q& MTBE
Medium BTU Gas Vinyl Acetate H+
Methane Methyl Formate
Higher (C,-Cd Alcohols Formic Acid Acetylene n Aaylic Acid
Gasoline Ethanol Ni
Diesel Fuel Dimethyl Carbonate
Isobutanol Dimethyl Oxalate Olefms &_ Highly-branched Acids
Gasoline co -
Diesel Fuel
Ethylene RCOOH +2Q RCH.#OOH
Propylene RlQ&
m-x
Chloromethanes Nimxuomatics a Isocyanates
Methylamines Pd
Methyl Glycolate
Ethylene Glycol Teqhtbalic Acid CH,OH Dimethyl Taephthalate

Fig. 5. Fuels and chemicals from synthesis gas.


I. Weruler/Fuel Processing Technology 48 (1996) 189-297 195

This report classifies syngas reactions into six categories which, although they may
occasionally overlap, are relevant enough to stand apart. The choice of categories rests
to a great extent on the amount of syngas used and the commercial importance of the
products. The categories are listed below.
1. Synthesis gas as a source of hydrogen.
2. Direct conversion of syngas to fuels and chemicals.
3. The hydroformylation (0x0) reaction.
4. The Mobil methanol-to-gasoline (MTG) and related processes.
5. Methanol plus syngas or carbon monoxide for the synthesis of chemicals.
6. Miscellaneous reactions.
A summary of fuels and chemicals obtained from syngas reactions is given in Fig. 5.

1. Introduction

I .I. Background

The main body of this review will not deal in detail with the manufacture of syngas
and of the separation of hydrogen and of carbon monoxide from syngas. Various sources
and processes for syngas production yield different ratios of H, to CO and the
availability of source material for syngas manufacture varies at different places in the
world. Matching the specific method of production of syngas with the use in a particular
synthesis is an important factor in determining the technical and economic value of the
various processes (Haag et al., 1987).
Natural gas is the largest source of syngas at present and its use for this purpose is
growing. Methane is the chief constituent by far of natural gas. Petroleum fractions are
the next largest syngas source at present and significant quantities of syngas are being
made from coal.
Thermodynamically, methane is an exceptionally stable molecule, but it can be
activated by reaction with such reagents as oxygen or water. A great deal of research has
been carried out on the reaction of methane with oxygen but total oxidation to CO, is
usually the major product. The equilibrium constant for the reaction of methane with
water, the well-known steam reforming reaction, is more favorable, with an equilibrium
constant of unity at 627°C.
The conversion of methane to syngas is carried out commercially on a very large
scale in the presence of a nickel catalyst at about 700°C under pressure. The CO,
concentration is reduced by the presence of H, in the reaction product.
As Notari (1991) has pointed out, “ALL industrial production for which methane is
the raw material proceeds through the intermediate production of syngas: it is the ONLY
reaction by which methane can be transformed into reaction intermediates (H, and CO)
with limited amounts of undesired CO,.”
The route to methanol, which is made commercially in large plants throughout the
world, is through syngas in essentially all cases, with natural gas as the chief source of
the syngas. Since natural gas has been in short supply in Europe, petroleum fractions
have been used as a source of syngas. The near-panic programs pursued following the
1% I. Wader/ Fuel Processing Technology 48 (1996) 189-297

1973-1974 oil embargo involved coal as the source of fuels and chemicals, including
increased gasification of coal to syngas. Coal utilization in the United States and
worldwide increased and this increase has been sustained, but coal research and
development have decreased with the decline in oil prices and the increased availability
of natural gas. The long-term instability of oil and gas markets is a reason for sustaining
research, development and technology for the use of syngas (often referred to as C-l
chemistry).
There are a few salient points, mostly well known, that apply to the supply of fuels
and chemicals in general and to syngas in particular: (i) perhaps the most precarious
situation involves the availability of petroleum, as the largest oil resources exist in
politically unstable areas far from the United States. In terms of availability, natural gas
and coal constitute large resources and are ubiquitous. (ii> Petroleum, natural gas and
coal will continue to compete for world markets for several decades. Flexibility will be a
key so that a particular application will not depend on the availability of one kind of
fuel. As all of these resources are convertible to syngas, applications based on syngas
use will continue to grow. (iii) It is not possible to separate the fuels and chemicals
businesses as some are prone to do. The use of fuels will govern the use of chemicals, as
is the case in the petroleum industry. Fuels are sold by the barrel or by the ton;
chemicals are of higher intrinsic value and are sold by the pound. (iv) With regard to
coal use, gasification processes are more forgiving than direct liquefaction. Variability in
coal source, rank, mineral matter, etc. are more easily accommodated by gasification
processes. Indirect liquefaction (via syngas) products are relatively easily converted to
clean gaseous and liquid fuels and chemicals. (v) The major cost of almost all uses of
syngas lies in the gasification and purification of the syngas. (vi) Syngas will continue to
be the source of methanol, although there is significant ongoing research on the direct
oxidation of methane to methanol and products derived from methanol, such as
formaldehyde. (vii) Syngas will continue to be the world’s principal source of hydrogen.
Hydrogen has a great many uses, including primary application for the manufacture of
ammonia (NH,). (viii) Environmental and economic factors are the eventual determi-
nants in fossil fuel use. Even a large amount of an available resource, if environmentally
harmful, will tend to lose out sooner or later. Environmental constraints are found in gas
clean-up and in the nature of gasification residues. The latter should be characterized, be
non-leachable and, if possible, be useful in some way. (ix) It is almost inevitable that
future power plants will allow the use of any clean fluid fuel: natural gas, oil, syngas,
petroleum, methanol, or even a Fischer-Tropsch derived liquid. There is a growing
tendency for coal and petroleum coke to be converted to syngas and then used in a type
of integrated gasification combined cycle (IGCC) mode, as first demonstrated at Cool
Water, California. This procedure could result in the proliferation (often nearby) of
plants that convert the syngas to high-quality transportation fuels or to chemicals, or use
the gas as a clean industrial fuel. (XI The modem petrochemical industry is based on
natural gas and the by-products of petroleum refining. The so-called petrochemical
feedstocks are comprised of ethylene, propylene, the butenes, benzene, toluene, and the
xylenes. All of these and liquid fuels can be made from syngas, and it is therefore
possible to envision a petroleum-less refinery. (xi) As tetraethyllead disappears from
automobile gasoline tanks, the demand will grow for oxygenated chemicals with high
I. Wender/Fuel Processing Technology 48 (1996) 189-297 197

octane ratings. Economical syntheses of such fuel-blending agents as MTBE, other


ethers or higher alcohols are needed. (xii) Sulfur and nitrogen compounds are removed
from crude syngas and can be sold as solid sulfur or as ammonia, respectively. (xiii) A
major factor is the discovery of large natural gas supplies. The use of natural gas as a
source of syngas has the advantage of producing less carbon dioxide, a greenhouse gas,
than syngas from other sources. It is considerably cheaper to obtain syngas from natural
gas than from naphtha or coal. (xiv) Fossil fuels will remain the dominant source for the
production of energy; technologies that are economically feasible and minimize pollu-
tion must be developed.
There are many issues associated with the integration of gasifiers with syngas
conversion processes (Haag et al., 1987). The downstream conversion process dictates a
minimum partial pressure of the active compounds (HZ and CO) in the syngas. This
minimum partial pressure is affected by three factors in the gasifier: gasifier pressure,
purity of the oxygen used, and the amount of light paraffins (mainly methane) generated
in the gasifiers. In addition, the syngas pressure is changed by using either a gas
compressor or an expander.
The HZ/CO ratio of the syngas is another important issue. It is costly to decrease the
HZ/CO ratio because of thermodynamic constraints. However, the use of a water gas
shift (WGS) unit to increase the HJCO ratio is a proven technology that adds
substantially to the process cost. Furthermore, syngas conversion processes use catalysts
that usually require syngas with less than a defined minimum threshold level of
impurities (usually H,S and COS). Expensive syngas purification units are needed to
remove excess impurities.
Another important integration issue is the co-production of energies in various forms.
There are actually two sub-issues involved here. One is the co-production of fuel gas
and syngas with liquid fuels to achieve high thermal efficiency. In a coal based synfuel
plant, it is almost always more costly to have a process designed to produce only liquid
fuels. By recycling the light hydrocarbons to make additional syngas, 3040% of the
energy in the hydrocarbons is easily lost. The other sub-issue is the co-production of
steam and electricity, as in cogeneration. With proper integration and optimization of a
gasifier with a syngas conversion process, the resulting high thermal efficiency can
result in a surplus of steam or electricity. The steam may be exported to nearby factories
as process steam or to nearby communities as utility steam, and the electricity is
exported to the local utility network.

1.2. H, + CO partial pressure requirements

The degree of importance attached to high H, + CO partial pressures in the syngas


depends strongly on the syngas conversion process. Molar contraction almost always
occurs with syngas conversion; consequently, conversions are thermodynamically more
favorable at higher H, + CO partial pressures. However, the importance of this require-
ment varies with the conversion process. Generally speaking, conversion reactions that
are limited by chemical equilibrium, such as the methanol synthesis, will require higher
partial pressures. Fischer-Tropsch (IT) reactions are less strongly constrained by
equilibrium considerations and can tolerate low syngas partial pressures.
198 1. Weruler / Fuel Processing Technology 48 (1996) 189-297

In commercial application, the synthesis of methanol requires a minimum syngas


pressure of 600-700 psig. Although FT reactions may proceed to high syngas conver-
sions at low pressures, moderate to high partial pressures (20-1000 psig) are usually
used to improve process economics.
The easiest way to obtain the required syngas partial pressure is to select gasifiers
that operate at the proper pressure. If a low-pressure gasifier is employed, the syngas is
compressed to the required value. It is more efficient to compress gasifier oxygen and
use high-pressure steam for gasification than to compress output syngas. The penalty for
gas compression is relatively smaller at high pressures (e.g., above 300 psig) because the
compression ratio is reduced. The best operating gasifier pressure ranges from = 300 to
600 psig.
Most impurities in syngas gasifiers are removed fairly easily. Removal of sulfur
compounds is essential because they poison the catalysts used in the downstream
processes. Established purification processes for removal of sulfur compounds have been
given by Kuo (1984). The maximum allowable sulfur levels for the synthesis of
methanol and for the Fischer-Tropsch reaction are about l-2 ppm by volume.
As mentioned earlier, different ratios of Hz/CO are obtained from various feed-
stocks.
From natural gas
CH,+H,Oe3H,+COAH= +49.0kcalmol-’ (1.1)
From oil fractions
-CH, - +H,O ZZ2H, + CO AH= +36.1 kcalmol-’ (1.2)
From coal
C+H,Ot3H,+COAH= +3l.lkcalmol-’ ( 1.3)
The energy required for these processes can be supplied by external heating or by
partial combustion of the feedstock.
At present, a great deal of research and development is aimed at reducing the cost of
producing and purifying syngas, as these operations constitute the major cost in the
conversion of syngas to fuels and chemicals (Rostrop-Nielsen, 1994, Sunset et al.,
1994).
Syngas and its separate components are involved in a great number of reactions and
chemical syntheses. It will be of help to memory, and to devising new chemical
syntheses, to separate syngas conversions into presently commercial, potentially com-
mercial, and processes which have been investigated but, at present, do not appear to be
able to replace existing processes not based on syngas. Syngas conversion chemistry has
undergone improvement over the last decade but, as will be shown, many attempts to
convert syngas directly to important uses, especially to chemicals, have not been
successful as yet.

1.3. Carbon monoxide: structure and reactivity

Before discussing syngas reactions, a short consideration of the structure and


reactivity of the carbon monoxide molecule is in order. The CO molecule contains a
I. Weruler/ Fuel Processing Technology 48 (19961 189-297 199

total of 14 electrons, six contributed by carbon and eight by oxygen. Its structure may be
represented as :C:::O:, in which only the valence electrons are depicted. Oxygen
transfers an electron to carbon producing a structure with three bonds; extra stability is
provided by coulombic interaction of the oppositely charged atoms. One of the principal
arguments in favor of a triple bond in carbon monoxide is its analogy with the nitrogen
molecule which also has 14 electrons and unequivocally can be assigned a triple bond
structure. Indeed there is a striking similarity between the physical properties of CO and
N, (Orchin and Wender, 1957). The bond dissociation energy of N, is 225.1 kcal, and
that of CO is 255.7 kcal. This is consistent with the triple bond formulation of CO. The
bond energy of CO is the largest observed bond energy of any diatomic molecule; the
CO molecule does not dissociate in the synthesis of methanol or in many carbonylation
reactions.
It is therefore remarkable that CO does dissociate readily over certain transition
metals. This is a necessary condition that must take place on a Fischer-Tropsch catalyst.
There are important variables which are necessary for this to take place; occupation of
the CO 2n (antibonding) orbital of CO and bonding of the dissociated carbon and
oxygen atoms to transition metal catalysts. Dissociation of CO also occurs in the
methanation of syngas, in the Boudouard reaction and in the conversion of syngas to
higher alcohols over heterogenous catalysts.
Bonding in metal-CO complexes and adsorption on transition metals involve both o
and rr components. Transition elements readily form strong bonds with compounds
(CO, ethylene) that contain n-electron systems or which have orbitals of suitable energy
and symmetry to form dn-bonds. Bonding between transition metal atoms and CO is of
great importance because transition metals, deposited on supports or as discrete com-
plexes, are catalysts for the reaction between CO and most other molecules. Multiple
bond character in the metal-carbon bond occurs by formation of metal-CO rr bonds by
overlap of metal dr orbitals with empty antibonding CO orbitals (Fig. 6).
Hydrogen plays at least two important roles in syngas reactions (Bartholomew, 1988).
It acts as a reductant to activate the catalyst; this affects the initial chemical catalyst
states in ensuing reactions. Secondly, as H, is a reactant in the hydrogenation of CO, its
adsorption kinetics, energetics and interactions with CO/carbon surface species greatly
affect the activity and selectivity behavior of the catalysts used.
Carbon monoxide reacts as a weak base (an electron donor) in many acid-catalyzed

filled
d orbital
metal V
empty CO
v* orbital

Fig. 6. Molecular orbital representation of a carbon monoxide transition metal bond.


200 I. Weder / Fuel Processing Technology 48 (1996) 189-297

reactions, e.g., with carbenium ions: R+ +:C=O --) (RCO)+. It reacts less readily as an
electron acceptor, as in the synthesis of formates; here CO acts as an electron acceptor,
with an empty orbital on the carbon atom: B- + C=O + (BCOl-.
There are four commercial processes for the purification of CO. Two are based on
CO absorption by salt solutions, one uses low temperature condensation or fractionation
and the fourth involves adsorption of CO on a solid adsorbent material (Pierantozzi,
1993). Pressure swing adsorption on high-area materials is being adopted increasingly in
preference to cryogenic separation. Similar techniques to remove impurities are used in
all four processes.

2. Classification of syngas reactions

2.1. Major commercial uses of syngas

This report will classify syngas reactions into categories which, although they may
occasionally overlap or seem to intrude on one another, are relevant enough to stand
apart. The choice of the categories rests to a great extent on the amount of syngas used
and the commercial importance of the products. Another factor is whether the syngas is
converted directly to fuels and chemicals (direct synthesis to products) or whether the
product of the direct synthesis reacts further with either syngas or carbon monoxide (i.e.
syngas is first converted to methanol which then reacts with carbon monoxide to form
acetic acid).
The largest commercial use of syngas is in the manufacture of hydrogen gas, more
than half of which is used in the synthesis of ammonia. The second largest commercial
use of syngas is in the synthesis of methanol. The third largest industrial use of syngas is
in its conversion to paraffins, olefins and oxygenates via the Fischer-Tropsch reaction.
There is recent renewed interest in this synthesis with the start of operation of new
plants. These three uses account for the main industrial commercial utilization of syngas.
The fourth largest industrial use of syngas, the hydroformylation (0x01 reaction,
although considerably smaller than the uses given above, is a viable, flexible reaction
that is used in a large number of syntheses of important industrial products. Hydro-
formylation consists of a general and widely applied conversion of olefins plus syngas to
aldehydes and alcohols. It is often used to synthesize intermediates in a sequence of
synthetic reactions. Hydroformylation plants are found in many countries throughout the
world.
The use of syngas in the generation of electricity via integrated gasification combined
cycle (IGCC) from coal and petroleum coke has the potential for considerable growth.

2.2. Categories of syngas reactions

Category 1: syngas as a source of hydrogen.


Category 2: direct conversion of syngas to fuels and chemicals: includes the synthesis
of methanol and the Fischer-Tropsch reaction.
I. Wmder/Fuel Processing Technology 48 (1996) 189-297 201

Category 3: the hydroformylation (0x0) reaction: olefins plus syngas to aldehydes and
alcohols.
Category 4: the Mobil methanol-to-gasoline (MTG) and related processes.
Category 5: methanol plus syngas or carbon monoxide, for the synthesis of chemi-
cals.
Category 6: miscellaneous carbonylation reactions.

3. Category 1. Manufacture of hydrogen

Syngas is our chief source of hydrogen gas. When syngas is treated with water in the
presence of generally nickel based catalysts, the CO reacts via the water-gas shift
(WGS) reaction to yield a second mole of hydrogen, if required. The WGS reaction is
also used to obtain hydrogen to carbon monoxide ratios desired for particular syntheses:
about 2H,: ICO for the synthesis of methanol and for the Fischer-Tropsch reaction and
about 3H,: 1CO for the synthesis of methane.
Approximately ten trillion (1012> standard cubic feet (SCF) of hydrogen are con-
sumed annually in the world; about four are used per year in the United States.
Worldwide, most hydrogen is used for the production of ammonia (almost six trillion
SCF of syngas per year). Refinery needs for H, have been expanding rapidly, meaning
significant growth in hydrogen demand. New environmental regulations are driving
more demand for hydrogen for better gasoline quality and improved diesel. Another very
large growth in demand for hydrogen will be for process needs and syngas (for
methanol) to satisfy the enormous growth in MTBE use. Detailed uses of hydrogen as a
basic material for large-scale processes are: ammonia synthesis, manufacture of methanol,
hydroformylation of olefins to aldehydes and alcohols (the 0x0 reaction), hydrogenation
to form organic intermediates (amines, cyclohexane, aliphatic alcohols and natural fats
and oils), iron ore reduction, protection gas, petroleum processing (hydrocracking,
hydrodealkylation, hydrodenitrogenation, hydrodesulphurization, and hydrorefining).
Fischer-Tropsch synthesis, methanation, coal hydrogenation, heavy oil hydrogenation,
in rocket propulsion, etc. Hydrogen is also expected to play a major role as an energy
carrier. Estimates of future hydrogen demand in the non-energetic area are outlined by
the assumed growth rates of various products. Specifically, hydrogen demand for
synthesis processes is likely to grow because of the already existing high level of
technological development. Environmental regulations are in part responsible for in-
creased hydrogen demand.
Hydrogen used in the chemical and petrochemical industries is produced primarily
from fossil energy carriers (natural gas, crude oil, and coal) and steam. Water electroly-
sis to produce hydrogen is restricted to areas with low hydrogen demand, since electric
power costs are almost prohibitive. However, significant amounts of hydrogen from
water electrolysis for ammonia synthesis are produced in a few countries with sufficient
electric power from hydroelectric plants. The principal commercial processes for
hydrogen manufacture from fossil energy carriers are catalytic steam reforming, pressure
partial oxidation of hydrocarbons, and coal gasification. In the United States, most
industrial hydrogen is manufactured by steam reforming of natural gas. Relatively small
202 I. Wemler/ Fuel Processing Technology 48 (1996) 189-297

amounts of hydrogen are produced by steam reforming of naphtha and by partial


oxidation of oil. Processing difficulties and manufacturing costs increase as the feed is
changed from natural gas to liquid hydrocarbons and then to coal; partial oxidation and
coal gasification processes require higher capital costs. Because of the newly created
hydrogen demand for the gasoline and diesel pool, improved reformer designs are
underway to reduce the cost of hydrogen production. A cheap way of producing
hydrogen from water is a most desirable goal.
The product gas leaving the hydrogen manufacture unit from fossil energy carriers is
a mixture of hydrogen, carbon monoxide, carbon dioxide and small amounts of methane.
Purification steps of hydrogen include WGS catalysis, CO, removal, methanation,
pressure swing adsorption, and cryogenic separation. The CO is converted to CO, and
hydrogen via the WGS reaction (CO + H,O + CO, + H,) which has the advantage of
increasing the hydrogen yield at the expense of carbon monoxide. Carbon dioxide is
scrubbed out by ethanolamine solutions or other processes such as Rectisol, Selexol,
Sulfinol, amine guard or hot potassium carbonate scrubbing. Remaining small amounts
of carbon oxides are converted to methane over supported nickel methanation catalysts.
Pressure swing adsorption (PSA) is a single step process in which all the constituents,
except hydrogen, are simultaneously removed as the reformer effluent moves through
the adsorbent bed. In a cryogenic purification system, the feed is cooled by indirect heat
exchange and impurities are condensed and removed as liquid by-products.
The chief commercial uses of syngas are the following in decreasing order: as a
source of molecular hydrogen > in the synthesis of methanol > Fischer-Tropsch synthe-
sis > in the hydroformylation of olefins to aldehydes and alcohols > in carbonylation
reactions in which carbon monoxide reacts with other molecular species. There is a large
potential for a rapidly growing use of syngas as a fuel for electricity generation via
integrated gasification combined cycle (IGCC).

4. Category 2. Direct conversion of syngas to fuels and chemicals

4.1. Background and scope

There are five major direct uses of syngas which do not involve a third chemical
reactant, i.e. syngas is used as such. They are:
- Synthesis of methanol-commercial
- Fischer-Tropsch and related reactions-commercial
. The Isosynthesis-not commercial
* Direct combustion for generation of heat and/or electricity-commercial
* Methanation of syngas to synthetic natural gas (SNG)-one plant in the US.
The synthesis of methanol and the Fischer-Tropsch synthesis (FTS) are the two
major direct uses of syngas for the production of fuels and chemicals. Each will be
discussed separately, but it is of considerable interest to point out similarities and
differences of these two important syntheses. The original patents for both syntheses
were obtained by BASF in Germany in the same year, 1913, (Anderson, 1984; Lee,
1990). The first commercial plant for the conversion of syngas to methanol was built in
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 203

1923 by BASF. The first commercial plant for the FTS began operation in Germany in
1935.
Both syntheses are conducted essentially with a syngas ratio of 2H,: ICO.
The methanol synthesis is equilibrium limited; the ITS is limited by Anderson-
Schulz-Flory selectivity and kinetics.
The methanol synthesis, as practiced commercially, yields methanol in over 99%
selectivity; the FIS yields a multitude of products (and incidentally, very little
methanol)
Transition metal catalysts are used in both syntheses. Methanol synthesis catalysts,
found in the right side of the periodic table, do not split the carbon-oxygen bond in
carbon monoxide. In a basic difference, FT catalysts, found in the left side of the
periodic table, proceed mainly by actual breaking (dissociation) of the carbon-oxygen
bond in carbon monoxide.
The weight retention of syngas (2H, + lCO> as a feedstock for methanol approaches
100%. In the FT reaction with iron catalysts, each methylene (-CH,-) group in the
formed carbon chain is accompanied by formation of water or carbon dioxide
depending on the catalyst promoter level and operating temperature. With a cobalt
catalyst in the FT process, water rather than carbon dioxide is formed.
Methanol is the source material for a large number of chemicals, mostly oxygenated
compounds. Many of the chemicals involve reaction of methanol with one or more
other compounds.
A large variety of chemicals is also produced in the FI’S but complex and sophisti-
cated separations and upgrading schemes are necessary to obtain these chemicals.
Methanol, when first produced from syngas, was used almost entirely as a chemical
and chemical precursor. In the last decade and in the upcoming years, the fuel uses of
methanol, largely to form methyl ethers, will overtake its use for chemical produc-
tion.
The FT synthesis began as a process to produce liquid fuels. In the last decade,
however, the production of chemicals and of widely used waxes via the FT reaction
has grown in importance.
Methanol use has expanded rapidly as a component (35.3 percent) of methyl
tert-butyl ether (MTBE), and other ethers which are oxygenated octane enhancers and
fuels.
The FIS has new potential in the synthesis of high molecular weight products which
are then cracked to yield high-grade diesel fuel in addition to gasoline.

4.2. Methanol manufacture

4.2. I. Introduction
The world consumption of methanol is over 24 million tons per year and is still
growing (Gee& et al., 1990). Methanol is one of the top ten organic chemicals
manufactured in the world. The alcohol is a fuel itself, an octane extender, it is used in
the manufacture of other fuels such as methyl t-butyl ether and higher alcohols, and it is
a precursor to a large and varied number of high-value chemicals and polymers. The
204 I. Weder / Fuel Processing Technology 48 (1996) 189-297

many uses of methanol will be discussed later. Almost all methanol is synthesized from
syngas containing 2-8 ~01% of CO, (H,/CO/CO,) derived to a very large extent by
steam reforming of natural gas. Small amounts of syngas are obtained by steam
reforming of naphtha and even smaller amounts by gasification of coal. The ratios of
H&O/CO, depend on the source.
Methanol is catalytically synthesized by the reactions

CO+2H,+CH,OH, AH&,= -24.0kcalmoll’


(4.1)
AG&. = + 10.8 kcal mol- ’

CO, + 3H, --) CH,OH + H,O, AH;,,., = - 14.7 kcalmol-’


(4.2)
AG&,oC = + 14.8kcalmoll’

The water gas shift (WGS) reaction occurs simultaneously with methanol synthesis.

CO + H,O + CO, + H, AH~,,., = -9.3 kcalmol-’


(4.3)
AG”,,,., = -3.9kcalmol-’

The synthesis of methanol from syngas is one of the technically very well-developed
industrial processes. Although a great deal of information has been published on its
chemistry, the mechanism of the commercial synthesis of methanol from syngas has
been the subject of much controversy.
In France, Patart (1921) obtained a patent covering the hydrogenation of CO to
methanol and other oxygenated compounds at elevated temperatures and pressures with
catalysts containing Cr, Co and Mn in the metallic form in addition to various oxides
and other compounds.
In 1923, BASF obtained a patent for the synthesis of methanol and built the first
commercial plant for the conversion of syngas to methanol using a zinc oxide/chro-
mium catalyst (BASF, 1923; Lormand, 1925). As recently as 1923, methanol was
produced by the distillation of wood. How many trees have been saved by the
development of catalysts for the methanol synthesis-ecological benefits through cataly-
sis (Waugh, 1992)!
At this stage, realizing that methanol could be manufactured more economically
catalytically, DuPont and the Commercial Solvents Corporation started experimenting
with the catalytic synthesis of methanol and, in 1927, commercial production by the
high-pressure process began in the United States (Wade et al., 1981). In the Commercial
Solvents process, methanol was produced from a HZ/CO, feedstock obtained from a
butanol fermentation plant at 30.6 MPa with metal oxide catalysts (Lee, 1990).
The BASF process used a high temperature catalyst operating at 3.50-400°C because
of its low catalytic activity. These catalysts had to operate at high pressures (25-35
MPa) because of low syngas conversion resulting from less favorable thermodynamic
equilibrium limitations at high temperatures. The catalysts are less active than copper-
zinc based catalysts but are more tolerant towards poisoning by compounds containing
sulfur.
I. Weruler/ Fuel Processing Technology 48 (19961 189-297 205

ICI introduced the more active copper-zinc based catalysts in 1966 (Davies et al.,
1966). Researchers in the 1920s had found that copper was the key component for the
methanol synthesis; newer developments made possible the use of copper/zinc oxide
catalysts. The source of syngas had shifted from coal to liquid or gaseous hydrocarbons
such as steam reforming of methane or naphtha. The syngas feed could be made pure
enough for copper catalysts (sulfur and arsenic compounds were now removed in
addition to chlorine). The copper-zinc oxide catalyst can be permanently deactivated at
high temperatures, so that proper control of reactor temperature is necessary. However,
even with the most carefully prepared catalysts, small amounts of methane, dimethyl
ether and traces of higher alcohols appear among the products.
ICI now had a new generation of low-pressure (< 10 MPa) and low-temperature
(220-270°C) Cu/ZnO/Al,O,-catalyzed methanol synthesis established as a commer-
cial process. A methanol process based on a Cu/ZnO catalyst was put into operation by
Lurgi in 1971; Haldor Topsije and BASF also practice low-pressure methanol synthesis
using the Cu/ZnO/M,O, (M is Al or Cr) catalysts. A basic difference between the ICI
and the Lurgi process is the type of reactor used; ICI uses multi-quench reactors and
Lurgi employs multi-tubular reactors. Modem methanol plants can yield about 1 kg of
methanol per liter of catalyst per hour, stated to be the “magic figure” for the
low-pressure process (Herman, 199 1).
The modem synthesis of methanol is very selective; > 99.5% conversion to methanol
is obtained. This high selectivity is quite an achievement in view of the fact that
methanol is thermodynamically the least probable product of the syngas conversion, i.e.,
other compounds are formed with a more negative free energy change than methanol. A
graphical example is given in Fig. 7, which shows the standard Gibbs free energy
change at 600°K (327°C) in kcal mol- ’ of carbon as a function of chain length n in the
product alcohols or paraffins (Wender and Klier, 1989).

nC0 + 2nH, + CH,(CH,),_ ,OH + (n - l)H,O (4.4)

(2n - 1)CO + (n + l)H, --j CH,(CH,)np ,OH + (n - l)CO, (4.5)

nC0 + (2n + l)H, -+ CH,(CH,),,_,CH3 + n&O ( 4.6)

2nCO+(n+ l)H, -+ CH,(CH,),,_,CH, + nCO, (4.7)

For long chains, alcohols plus water formed in Eq. (4.4) tend to have the same
standard free energy as the paraffins plus water of Eq. (4.6); alcohols plus CO, formed
in Eq. (4.5) have the same standard free energy as paraffins plus CO, in Eq. (4.7). The
difference for any single reaction type with water or CO, coproduct is given by the
standard free energy of the WGS.
For short chains, the free energies of formation for alcohols and hydrocarbons from
CO/H, diverge, with hydrocarbons being significantly more favored. The greatest
thermodynamic driving force is for production of methane plus CO, and the least
thermodynamic driving force, given by a positive Gibbs free energy, is towards
methanol. Because of the negative volume change of Eq. (4.1) and Eq. (4.21, the
methanol synthesis can be thermodynamically driven against positive free energy by
206 I. WeruIer/ Fuel Processing Technology 48 (1996) 189-297

GIBBS FREE ENERGY OF


CARBON MONOXIDE
HYDROGENATION TO

PARAFFINS + CO1

9 I I , 1 , ! I ,
I ; 3 4 5 6 7 6 5 IO

NUMBER OF CARBON ATOMS IN THE


HYDROGENATED PRODUCT

Fig. 7. Gibbs free energy of hydrogenation of carbon monoxide.

high pressures, but the catalyst must kinetically prevent formation of all thermodynami-
cally more favored products, i.e., hydrocarbons and C, + alcohols.

4.2.2. Existing technology


Methanol is considered a rather cheap chemical (about 50 cents per gallon in July
1996). New plants are being constructed all over the world. Significant new or revamped
annual capacity of neat methanol will be completed, engineered, or planned in the
1995-1998 period and beyond, evidently with confidence in the future world market for
methanol in fuel uses.
Methanol synthesis technology based on steam reforming of natural gas has advanced
to the point where room for further gains in energy efficiency is small. Overall
efficiency, because of continuing development, has risen from 58% in early plants to
72% with current designs (Abbott et al., 1989). But the added extra heat recovery
equipment has been accompanied by rising capital costs. World-scale methanol plants
are capital intensive, which has a large influence on production costs. The objective of
major producers of methanol is to reduce capital costs significantly without loss of
efficiency.
A flowsheet for a world-scale methanol plant with the distribution of capital costs
allocated to the principal process stages has been outlined by workers at ICI (Fig. 8).
Other methanol producers have similar flowsheets. Steps involving desulfurization and
distillation will not yield large savings. The focus of development for the past 15 years
by methanol producers is in the stage of manufacture of the synthesis gas. Over half the
I. Wender/Fuel Processing Technology 48 (1996) 189-297 207

Reforming/
Dcsulphurisatiou Gas Cooling Compression Synthesis Distillation

32% 22% 6%
t-ii,
I
FFEDSM)CK PRODUCT

i
r-27-t 0 -----

Steam Raising

Fig. 8. ICI methanol plant capital cost breakdown (ICI brochure).

investment in process equipment, with considerable cost-saving potential, is in the


reforming stage, with its associated gas cooling and steam raising equipment. We shall
deal with these aspects of the methanol synthesis only in passing.
A new proposed liquid-phase system for the production of methanol is being built in
Kingsport, TN, in the Eastman Chemical Co. facilities. This is the result of a joint
venture between Eastman and the developers of the technology, Air Products and
Chemicals. This is an example of the commercialization of a project co-sponsored with
the US Department of Energy. Use of a slurry bubble column instead of a multi-tubular
bed significantly reduces the capital cost of the reactor and decreases the compression
cost by reducing the pressure drop across the reactor. The fact that it is a fluidized
reactor allows better control of the reaction temperature, avoiding temperature runaways
and cold spots. Although this technology has been proven in different situations at the
LaPorte, TX facilities by Air Products, continuous operation would be the test that
methanol producers may be waiting to see.
The ICI and Lurgi low-pressure technologies are briefly discussed below. The earlier
high-pressure technologies are omitted in the present review because all new methanol
plants built after 1967 utilize the more modem low-pressure technology.
The service life of modem ICI catalysts is 3-4 years of continuous operation. Two
large ICI methanol plants are a part of the methanol-to-gasoline (MTG) complex in New
Zealand. Plants ranging in size from 47000 to 580000 tons per year are in operation,
and more are being built. A methanol plant using spherical reactors reported to weigh
about one-half as much as the usual cylindrical reaction vessels containing the same
catalyst volume is now in operation in Chile with a capacity of 2500 tons per day,
possibly the world’s largest single-train methanol synthesis plant (LeBlanc and Rovner,
1990). A methanol synthesis plant producing 750000 tons per year went on stream in
Venezuela in 1994 (Hydrocarbon Processing, 1992).
In most existing plants, natural gas is steam-reformed to hydrogen-rich syngas,
HJ(2CO + 3C0,) > 1. Alternatively, the naphtha reforming process yields a nearly
stoichiometric syngas, H,/(2CO + 3C0,) = 1. Coal or heavy oil can be partially
oxidized to syngas rich in carbon, HJ(2CO + 3C0,) < 1, which contains considerable
quantities of sulfur. Because copper based catalysts are extremely sensitive to sulfur
poisoning, the coal-derived syngas must be purified to bring the sulfur content below 0.1
ppm. This can be achieved by several purification processes which operate by physical
or chemical adsorption of acid gases, followed by a catalytic purification stage. The
208 I. Wender / Fuel Processing Technology 48 (1996) 189-297

Crude

Fig. 9. Diagram of a modem plant for methanol production from natural gas. (Reprinted with permission from
Chinchen et al., copyright 1990, American Chemical Society.)

adjustment of hydrogen-to-carbon ratios suitable for methanol synthesis can be achieved


by the WGS reaction between hydrogen-poor syngas and steam. A flow diagram for a
typical low-pressure methanol synthesis plant from natural gas is shown in Fig. 9
(Chinchen et al., 1990).
In the ICI process, the synthesis loop contains a circulator, a converter, and a heat
exchanger. There is a single catalyst bed with lozenge distributors for the injection of
cold quench gas located at optimal depths of the catalyst bed. Good mixing of gases and
temperature distribution in the reactor are ensured by this design. The distillation plant
consists of a unit that removes volatile impurities such as dimethyl ether, esters, ketones,
and iron carbonyl, and a unit which removes water and higher alcohols. After the first
Billingham methanol plant was operated at 5 MPa since 1966, the pressure of 10 MPa
was selected for a second, larger plant, with the carbon efficiency, defined as 100 X (mol
of methanol produced)/(mol of CO + CO, in the synthesis gas), 17% percent higher
than that of the 5 MPa process. Pinto and Rogerson (1977) have pointed out, however,
that the above pressure advantage in efficiency holds only for hydrogen-rich syngas
from natural gas or naphtha and not for coal-derived carbon monoxide-rich syngas.
Coal gasification conditions are usually such that syngas with high CO/CO, ratios is
obtained, which results in high carbon efficiencies over a wide range of pressures in the
synthesis loop. Methanol synthesis is adaptable, without loss of carbon efficiency, to
match a range of output pressures from various coal gasifiers. The general range of
operating conditions of ICI low-pressure methanol plants is 5-10 MPa, 220-280°C
I. Wen&r/ Fuel Processing Technology 48 (1996) 189-297 209

GHSV 5000-60000, and H,/(2CO + 3C0,) ratios 2 1 but adaptable to < 1. Eco-
nomic considerations have to take into account the energy and capital costs; an
advantage associated with the use of coal-derived syngas is that the syngas compression
can be avoided without penalty in carbon efficiency.
Although Cu/ZnO/Al,O, catalysts have been optimized for maximum selectivity,
space time yields, and long service life, they still deactivate, but detailed reports on the
rate of deactivation are not available. The main deactivation mechanism probably
involves poisoning by chemical impurities, includin g traces of iron carbonyl, even in
copper-lined reactors with pre-purified gases. The Lurgi low-pressure technology also
utilizes copper based catalysts, principally Cu/ZnO/CrzO,, the detailed preparation
and additional promoter composition of which are not disclosed. The Lurgi reactor is a
multi-tubular type, the tubes being filled with the catalyst and cooled with pressurized
boiling water on the outside. The reactor operates at 240-270°C and good heat transfer
to the pressurized boiling water (which yields steam) is achieved. Methanol is produced
at space time yields (STYs) close to 1 kg of methanol per liter of catalyst per hour. The
crude methanol product is condensed, cooled, and distilled. The reactor has been called
“quasi-isothermal” because of the smooth temperature profile along the reactor tubes,
with exotherm variations not exceeding 10°C (Supp, 1981). Catalyst life is 3-4 years.
The process can be easily adapted to utilize coal-derived syngas by replacing the steam
and autoreformers by coal gasifiers and a purification plant, mainly to remove sulfur
from the syngas.
In the low-pressure industrial plants, the feed composition is adjusted to contain
4-8% CO, in addition to H, and CO. Only 4-7 ~01% methanol is obtained per pass;
the remaining gas is recycled. Flow sheets outlining the Nissui-Tops&, the Mitsubishi
Gas Chemical and the Japan Gas-Chemical Company Methanol synthesis processes are
to be found in the book by Lee (1990).

4.2.3. On the mechanism of the methanol synthesis


Before discussing other catalytic systems for the synthesis of methanol, it is worth-
while to examine the mechanism of the commercial system using the Cu/ZnO/Al,O,
catalyst. The selective synthesis of methanol over this copper-containing catalyst is a
well-developed technology, but there are a number of scientific issues that pervade the
literature to this day. Questions include the nature of the active catalytic centers, the
reactive carbon-containing component (CO or CO,) and the overall mechanism of the
methanol synthesis.
The Cu/ZnO/AI,O, catalyst will make methanol in pure H/CO mixtures or when
only H,/CO, is used. Most workers in the West assumed that methanol was a product
of the reaction of CO in H,/CO/CO, mixtures and developed kinetics and mecha-
nisms based on this premise (Kung, 1980; Klier, 1982; Bart and Sneeden, 1987; Bridger
and Spencer, 1989; Chinchen et al., 1988). Russian workers (Kagan et al., 1975;
Rozovskii, 1980) however, obtained evidence that methanol was synthesized from CO,
and that little or no methanol was formed from H,/CO mixtures with a Cu/ZnO/Al,O,
catalyst.
A definite answer cannot be obtained from reaction kinetics because the WGS
reaction takes place during the synthesis of methanol, so that CO and CO, can
210 I. Wender / Fuel Processing Technology 48 (1996) 189-297

soecllic
Radioaclivity
K/m mot
0.14-
Inlet CO2
0.12-

5XiO4 1ox;04
Swce Velocity (h “1

Fig. 10. Specific radioactivities of carbon monoxide, carbon dioxide and methanol as a function of space
velocity. Reactant mixture 10% CO/IO% CO/SO% H,, containing 14C0, at 50 bar, 523 K (Chinchen et al.,
1987).

interchange. Chinchen et al. (1987) and Chinchen et al. (1990) used carbon-14 labeling
of 14C0 or 14C02 to attempt to find the answer. At low residence times, and thus very
low conversions, the methanol had the same 14C content as the CO, (Fig. 10). They
concluded that the Russian workers were correct; all methanol under these conditions
and at higher conversions is made from CO,. They also found that the WGS reaction
and the methanol synthesis do not have a common carbon-containing intermediate on
the surface of the catalyst. All methanol is made immediately from CO,; CO is
converted via the WGS to CO,, which is then converted to methanol.
Waugh (1992) has published a review of the results of many years of work at ICI on
the kinetics and mechanism of methanol synthesis over Cu/ZnO/Al,O, and other
oxide supports. There is little doubt that copper metal is the important catalyst
constituent. Waugh and his co-workers concluded that it is the CO, part of the CO, CO,
mixture that is the precursor to methanol, being adsorbed on the partially oxidized
copper as a symmetrical carbonate which is then hydrogenated to a formate species
adsorbed on the copper. The formate is possibly the longest lived intermediate leading to
methanol. Hydrogenation of this formate species is probably the rate-determining step in
the synthesis.
The specific activity of copper in the methanol synthesis is apparently not signifi-
cantly affected by the nature of the oxide support, so that there may be no unique
synergy to the Cu/ZnO combination. The role of CO in the methanol synthesis is
probably to maintain copper in a more reduced, hence more active, state than reduction
with hydrogen alone could achieve.
Nevertheless, we do read that the catalyst of choice is Cu/ZnO/Al,O,. Zinc oxide
seems to have a number of functions, perhaps performing somewhat better than other
I. Weder/Fuel Processing Technology 48 (1996) 189-297 21 I

supports such as CrO,, SiO, or MnO. The functions that zinc oxide plays have been
listed by Chinchen et al. (1990). Zinc oxide helps give the catalyst a high surface area,
especially with Al,O, present, is suitably refractory and hinders the copper particle
agglomeration that is inevitable in the life of the catalyst. Zinc oxide also acts as a sink
for sulfur and chlorides, which poison copper catalysts. In addition, the oxide interacts
with Al,O, in the catalyst to cut down on the conversion of methanol to dimethyl ether.
Chinchen and Spencer (1991) have proposed the following scheme for the synthesis
of methanol from carbon dioxide and hydrogen over a Cu/ZnO/Al,O, catalyst.

Hz(g) * 2H(a) (4.8)


CO,(g) * CO,(a) (4.9)
CO,(a) + H(a) # HCO,(a) (4.10)
HCO,( a) + 2H( a) e CH,O(a) + O(a) (4.11)
CH,O(a) +H(a) 8 CH,OH(g) (4.12)

CO(g) + O(a) *CO,(g) (4.13)

Hz(g) + O(a) @ H,O(g) (4.14)


Some of these reactions include several steps. The reaction in Eq. (4.1 l), the
hydrogenolysis of adsorbed formate, may be the critical step.
The synthesis of methanol from carbon dioxide over Cu/ZrO/Al,O, is an insensi-
tive reaction (Burch et al., 19901, meaning that the specific activity (turnover frequency)
under standard conditions depends little on catalyst parameters such as catalyst composi-
tion, metal crystallite size, nature of catalyst support, method of catalyst preparation,
promoters or poisons (impurities), etc. Sensitive catalytic reactions are strongly depen-
dent on catalyst parameters. Methanol synthesis over Cu/SiO,, for instance, is a
sensitive reaction (Nonneman and Ponec, 1990), and this may account for differences in
activity of various supported copper based catalysts.
In summary, methanol is formed from CO, in the H/CO/CO, feed. Copper is the
active catalyst component. The CO, molecule adsorbs dissociatively (CO, + CO f 0)
on the copper to form a bidentate formate which is hydrogenated to methanol in a
rate-determining step. The role of CO is to keep the copper in a highly reduced state.
Oxygen coverage of the copper is a function of the CO/CO, ratio. Although methanol
can be made from H/CO feeds on Cu/Zn/Al,O,, the rate is about 100 times slower
than when CO, is present. The presence of CO, enhances the durability of the catalyst;
in the absence of CO,, the catalyst deactivates more rapidly.
Studies devoted to the mechanism of the synthesis of methanol remain under active
investigation. The mechanism given above is still under vigorous study.
The kinetics of the methanol synthesis are complex and are affected by a number of
variables, such as the nature of the catalyst, the physical changes of the catalyst as the
reaction progresses, the composition of the gas (which is also constantly changing in the
reactor), temperature, and pressure. Modem methanol syntheses use copper-zinc low-
pressure catalysts and, as the synthesis reaction proceeds to thermodynamic equilibrium
very rapidly, the kinetic behavior of the catalyst may not be of great importance.
Commercial catalysts and data on their kinetic behavior are more or less proprietary.
212 I. Wender / Fuel Processing Technology 48 (I 996) 189-297

The production of methanol is an established commercial technology. Nevertheless,


constant improvements are being made in process technology and reactor design for
better recovery, in lower compression costs, and in processing of the raw methanol. Lee
(1990) and Satterfield (1991) have summarized information on these subjects.
It is particularly important to avoid contamination of the methanol catalyst by metals
that are Fischer-Tropsch catalysts. Care is required in catalyst preparation so as to
obtain pure methanol. Nickel and especially iron, both of which form volatile metal
carbonyls, Ni(CO), and Fe(CO),, respectively, must not be allowed to come into contact
with the syngas under reaction conditions. The carbonyls form at lower temperatures and
decompose to the metal and CO at higher temperatures, possibly in upstream heat
exchangers, etc. The presence of Fischer-Tropsch metals in the catalyst or on the
reactor walls will result in the formation of methane and also of higher hydrocarbons
and higher molecular weight oxygenated products. Methanol synthesis reactor shells are
typically lined with copper, although internal parts may be constructed of 18-8 stainless
steel.
The formation of higher alcohols (ethanol, propanol, etc.) can be suppressed by
careful exclusion of alkalis from the catalyst. Dimethyl ether is formed by the dehydra-
tion of methanol or by the hydrogenation of CO and may form in the presence of Al,O,.
However, if a Cu/ZnO low-pressure catalyst is employed with about 7.5% Al,O, as
stabilizer and promoter, ether formation is negligible. The copper-zinc catalysts vary in
metal composition and contain different amounts of other metals, such as Cr, Al, Mn, V,
Ag, etc. Nonneman and Ponec (1990) have shown that a pure copper catalyst is not
active for the synthesis of metals; the presence of small amounts of (alkali) promoters is
necessary for catalytic activity.
The Boudouard reaction, 2C0 @ C + CO,, which results in carbon laydown, is not
significant if the temperature is carefully controlled, despite a highly favorable thermo-
dynamic tendency.

4.3. Developments in methanol synthesis technology

Turnkey plants for the synthesis of methanol are available from a number of
companies throughout the world. Newer units for methanol synthesis built by separate
manufacturers generally embody what may be termed evolutionary improvements, so
that there is a continuous incorporation of new technology. These efforts maintain a
well-entrenched viable industrial synthesis of methanol. Completely new routes for the
synthesis of methanol are under study but they must show considerable advantages to be
considered for commercialization. Meanwhile basic and applied research on the mecha-
nistic aspects of the synthesis of methanol continues to attract the attention of scientists
and engineers.
Conversion of syngas to methanol in current commercial plants is limited to about
25% by thermodynamic considerations. A higher conversion of syngas per pass can be
achieved by operation at lower temperatures to increase allowed levels of methanol.
Methanol equilibrium as a function of temperature and pressure is shown in Fig. 11.
Efforts are made to keep operating temperatures as low as possible. More active
catalytic systems which operate at lower temperatures are being explored, but these are
I. Wmder/ Fuel Processing Technology 48 (1996) 189-297 21.3

2.5 6 10 20 30

Pressure, MPa

Fig. I 1.Methanol equilibrium as a function of temperature and pressure.

generally deactivated by low levels of CO, (< 100 ppm). The expense of removing
CO, from syngas hampers the adoption of these routes for methanol synthesis.
The thermodynamic constraint to methanol synthesis can be overcome by shifting the
equilibrium by removing methanol during reaction so as to attain higher conversions per
pass. This promising approach may well be incorporated into newer commercial units.
Reviews of newer catalytic systems have been published by Mills (1988, 1993) and by
Trimm and Wainwright (1990).

4.3.1. Alloys as catalysts


Baglin et al. (1981) discovered that Cu/ThO, catalysts prepared from intermetallic
CuTh, alloys were active in methanol synthesis. This system was also studied by Daly
(1984). Owen et al. (1987) probed into a larger class of intermetallic CUM,~ (M is Ce,
La, Pr, Nd, Gd, Dy, Zr, Ti, or Th) alloys as precursors of Cu/M,O, and Cu/MO,
catalysts, and found that the CuCe,,, CuLa,,,,, and CuPr,,, precursors also resulted in
highly active catalysts for the synthesis of methanol. Methanol synthesis using Cu/La
catalysts was achieved at temperatures as low as 100°C. These catalysts were severely
poisoned by very small amounts of CO, and often contained copper metal of low
dispersion (very large particles, Cu metal area < 1 m2 g- ‘1. An electropositive copper
species analogous to that proposed by Klier et al. (1986) for the Cu/ZnO catalyst, by
Baglin et al. (1981) for the Cu/ThO, catalyst, and by Shibata et al. (1984) for the
Cu/ZrO, catalyst was tentatively suggested to be the active component. Alternative
candidates for the site for methanol synthesis were suggested to be extremely small
copper particles ( < 1 nm diameter) or inter-metallic hydrides.
Because of the lack of tolerance to CO,, the Cu intermetallics-derived catalysts.
although very highly active, are not considered practical, as industrial syngas invariably
contains significant amounts of CO,, and CO, removal to very low levels adversely
affects the process economics. The oxygenate selectivities of these catalysts of 80-98%
are acceptable in their upper limit but not in the lower limit. Taking into account their
very high activities, the lack of precise determination of CO, effects at low concentra-
tions, and the lack of data on water effects on the synthesis, it seems that these novel
copper based catalysts have not been studied in detail and further research into their
possible improvement is warranted.
214 I.Wender/Fuel Processing Technology 48 (1996) 189-297

4.3.2. Cs / Cu / ZnO catalysts


Cu/ZnO catalysts can be doped with the more active alkalis to increase the yield of
methanol from syngas low in CO, (Klier et al., 1986; Nunan et al., 1988; Klier et al.,
1982). Catalytic activity depends on the level of alkali doping of Cs in the Cu/ZnO =
30/70 mol% catalyst and the ternary Cu/Zn/Cr = 30/45/25 mol% catalyst. Methanol
selectivity at 250°C and 7.6 MPa was over 97% with H.&O = 2.33 at optimum Cs
levels. Side products were mostly ethanol and methyl formate. The high selectivity to
methanol can be shifted towards higher oxygenated compounds by changing the reaction
conditions. The use of potassium instead of cesium gives poorer results.

4.3.3. Supported Pd catalysts


Poutsma et al. (1978) showed that methanol could be formed selectively over
Pd/SiO, catalysts with a 70:30 ratio of H,/CO at 260-350°C 5-110 MPa and
GHSV = 3300 h- ’. A number of different supports and alkali additives have been
studied. van der Lee et al. (1986) have discussed this somewhat surprising methanol
selectivity of Pd based catalysts in view of the tendency of Group VIII transition metals
to function as IT or methanation catalysts. Although Pd catalysts can selectively yield
oxygenates from H,/CO = l-3 syngas under conditions of commercial interest, yields
of methanol are low and industrial incentive to pursue their use is lacking.

4.3.4. Brookhaven liquid low-temperature and related syntheses


This methanol synthesis was originally based on NaI-I/RONa/M(OAc), catalysts at
the Brookhaven National Laboratory (Slegier et al., 1984; Sapienza et al., 1986). The
catalyst is completely dissolved in a solvent medium, unlike heterogeneous catalyst
systems. It is probable that the later catalytic systems are composed of an alkali
methoxide such as KOCH, and a nickel salt in a solvent such as triglyme. The synthesis
takes place at 80-12O”C, yielding > 90% syngas conversion per pass with over 95%
selectivity to methanol. A selectivity to methyl formate of up to 78 mol% can be
achieved in the same system (Mahajan and Mattas, 1992).
With nickel as part of the catalyst, the formation of the extremely toxic nickel
carbonyl must be guarded against. The process has been operated using syngas made by
partial oxidation using air to avoid the expense of an oxygen plant. The system is
sensitive to small amounts of CO, which, as stated, is expensive to remove from syngas
to low levels. However, this process is a significant departure from the commercial
synthesis of methanol (Trimm and Wainwright, 1990).
Marchiomra et al. (Marchiomra et al., 1988, Marchionna et al., 1992) have investi-
gated systems which produce methanol from syngas at similar low temperatures and
pressures (80-120°C lo-50 atm). The active catalytic system is again homogeneous
and involves a Ni(CO),/NaOCH 3 combination. Infrared investigations indicated that
attack by CH,O- on Ni(CO), under a hydrogen atmosphere produces (HNi(CO),)- and
methyl formate through the intermediate formation of (Ni(CO>,(COOCH,)-. Methyl
formate is also synthesized by the NaOCH,-catalyzed carbonylation of methanol. Then
the anionic hydronickel carbonyl species catalyzes the reduction of methyl formate to
methanol via the intermediate formation of formaldehyde (HCHO).
The extremely mild methanol synthesis conditions were postulated as being due
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 215

chiefly to the more facile reduction of the activated carbon-oxygen bond in methyl
formate as compared with that of CO coordinated to a metal.
This low-temperature methanol synthesis resembles that of the Brookhaven synthesis.

4.3.5. Concurrent methyl alcohol/methyl formate synthesis


Christiansen (1919) patented a two-step synthesis of methyl alcohol via the following
reactions
CH,OH + CO * NaOCH, + HCOOCH, (4.15)
HCOOCH, + 2H, P 2CH,OH (4.16)
Net 2H, + CO @ CH,OH (4.17)
The first step is catalyzed by an alkali alkoxide (NaOCH,, KOCH,, etc.); the
hydrogenolysis reaction in the second step is carried out in the liquid phase using a
catalyst such as copper chromite. Both reactions in Eqs. (4.15) and (4.16) have been
investigated in some detail (Tonner et al., 1983).
The alkoxide catalysts react with water and with CO, to form alkali formates (e.g.
HCOOK) and alkali methyl carbonates (e.g. KOCOOCH,), respectively (Liu et al.,
1988, Liu et al., 1989) so that great care must be taken to remove water and CO, to
levels of about 1 ppm and 10 ppm, respectively. These deactivating effects have deterred
investigations in which both steps, Eqs. (4.15) and (4.161, would be carried out in a
single reactor. However, Liu et al. (19881, Onsager (1984) and Aker Engineering (1982)
have reported the synthesis of methanol directly from syngas in a single reactor using a
catalyst comprised of an alkali methoxide and copper chromite; this has been referred to
as the concurrent synthesis of methanol. Under the conditions of this reaction, 1OO- 180°C
and 50-65 atm, there is an interaction between the carbonylation and hydrogenolysis
catalysts, with the overall rate being higher than that predicted for either reaction at
these conditions. The copper chromite catalyst regenerates the carbonylation catalyst
which has been deactivated by water and also removes water by the WGS reaction
(Palekar et al., 1993a). Indeed, potassium formate and potassium hydroxide used with
copper chromite are as active as the potassium methoxide/copper chromite mixed
catalyst in this concurrent synthesis (Palekar et al., 1993b). The reaction is carried out in
a slurry with consequent good heat transfer rates and with a rate of reaction at low
catalyst loadings which is comparable with that in the commercial synthesis. High per
pass conversions can be obtained, and the only significant products are methanol and
methyl formate, which can be separated easily. However, deactivation by water and CO,
is severe.

4.3.6. The liquid-phase methanol synthesis


Sherwin and Frank (1976) at Chem Systems, Inc., developed a liquid-phase methanol
synthesis (LPMEOH’“). This slurry-phase process has a number of advantages: it can
use syngas rich in CO as obtained from modem coal gasifiers, enhanced heat transfer of
the highly exothermic heat of reaction, and a high conversion per pass.
Researchers at the University of Akron have been investigating various aspects of this
process in a 1 liter slurry reactor that closely approximates CSTR conditions (Lee, 1990;
216 I. Wender / Fuel Processing Technology 48 (1996) 189-297

Lee et al., 1992). Using a Cu/ZnO/Al,O, catalyst, they established reaction chemistry
for diverse gas feeds including syngas with high H,, high CO, CO,-free and CO-free
environments. Their studies led to the suggestion of a liquid entrained reactor with a
catalyst slurry for commercial operation of the liquid-phase methanol synthesis.
In 198 1, the US Department of Energy began supporting research on a liquid-phase
methanol synthesis in a process development unit at the LaPorte, TX plant operated by
Air Products and Chemicals, Inc. (Brown et al., 1991). The Electric Power Research
Institute co-sponsored the program because of their interest in the possible coproduction
of methanol and electricity in integrated gasification combined cycle (IGCC) systems.
Operational conditions for methanol synthesis were studied in a 10 ton per day unit with
a Cu/ZnO/Al,O, catalyst. In situ catalyst activation was achieved with a l-10 pm
sized catalyst in Witco 70 oil with reaction temperatures of 225-265°C (Lewnard et al.,
1990). An extensive series of tests was run with various CO/H, ratios. Catalyst slurries
of 20-45 wt% were used at 250°C with CO-rich syngas. Methanol productivity was
optimum when the CO, content in the syngas with a H,/CO ratio of 0.69 was 5-8
mol% (Herman, 1991). This work demonstrated the viability of the slurry bed concept,
attaining a CO conversion to methanol of about 13% per reactor per pass in a
hydrogen-rich feed.
A primary output of the Eastman Chemical Company plant in Kingsport, TN is
carbonylation-derived acetic anhydride (see Section 7). At present, methanol for acetic
anhydride is produced from syngas by a gas phase process. The Air Products’ liquid-phase
methanol synthesis enables sufficient heat removal without the need for a a WGS
reactor, as mentioned earlier. Construction of a demonstration plant at Kingsport, with
partial funding by the US Department of Energy, is underway in a joint Air
Products/Eastman venture (Cook, 1995). Start-up of this plant may take place in 1997.

4.3.7. Methanol removal to obtain higher conversions


The synthesis of methanol is a classic example of an exothermic equilibrium-limited
reaction. There is recent emphasis on improving the synthesis of methanol by removing
the alcohol as it is formed, so lowering the thermodynamic constraints on methanol
conversion.
Berty et al. (1990) introduced the concept of “beating the equilibrium” by introduc-
ing a high-boiling inert solvent such as tetraethylene glycol dimethyl ether (tetraglyme)
into a reactor containing a fixed catalyst bed. The solvent, flowing concurrently with the
syngas stream, absorbs methanol as it is formed. This results in an equilibrium shift to
methanol, as the activity of methanol in the reactor remains low. This concept, based on
the classical idea that the equilibrium limitation in a reversible reaction can be overcome
by product removal on its formation, works well; syngas conversion is enhanced and
recycle is almost eliminated. The principle is sound but has not so far been used in a
commercial synthesis of methanol.
Westerterp and Kuczynski (19861, using a gas-solid trickle flow reactor (GSTPR),
have obtained methanol yields of 100% by removing methanol by selective adsorption
on an amorphous SiO,/Al,O, cracking catalyst in the reaction zone. The same group
(Westerterp et al., 1988; Westerterp et al., 1989) has employed another approach to the
separation of product methanol by use of a reactor system with interstage removal of the
I. Wmder/ Fuel Processing Technology 48 (1996) 189-297 217

alcohol between two packed tubular reactors in series (RSIPR). Methanol is absorbed at
reaction temperature in a countercurrently operated packed bed absorber using te-
traglyme (boiling point 275°C) as the solvent. Here a liquid rather than a solid is used to
absorb product methanol. The authors envisage significant energy and raw material
savings on scale-up of this system.

4.3.8. Condensing methanol principles


Hansen and Joensen (1991) at Haldor Topside A/S have carried out a study to
ascertain the best conversion to methanol obtainable per pass with a 2:l Hz/CO gas
mixture and with an acceptable space time yield. This approach is essentially based on
the finding that maximum conversion obtainable for a stoichiometric gas feed for the
synthesis of methanol decreases as the CO, content increases. A distinct carbon
conversion maximum is observed at about 2% CO, in the make-up gas, in line with the
findings of Klier et al. (1982).
These workers generated a syngas with about 67% H,, 29% CO and 3% CO,.
Conversion to methanol was calculated assuming that both the synthesis of methanol
and the WGS were in equilibrium. Calculated conversion corresponded to such high
partial pressures of methanol that condensation of the methanol and water would occur.
The calculated dew point lines as a function of temperature, conversion level and
pressure are depicted in Fig. 12, as are the equilibrium curves. Condensation takes place
above and to the left of the dew point line. Proximity to the critical point of methanol
(239.43”C, 79.9 atm) is responsible for deviation from an essentially straight dew point
line for 12.5 MPa at high conversion.
Results of laboratory experiments with conversions at 230°C and 240°C carried out in
isothermal reactors agreed well with those predicted for gas-phase equilibria (Fig. 13).
At 220°C and below, levels of conversion were high enough so that condensation
occurred and the dew point line was crossed. At 200-22O”C, conversions were higher
than those predicted by gas-phase thermodynamics. These findings are explainable only
by the highly non-ideal properties of the liquid phase on the catalyst, essentially by the
fact that the activity coefficients in the liquid phase deviate to a large extent from unity.

Fio 12. Methanol + shift equilibria and dew points. (Reprinted with permission of Elsevier Science Publishers
fr& Hansen and Joensen. 1991).
218 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

s .Bfeamld
m
lm
356
lso
374
an
392
zlo
410 z iii 24072
44°F

Fig. 13. Once-through conversion of CO and CO, 9.6 MPa; CO = 30%. (Reprinted with permission of
Elsevier Science Publishers from Hansen and Joensen, 1991).

Haldor Tops& carried out pilot plant work and concluded that their methanol
synthesis with a newly developed catalyst, combined with two-step reforming (a smaller
first-stage steam reforming followed by an autothermal reforming (Hydrocarbon Pro-
cessing, 1994) could be an economic way of producing methanol.

4.3.9. Methanol / dimethyl ether from syngas


The physical removal of methanol is one way of overcoming thermodynamic
limitations in the synthesis of methanol. Another interesting and promising way is to
convert methanol into a chemical species whose removal affects the equilibrium
conditions. The in situ dehydration of methanol to dimethyl ether (DME), in the
presence of an added acid catalyst such as gamma-Al,O,, is based on the second option.
The direct synthesis of DME from synthesis gas by dual function catalysts has been
investigated by a number of companies (Mobil Oil, 1975; Slaugh, 1983; Haldor Top&,
1985; Brown et al., 1991). The synthesis of DME in a single reactor is based on a
combination of an equilibrium-limited reaction (synthesis of methanol) and an equilib-
rium-unlimited reaction (methanol dehydration over the added acid catalyst) (Lee et al.,
1992). The reactions involve methanol synthesis, methanol dehydration, and the WGS
reaction

4H, + 2C0 e 2CH,OH AH = -43.2kcal (4.18)

2CH,OH @ CH,OCH, + H,O AH = -5.6kcal (4.19)

CO+H,0+H2+COZ AH= -9.8kcalmol-’ (4.20)

Net 3H, + 3C0 F? CH,OCH, + CO, AH = -58.6kcal (4.21)

Here one of the products in each step is a reactant for the other; hydrogen formed as
in Eq. (4.20) is a reactant for the methanol synthesis. This is a strong driving force for
allowing very high conversions of syngas for the overall reaction. The combination of
reactions results in a synergistic effect that eases the thermodynamic constraints for the
1. Wender/ Fuel ProceuGng TecAnolqy 48 (1996) 189-297 219

% Conversion (CO + CO 2 >


MeOH + DM3

MeOH

0 1 2 3 4 5 6 7 6

Pressure @IPa)

Fig. 14. Equilibrium conversion of syngas versus pressure at 240°C. (Reprinted with permission of Elsevier
Science Publishers from Hansen and Joensen, 1991).

methanol synthesis. The data in Fig. 14 show the advantage of combining the methanol
and DME syntheses.
Conversion levels similar to those achieved in conventional methanol synthesis may
be attained at much lower pressures. The process is thus an attractive front end for those
syntheses where DME may be employed as a feedstock instead of or with methanol
(Hansen and Joensen, 1991). DME has potential for use as a motor fuel, for energy
storage, as a propellant in aerosol containers, in the TIGAS process and as an
intermediate in the synthesis of chemicals (Rostrop-Nielsen, 1994).
Brown et al. (1991) compared the results of the liquid-phase DME (LPDME) system
with those of the LPMEOH process using a number of methanol and dehydration
catalysts. They found that the conversion of CO in the LPDME system was almost twice
as high as that achieved with LPMEOH and much higher than the conversion obtained
in the reaction of methanol alone. Work by Brown et al. (1991) and also the contribution
of Gogate et al. (1993) have shown that the single-step DME synthesis from coal-de-
rived syngas gives greater syngas conversions than that achievable in the LPMEOH
process. The route to methanol synthesis via DME has great potential for enhanced
overall process performance. Future work will be necessary to determine the effects of
process variables such as temperature, pressure and feed gas composition and also the
activity and stability of the mixed methanol/dehydration catalyst system. The reaction
of the methanol/DME system provides a challenge to conventional methanol manufac-
turers to design more active systems for the synthesis of methanol.

4.4. Fuel uses of methanol

Methanol is a fuel and a chemical and is used in the synthesis of a wide variety of
fuels and chemicals. As mentioned earlier, fuels are sold by the ton, chemicals by the
pound. The huge amount of methanol consumed worldwide per year (24 million tons
and still growing) can only mean that the main growth outlets for the uses of methanol
must be as a fuel, essentially in liquid transportation fuels.
The chief uses of methanol for chemicals will be discussed in a later section of this
review.
220 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

The main uses of methanol as a fuel, essentially in transportation fuels, are listed
below.
* Neat methanol or in mixture with gasoline in vehicles-commercial
- Methyl ethers as octane enhancers and fuels-commercial
- Higher alcohols (C ,X6) as octane enhancers and fuels-not now commercial
4.4.1. Methanol, neat or in mixture with gasoline
In the limited time of high gasoline prices, there was a strong incentive to use
methanol or methanol-containing products in gasoline. The ban on the use of tetraethyl-
lead and the Clean Air Act Amendments (CAAA) of 1990, which advocate a lowering
of aromatics in gasoline, resulted in a turn to the use of octane-enhancing oxygenates in
gasoline.
It appeared that a new potential market for methanol was opening up as a motor fuel;
emissions of CO, NO, and hydrocarbons are claimed to be low, and it has octane-en-
hancing properties. A review of research on and the industrial activity of methanol has
been published by Sinor Consultants (1990).
It is well to pause here to consider that the enactment of the CAAA in 1990 is
somewhat revolutionary in scope. The United States government, for the first time, is in
a sense regulating the composition and properties of gasoline and the kinds of trans-
portation fuels sold as these concern environmental and health policy (Peeples, 1991).
Provisions of the Act require manufacture and sale of clean fuels which will reduce
refueling, evaporative and exhaust emissions of ozone-forming compounds, airborne
toxic compounds and CO from motor vehicles.
The petroleum industry and car manufacturers have cooperated in meeting the
challenges raised by the CAAA. ARC0 led the way, followed quickly by other major
manufacturers of motor fuels. Refinery blending of methyl ethers such as methyl t-butyl
ether (MTBE) helped meet the regulated oxygen requirements, allowed the removal of
high-octane aromatics and helped lower or balance the vapor pressure and the light
olefin content of gasoline. The oxygenates that are most likely to be included in
gasolines in the United States are shown in Table 1 (Unzelman, 1992).
As will be mentioned later, alcohols with four to six carbon atoms blend well with
gasoline and can be added directly to gasoline.

Table 1
Oxvgenates most likelv to be used in United States gasoline
Blend Blend RVP/psi Blend b.p./“F Oxygen/W%
octane, (R + M)/2
Alcohols
TBA 101 10-15 181 49.9
Ethanol 113 17-22 172 34.7
Methanol 116 50-60 149 49.9

Ethers
MTBE 109 8-10 131 18.2
TAME 104.5 3-5 187 15.7
ETBE 110 3-5 161 15.7
I. Wender/Furl Processing Technology 48 (1996) 189-297 221

Methanol, however, has a number of drawbacks; the most serious is phase separation,
which would create serious problems in transportation, distribution and use in mixture
with gasoline. With pure methanol, 500 ppm of water is enough for two phases to form.
Corrosion caused by methanol-gasoline blends is also a problem.
In the early 198Os, ARC0 sold a methanol plus cosolvent mixture with gasoline in
parts of New York State and Pennsylvania under the trade name “Oxinol”. It consisted
of 4.75% of methanol and 4.75% of t-butyl alcohol (TBA) in unleaded gasoline.
Meanwhile Midwestern refiners sold ethanol-blended gasoline. A drop in demand for
Oxinol resulted, mainly through lack of consumer acceptance of methanol in gasoline
and higher methanol prices; it is no longer in use.
Racing cars use methanol with 15 ~01% of gasoline (M-8.5); gasoline is needed
because methanol burns without a visible flame. Neat methanol has been tested in
specially adapted cars in California and other places but distribution and costs, among
other problems, appear to hinder this approach from making a significant contribution to
our transportation fuel needs.

4.4.2. Methyl ethers and higher alcohols as octane enhancers and fuels
To overcome air pollution in certain cities in the United States, the concept of
reformulated fuels in automobiles has been introduced. Gasoline-fueled vehicles emit
small amounts of oxides of nitrogen (NO,), CO, hydrocarbons and fuel from tanks and
fuel-delivery systems (Seddon, 1992). By the mid- 1980s the concept of the use of
alternative, clean-burning fuels had gained increasing support. United States oil refiners
introduced “reformulated gasoline”, a concept embraced by the Clean Air Act of 1990.
According to this Act, 38 US cities with high levels of CO would have to use
reformulated gasoline (RFG) with an average oxygen content of 2.7 wt%. Gasoline in
cities with high levels of ozone would have to reduce volatile olefins by 15% by 1995
compared with 1990 levels. Oxygen content in cities with high ozone levels would have
to be an average of 2 wt%.
The CAAA legislation resulted in the introduction of alternative fuels not derived
from petroleum. Alternative fuels could be mixtures comprised of 100% gasoline, with a
composition primarily limited to a balanced mixture of normal butane, isoparaffins and
toluene to sustain octane ratings with low photochemical reactivity and low emissions. A
very large overhaul of refining operations would be required to achieve this goal.
Refiners responded with the introduction of RFG which would have performance
equivalent to that of alternative fuels, so avoiding the enormous capital expenditure and
massive write-off on fuel supplies and distribution (Seddon, 1992).
An illustration of RFG proposed compositions which pre-date the CAAA provisions
is given in Table 2 (Seddon, 1992). They were formulated with the purpose of lowering
levels of CO, which has a propensity to form ozone; an average of 2.7 wt% of RPG
would be required in cities with high CO levels. Volatile olefinic compounds (VOCs)
would be eliminated by 1995 compared with 1990 levels and benzene would be limited
to 1 ~01%. Total aromatics would be limited to 25 ~01%. The data in this table depict the
refiners’ efforts towards the marketing of oxygenates in gasoline with limited photo-
chemically active compounds.
All of the reformulated gasolines in Table 2 contain the compound methyl t-butyl
222 I. Wender / Fuel Processing Technology 48 (1996) 189-297

Table 2
Some previously proposed reformulated unleaded gasolines (Seddon, 1992, reprinted with permission from
Elsevier Science Publishers)
Company Arco Phillips Conoco Shell Exxon Chevron
Product EC- 1 Unleaded plus RXL SU2OOOE Supreme plus Supreme
Octane, (R + M)/2 a 88 87 86.5 91 89/91 91
Oxygen b/wt% I I .o-2.0 0.0-2.5 1 1 1
Oxygenate MTBE MTBE MTBE MTBE MTBE MTBE
Aromatics/vol% c 20 < 20 < 25 no limit no limit no limit
Olefms/vol% < 10 < 10 no limit no limit no limit no limit
Benzene/wt% < 1.2 I < 2.4 no limit no limit no limit
RVP/psig 8 8.5 8.5 8.0-8.5 8.5 reduced

a (RON + MoN/2.
b Lower value for summer, higher for winter.

ether (MTBE), the most common oxygenated compound in use in gasolines in the
world. In the period 1982-1992, MTBE production increased at an annual rate of 30%.
MTBE production, expected to reach 31 million metric tons by 1995, will then probably
level off into the late 1990s.
MTBE is produced by the acid-catalyzed reaction of methanol and isobutene;
methanol contributes 35.3 wt% of the ether (Eq. (4.22)). Mixed olefin streams from FCC
units are used in the manufacture of MTBE. There have been concerns about obtaining
an adequate supply of isobutene but ongoing research promises to solve this problem.
MTBE is an excellent octane booster, has little effect on volatility, helps reduce CO
emissions and has no adverse effect on vehicle system materials. It has favorable
physicochemical properties compared with alcohols, including a higher energy content
than methanol or ethanol.

CH,OH + (CH,),C= CH$CH, - 0 - C(CH,), (4.22)


isobutylene MTBE

There is a growing interest in t-amyl methyl ether (TAME), which uses the C,
fraction from FCC units and from steam crackers (ethylene plants). It is estimated that
TAME could supply about 70 000-90000 bpd of oxygenates in gasoline before 2000
(Unzelman, 19921, perhaps eventually equalling about a tenth of total ethers.
ETBE (ethyl tertiary butyl ether) and TAEE (tertiary amyl ethyl ether) can play roles
similar to that of MTBE. Indeed, they have the advantage of lower blending vapor
pressure (a lower RVP is of great importance) and higher boiling points.
At present, fuel ethanol is not produced from syngas but rather by fermentation of
starch. Gasohol, a 9O:lO mixture of gasoline and ethanol, now constitutes about 1% of
the gasoline pool. The manufacture of ethanol by fermentation carries a federal and
often a state subsidy.
To a large extent, then, the continuing growth in methanol synthesis is tied to the
increasing use of methyl ethers in reformulated fuels.
The synthesis of mixtures of methanol and higher alcohols from syngas by modifica-
tion of reactions in the synthesis of methanol and by alkali promotion of methanol
Table 3 .4
Research and development status in 1988 for Syngas to alcohol projects (Courty et al., 1990)
f
The catalysts Range of operating conditions
k
Company Key elements Alkali metals Others GHSV/h- ’ T,“C P /MPa H, /CO CO, removal State of development 7

I Enichem Zn K Mn, AI 3oOt- I500 350-420 12-16 OS-3 No Industrial plant 1


Snamprogetti Cr Na MO, Ti (18) (I5000 t/y, Italy) <
2
Haldor Topwe CU La, VCe x
$
2 Lurgi CuAlZn K, others No 2OOt-4000 270-300 7-10 l-I.2 Yes (1% synth. loop) Bench scale E’
h

3 IFP cuco NaK Al, Cr, Zn, 300%6OQO 260-320 6-10 l-2 Yes (OS-3%) Demonstration 2
$
and others unit 0
5
Idemitsu-Kosan CuNi K Ti, Mn, Zn, (synth. loop) (20 bbl/dayJ B
others
2
;=
4 Dow Chemicals MoCoS K H*S 5000-7000 290-310 12-14 1.1-1.2 Yes Bench scale
z
(Union Carbide) (28 I cat.) 8

5 C, Chemistry Group Rh, Mn + Cu, Zn Li 30000-45000 260-280 5 I .4 No Lab. scale %


(Japan) !!.I
s
224 I. Weder/ Fuel Processing Technology 48 (1996) 189-297

synthesis catalysts has been known since 1913 (Mittasch and Schneider, 1913). Indeed,
from 1927 to 1945, mixed alcohols were manufactured in Germany using alkalized iron
catalysts.
Early work on the higher alcohol synthesis (HAS) has been reviewed by Natta et al.
(19571, Anderson et al. (1952) and Stiles (1977). In the 1980s reviews of the HAS were
published by Haag et al. (1987), Xiaogding et al. (1987), Mills (1988) and Wender and
Klier (1989). More recently, excellent reviews of the HAS have been published by
Courty et al. (19901, Herman (1991), Forzatti et al. (1991) and Mills (1993).
The HAS consists of the conversion of syngas to methanol plus higher alcohols;
methanol is usually the chief component of these higher alcohols (Smith and Anderson,
1983; Stiles et al., 1991). It is pertinent to realize that, although there is growing use of
the addition of MTBE and other methyl ethers to gasoline in reformulated fuels, the
Snamprogetti/Enichem, Tops&e, Lurgi, Dow/Union Carbide, Institut FranGais du Pet-
role, and Vulcan projects on the HAS were all abandoned before a single commercial
plant was built (Notari, 1991). The status of research and development in 1988 of the
various projects for the synthesis of higher alcohols is given in Table 3. Mixed alcohols
are inferior as blending agents in gasoline and, although their use is not precluded, they
are not presently regarded as commercially promising. The composition of some of the
higher alcohols prepared from syngas is given in Table 4 (Mills, 1993).
The presence of lower alcohols in the HAS mitigates against their use. The large
amounts of methanol, in particular, in the C,-C6 alcohols increases their susceptibility
to extract out of gasoline under wet storage. Ethanol, for example, is not usually blended
with gasoline at the refinery but is added to gasoline just before delivery to the gasoline
pump (Piel, 1993).
El Sawy (1990) has published an evaluation of mixed alcohol production catalysts
and processes. It is difficult to rank the various HAS processes because of their different
responses to areas of concern for the EPA. These include responses and activity in
relation to exhaust emissions, evaporative emissions, driveability, materials compatibil-
ity and fuel stability (phase separation).
Methyl ethers such as MTBE and, to a lesser but growing extent, TAME and
especially ETBE are regarded as having much better properties as gasoline blending
agents. The ethers are more effective and have less costly pathways to deliver alcohols
and energy from domestic sources such as natural gas or coal into the liquid transporta-
tion fuel system of the United States (Piel, 1993).

Table 4
Composition of some fuel alcohols from syngas
Alcohol/% C, C, C, C, C, Catalyst
MAS (SEHT) 69 3 4 13 9 K/Zn/Cr
Substifuel (IFP) 64 25 6 2 2.5 K/&/Co/Al
Octamix (Lurgi) 62 7 4 8 19 akali/Cu/Zn/Cr
HAS (Dow) 26 a 48 14 3.5 0.5 CoS/MoS, /K

a Methanol can be recycled to extinction, increasing the amount of ethanol.


Dow: straight chain alcohols.
Lurgi: isobutanol is 70% of C, alcohols.
I. Wen&r/ Fuel Processing Technology 48 (1996) 189-297 225

Higher alcohols, with four to six carbon atoms, are good gasoline blending agents and
may be added directly to gasoline. They are essentially as good as methyl ethers in
blends with gasoline. Isobutanol, t-butanol (TBA), pentanols and hexanols may be used
in reformulated fuels as such. They have low water solubihty, low RVP properties, high
energy contents and are completely soluble in gasoline.
The potential octane contribution of oxygenates in gasoline as a function of their
oxygen contribution in gasoline is illustrated in Fig. 15 (Pie], 1994). Alcohols are
currently limited to 3.7 wt% oxygen (methanol to 2.5 wt%); ethers are limited to 2.7
wt%. With these limits, ethers and alcohols could replace from 1.5 to 20% of gasoline.
The replacement potential as energy is shown in Fig. 16 (Piel, 1993). The energy
replacement potential of the oxygenates is slightly lower than the volume replacement.
The C, alcohol isobutanol is of particular interest. It can be synthesized in fairly
good selectivity directly from syngas, is a good blending agent with gasoline, and can be
used in the manufacture of MTBE. There is a large amount of isobutanol in the C,
fraction of mixed alcohols; as much as 70% has been reported by Lurgi (Mills, 1993).
But the selectivity in the synthesis of isobutanol from syngas is not high in the HAS.
Thermodynamically, the amount of isobutanol from syngas can be high under certain
conditions; it is the dominant alcohol up to 500°C (Roberts et al., 1992). The concept of
reacting methanol and isobutanol to form methyl isobutyl ether (MTBE) is attractive;
unfortunately, methyl isobutyl ether (MIBE) in unleaded gasoline has the very low
BRON (blending research octane number) of 60.8 and a BMON (blending motor octane
number) of 67.0 (Herman et al., 1994). The corresponding values of MTBE are 120.1
and 96.3 respectively.

10

IPTBE

: ETBE

.’ :
: .’
MTBE
:
: :

Ethanol

T-Butanol

Current Oxygen Limit

2 3 4 5 6

Wt % of Oxygen in Gasoline

Fig. 15. Potential octane contribution of oxygenates in gasoline (Piel, 1994). IPTBE, isopropyl butyl ether;
ETBE, ethyl t-butyl ether. MTBE, methyl t-butyl ether; t-butanol, t-butyl alcohol.
226 1. Wender / Fuel Processing Technology 48 (1996) 189-297

35
?? Current Oxygen Level
IPTBE
30 - :
:
:
: ETBE
$ .’ :
g 2s
,’ :
Lu MTBE
: :
_m : ,’ TAA
::
E
2
20_
.’ : ,’.’
.’ : : TBA

0 1 2 3 4 5 6 7

Wt % Oxygen in Gasoline

Fig. 16. Potential volume contributions of oxygenates in gasoline. TAA, t-amyl ether, IBA, isobutyl alcohol
(Piel, 1993).

The market demand for oxygenated fuels shows a definite trend to higher molecular
weight ethers and the more highly branched higher alcohols which have a high energy
density, high octane rating, low RVP and high tolerance for water.
Typical properties of fuel alcohols and fuel ethers in gasoline are given in Table 5
(Piel, 1994).
The gasoline of the future is still of uncertain composition. Very likely, a number of
combinations of hydrocarbons and oxygenates will be used, but the situation is fluid at
this time. There will be many formulations proposed and used, for the transportation fuel
situation is still undergoing rapid change.

4.5. The Fischer-Tropsch and related reactions

4.5. I. Background and scope


The Fischer-Tropsch process was the first one to be used to convert syngas to liquid
fuels on a large scale. The two major reactions involving the hydrogenation of CO to
liquids are the Fischer-Tropsch synthesis (FTS) and the synthesis of methanol. The two
reactions differ in a major fundamental way: CO is adsorbed dissociatively, to a great
extent, in the ITS; in the synthesis of methanol, CO is exclusively adsorbed associa-
tively (without splitting of the C-O bond).
The seminal paper in regard to the catalytic hydrogenation of carbon monoxide (to
methane) was published by Sabatier and Senderens (1902). In 1913, BASF obtained
patents on the preparation of hydrocarbons and related oxygenated compounds by
I. Wender/Fuel Processing Technology 48 (1996) 189-297 227

Table 5
Typical properties of fuel alcohols in gasoline (Piel, 1994)
Methanol a Ethanol Isopropanol t-Butanol Isobutanol t-Amy1
and co-solvent alcohol
Octanes: blending (R + Ml/2 108+ Il.5 106 100 102 97
Vapor pressure near RVP ( lOOoF) 4.6 2.3 1.8 1.7 0.6 0.7
Blending RVP b 31f 18 14 9 5 6
Boiling pt./OF 148 173 130 181 226 216
Density/(lb gal- ’) 6.63 6.61 6.57 6.59 6.71 6.79
Energy density/(MBTU gal- ’) (LHV) 56.8 76.0 87.4 94.1 95.1 100.1
Heat of vapor/(MBTU gal-‘) at NBP 3.14 2.39 1.90 1.55 1.67 1.58
Oxygen content/wt% 50 34.8 26.7 21.6 21.6 18.2
Sohtbility in water/wt% I I I I 10.0 11.5

Typical properties of fuel ethers in gasoline


MTBE ETBE DIPE TAME IPTBE AEE
Octanes: blending (R + M)/2 110 112 105 105 113 100
Vapor pressure
near RVP ( 100°F) 7.8 4.0 4.9 2.5 2.5 1.2
blending RVP 8 4 5 2.5 2.5 1
Boiling pt./OF 131 161 155 187 188 214
Density/(lb gal- ’) 6.19 6.20 6.10 6.41 6.30 6.39
Energy density/(MBTU gal- ‘) (LHV) 93.5 96.9 100 100.6 NA NA
Heat of vapor/(MBTU gal- ’) at NBP 0.86 0.83 0.90 0.90 NA NA
Oxygen content/wt% 18.2 15.7 15.7 15.7 13.8 13.8
Solubility in water/wt% 4.3 2.6 2.0 2.0 NA NA

I: Infinite solubility; NA: not available.


’ Typical for methanol waivered blends with cosolvents.
b Blending RVP for 2.7% oxygen or higher in gasoline.

hydrogenation of CO at high pressure, mostly on oxide catalysts. In 1923, Fischer and


Tropsch obtained a large amount of oxygenated products from syngas using alkalized
iron and other catalysts; they called the product “Synthol”. In the same year, Fischer
and Tropsch synthesized higher hydrocarbons using nickel and cobalt at one atmosphere
(BASF had patents covering higher pressures). Several countries, including the United
States, England and Japan, initiated studies on the FTS starting about 1926. Fischer and
his co-workers at the Kaiser Wilhelm Institute for Coal Research (KWIK-now the
Max Planck Institute for Coal Research at Mtilheim) developed Ni-ThO,-kieselguhr
and Co-ThO,-kieselguhr catalysts for the FTS. Nickel catalysts gave too high yields of
methane and work with this catalyst was discontinued (Anderson, 1984).
In the decade following 1935, the FTS was operated in Germany using cobalt
catalysts, usually in the 0.5-2.0 MPa range (so-called medium pressure). Cobalt was
expensive and in short supply; after World War II cobalt catalysts were replaced with
alkalized iron catalysts. During this period, Pichler (1952) produced high molecular
weight hydrocarbon waxes using ruthenium catalysts at high pressures. Fischer and
Pichler, while investigating new catalysts and studying the mechanism of the FTS,
developed a process called the Isosynthesis (Cohn, 1956). Roelen (1943), in the course
228 I. Wender/ Fuel Processing Technology 48 (1996) 185-297

of studies on the mechanism of the ITS, discovered the homogeneously catalyzed


hydroformylation (0x0) reaction, now used extensively around the world to produce
aldehydes and alcohols from olefins and syngas (this reaction is discussed in Section 5).
The production of ITS products in Germany during WWII reached a maximum of
about 650000 tons per year in 1944. The standard catalyst used in all the plants had the
composition in relative mass units of Co:ThO,:MgO:kieselguhr = 100:5:8:200. Bomb
attacks reduced the output of these plants to about 10000 tons per year by early 1945.
All these FTS plants had ceased operation by the end of the war.
After the war, Ruhrchemie and Lurgi jointly developed a process based on the use, at
medium pressure, of iron catalysts in fixed bed units which was commercialized at Sasol
1 and later referred to as the Arge process. In 1948, Koelbel and Englehardt developed a
hydrocarbon synthesis from CO and steam to utilize the then available CO-rich blast
furnace gas (Koelbel and Englehardt, 1951; Koelbel, 1957). In 1935, the British Fuel
Research Station initiated work on the FYI’Susing both cobalt and iron catalysts in fixed
bed reactors (Hall et al., 1952). At about the same time, work on the FTS was started at
the Bureau of Mines in the US (Anderson, 1956, Anderson, 1984).
In 1950, US firms built a commercial (365000 tons per year) ITS plant, developed
by Hydrocarbon Research, in Texas (Keith, 1946). Syngas was obtained by reforming
natural gas. The process used a fluidized bed reactor and, after operating troubles
requiring the design of a new reactor, went on stream in 1953. Iron impregnated with
about 1 wt% of K,CO, was used as a catalyst. The price of natural gas rose during this
period and it became more profitable to sell the gas than to convert it to gasoline and
chemicals; the plant, using what was called the Hydrocol process, was shut down soon
after operation was deemed satisfactory.
At about the same time, in South Africa, the Sasol 1 Fischer-Tropsch plant was built
and commercial operation started in 1955.

4.5.2. On the chemistry of the Fischer-Tropsch synthesis (FTS)


Schulz (1985) has provided a systematic definition of FIS chemistry that differenti-
ates it from other reactions that involve the hydrogenation of CO: methanation, the
synthesis of methanol, the Isosynthesis and the hydroformylation reaction. The
Fischer-Tropsch synthesis (FTS) is catalyzed heterogeneously on metal catalysts which
are extremely sensitive to poisoning by sulfur, on which CO is strongly chemisorbed,
and which form metal carbonyls at high pressures but at temperatures too low for the
FTS. These metal catalysts must be able to dissociate CO (split the carbon-oxygen
bond). On certain catalysts under certain conditions, dissociative adsorption of CO is the
main route; on Pd, Pt and Cu, associative adsorption is most likely, and on some
catalysts (rhodium is probably the best example) CO is adsorbed in both ways. A better
understanding of these observations can be gained from an examination of Fig. 17
(Brod&n et al., 1976).
The dotted line at the left separates those metals (Cr, MO, etc.) that dissociatively
chemisorb CO at room temperature. As the temperature is increased, the line is shifted
to the right. At 200-300°C the borderline is between nickel and copper, rhodium and
palladium, and osmium and iridium. It is fairly well established (Kellner and Bell, 1981;
Katzer et al., 1981; Biloen and Sachtler, 1981) that metals to the left of the line catalyze
1. Wender/ Furl Processing Technology 48 (1996) 189-297 229

I
I 1

Cr Mn Fe : Co Ni j cu
,--_-: r_____l
I
MO Tc i Ru Rh ’ Pd
a Ag
,---_:
r----J I
w 1 Re OS j Ir Pt Au

i c
ROOM SYNTHESIS TEMPERATURE
TEMPERATURE 473-573 K

Fig. 17. Metals that adsorb CO dissociatively and non-dissociatively at ambient temperature and Fischer-
Tropsch reaction temperatures (BrodCn et al., 1976).

the ITS whereas those to the right are methanol synthesis catalysts. At still higher
temperatures, however, the line is shifted further to the right, and at these temperatures
CO is dissociatively chemisorbed even on copper and palladium. Poutsma et al. (1978)
showed that Pd, Ir and Pt catalyzed the synthesis of methanol with high selectivity at
higher synthesis gas pressures.
The IT reaction yields a wide spectrum of hydrocarbons and oxygenated compounds.
The major constituents of the hydrocarbons are paraffins and olefins, and primary
alcohols are usually the chief oxygenated products. Straight-chain paraffins, along with
some 2-methyl-branched paraffins, predominate among the saturated hydrocarbons; the
major olefin components are terminal olefins.
A large number of reactions occur during the IT reaction; the major ones are given
below
Paraffins(2n + l)H, + nC0 -+ CnH2n+2 + nH,O (4.23)
Olefins2nH, + nC0 -+ C,H,, + nH,O (4.24)
Alcohols2nHz + nC0 + C,H*,,+ ,OH + (n - l)H,O (4.25)

WatergasshiftCO + H,O -+ CO, + H, (4.26)


Boudouard reaction 2C0 + C + CO, (4.27)
Coke depositionH, f CO + C + H,O (4.28)

For catalysts that are good water gas shift (WGS) catalysts, such as alkalized iron, the
water formed in Eqs. (4.23), (4.24) and (4.25) reacts with CO to form H,, so that the
apparent HZ/CO usage becomes smaller. For a catalyst such as cobalt, which is not a
good WGS catalyst, water is the main reaction product.
The carbon number distribution of organic compounds is extremely wide and much
effort has been expended in finding ways to improve the selectivity of the FfS for
desirable products: gasoline, diesel fuel, C,-C, olefins and alcohols. Glycols are not
produced in the FI reaction. Aromatic products become significant only at high reaction
temperatures.
When several reactions involving the same reactants are thermodynamically possible,
the yields of various products depend upon their relative reaction rates and on the rates
of ensuing reactions. At the temperatures usually employed in the ITS, the actual
230 I. Wender/Fuel Processing Technology 48 (1996) 189-297

selectivity found differs very much from that expected from thermodynamic calcula-
tions. An excellent summary of the thermodynamics of the FTS has been given by
Anderson ( 1984).
Anderson (1984) and Frohning et al. (1982) have pointed out that data on the heats of
reaction are important, as the FTS is strongly exothermic. Indeed, removal of the heat of
reaction, about 25% of the calorific value of the syngas, is probably the chief problem in
practical application of the synthesis. Excessive catalyst temperatures can lead to
undesirable products, carbon deposition, catalyst deactivation or catalyst disintegration.
The heats of reaction per carbon atom of the products vary only slightly with
temperature, increasing with increasing carbon number for paraffins and decreasing with
increase in the carbon number of the olefins formed. Enthalpy changes for reactions
yielding hydrocarbons and CO, exceed those for the corresponding reactions yielding
H,O. The enthalpy difference, about 9 kcal, is due to the WGS reaction, which is
relatively independent of temperature.
The heats of reaction for the formation of alcohols, also somewhat independent of
temperature, in kcal mol- ’ per carbon atom are -23.9 for methanol and -29.5 for
ethanol (Wagmann et al., 1982).
The FTS is normally carried out at pressures from 0.1 to 4 MPa with c- 425°C as the
upper temperature limit. As FT reactions involve a decrease in the number of moles,
conversions at a given temperature increase rapidly with increase in pressure. The
pressure and temperature limits of the reaction are determined by changes in selectivity
and in the rate of catalyst deterioration.
Anderson (1956) presented equilibrium data for the hydrogenation of CO to hydro-
carbons and water, plotting AGO/n (AC per mole per carbon atom) versus tempera-
ture. The AGO/n curves for paraffins are about parallel but become more positive with
increasing carbon number (Fig. 18). Below 573 K, AGO/n becomes more negative with
increasing carbon number for olefins. In the FT temperature range, AC/n values for
aromatic and saturated cyclic hydrocarbons, which are not significant FI products, are
essentially the same as those for olefins of the same carbon number.
Large equilibrium yields are possible for all primary, straight-chain alcohols except
methanol. Methanol is formed in very small amounts in the FIS.
As kinetic rather than equilibrium considerations are controlling in the FI’S, only
limited amounts of useful information are furnished by thermodynamics. The products
obtained depend largely on the selectivity of the catalyst. But certain conclusions
derived from thermodynamics help us understand these syngas reactions
?? Methane is the preferred product at all FIX conditions.
* The relative stability of various compounds varies with the temperature of the FTS
(25-425°C). The order is paraffins > olefins > alcohols.
?? Although methane is the preferred paraffin, for olefins the favored products are
compounds with the highest carbon number. Above = 425°C at 0.1 MPa this
reverses and light olefins are favored.
* The favored alcohol is the one with the highest carbon number.
?? The equilibrium conversion of syngas increases with pressure. Owing to the effect of
pressure (and temperature) on the activity and selectivity of the catalyst, the upper
limits of usefulness for iron catalysts are 3-4 MPa and = 400°C.
231

-3 -
1 k k b 1
10 I
12

Carbon Number

Fig. 18.Plot of AC/n versus carbon number and temperature (adapted from Stall et al., 1969; Janaf, 1971).

The water gas shift reaction is favored under FIS conditions; iron is a particularly
good catalyst for this reaction.
The Koelbel-Englehardt reaction (Koelbel and Englehardt, 195 11, the reaction of CO
plus H,O to produce hydrocarbons, is thermodynamically more favored than the
usual ITS from syngas. This reaction will be discussed later.
The concentrations of n-paraffins and terminal olefins greatly exceed equilibrium
values in the FTS. Subsequent isomerization reactions are not important.
The reaction of ethylene and of ethanol with syngas is thermodynamically possible at
FI temperatures. Incorporation of higher olefins or higher alcohols is less favored.
Thermodynamically, methanol incorporation is favored over that of ethanol. The term
“incorporation” denotes the building-in of an organic molecule together with CO
andH, intheFIS.
As with higher hydrocarbons and olefins, the amounts of oxygenated compounds
(alcohols, aldehydes, acids, ketones) are also formed in much higher concentrations
than predicted from thermodynamic calculations. However, these oxygenated com-
pounds readily interact with each other under ITS conditions (Weitkamp and Frye,
1953).

CH,CH,OH e CH,CHO + H, (4.29)

CH,CHO+H,O~CH,COOH+H, (4.30)

CH,COCH, + H, e CH,CHOHCH, (4.31)

Olefin hydrogenation and alcohol dehydration are thermodynamically favored under


232 I. Wet&r/ Fuel Processing Technology 48 (1996) 189-297

FT conditions. Olefins and paraffins may be formed by dehydration of alcohols and


hydrogenation of the olefins, respectively, and by primary reactions.
. The Boudouard reaction, in which CO can form free carbon and CO,, is favored
under all synthesis conditions. It is, however, possible to suppress this reaction in
many FI reactions.
It is difficult to arrange experiments amenable to analysis on a group of reactions of
such a high degree of complexity as the FI reaction. Anderson (1956) has discussed the
kinetics of the FIS, their goal being to find a fundamental rate equation relating the
differential rate to the partial pressures or concentrations of reactants and of products.
However, there are too many catalyst variables affecting the kinetics, including the
method of preparation, composition and ageing, and also the operating conditions and
transport phenomena involving the catalyst and the reactants, intermediates and prod-
ucts. Starch et al. (1951), Anderson (1956) Anderson (1984), Kuo (1984), Vannice
(1982), and Dry (1981) have summarized kinetic expressions derived for nickel, cobalt,
iron and ruthenium catalysts.
Vannice (1975), using Al,O, as the support, reported that the activities of Group VIII
metals declined in the order Ru, Fe, Ni, Co, Rh, Pd, Pt and Ir. With SiO, as the support,
the activity declined in the order: Co, Fe, Ru, Ni, Rh, Ir and Pd (Vannice, 1977).
Most studies, however, have been carried out using small laboratory units, usually at
low levels of conversion. Dry (1981) Dry (1996) reinvestigated the FTS over iron
catalysts in large pilot plant reactors over wide ranges of conditions at high levels of
conversion (including commercial conditions). He found that, at low levels of conver-
sion, the rate was directly proportional only to the partial pressure of H,; at higher
levels of conversion, water vapor pressure had a strong negative effect, but CO, had
only a small influence on the rate equation

r=mPcoPH2/(Pco+~PH*o)
This resembles the equation proposed by Anderson (1956) and fits Sasol research
findings. When PHZO is low (at low conversion levels), the equation becomes
r=mPH2 (4.33)

With the full rate equation plus the WGS, the reaction profiles of pilot and
commercial-scale reactors (fixed and fluidized bed units) could be accurately simulated
in a computer model (Dry, 1981). Satterfield (1991) has proposed a somewhat different
version of the FI rate equation.

4.5.3. On the mechanism of the Fischer-Tropsch synthesis (FTS)


The intrinsic kinetic feature of the FTS is stepwise chain growth, in effect a
polymerization of -CH,- groups on the catalyst surface. This is valid regardless of the
products that are formed, paraffins, olefins or alcohols.
There has been much research on chemisorbed intermediates in the FTS, reflecting on
progress in experimental methods, work that has added to our understanding of the
synthesis. Although it may appear that the FI proceeds by simple polymerization of
methylene groups, a large number of identified or assumed species exists on the catalyst
surface during reaction, some of which are depicted in Fig. 19 (Schulz, 1985).
1. Weruler/Fuel Processing Technology 48 (19961 189-297 233

H, H, ,,OH, ,OH ,CHTR R. ,O Rt


0 c c 0 C CH-0

u L8 19 z! 21 22

Fig. 19. Some surface species of Fischer-Tropsch CO hydrogenation (Schulz, 1985).

The large number of these surface species has led to the consideration of several
mechanistic pathways for the FTS (Fischer and Tropsch, 1926; Starch et al., 1951;
Anderson, 1956, Anderson, 1984; Pichler and Schulz, 1970; Dry, 1981; Biloen and
Sachtler, 1981). However, there is general acceptance that chain growth in the FTS
proceeds by a stepwise process. Herington (1946) first introduced into FT studies the
probabilities of chain growth and chain termination, terms common to polymer chem-
istry. He considered the paraffins and olefins formed on a cobalt catalyst and postulated
that they were formed by stepwise addition of a methylene (-CH,-) entity to the
growing chain on the surface of the catalyst. Anderson (1956), Anderson (1984) and
Friedel and Anderson (1950) analyzed the product distribution of a large number of FT
runs using different fixed bed catalysts. They found that plots of 1ogWJn against the
carbon number n yielded straight lines over a fairly large range of products (W,, is the
mass fraction of a particular product). This showed that the probability of chain growth
(Y was essentially constant. Most FT mechanisms assume that the monomer unit is the
same weight wherever it is found in the chain. From this work an equation for mass
fraction can be written as
W,=n(l -*)2& (4.34)
This expression is equivalent to the Schulz-Flory equation (Flory, 1950; Schulz,
1935) which treats the Ff product distribution as a polymerization process. Eq. (4.34) is
now generally referred to as the Anderson-Schulz-Flory (ASF) equation. It is usually
written in the logarithmic form

w, (1 - cr)2
log- = n log (Y+ (4.35)
n a
234 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

It is evident that, if Eq. (4.35) for the FI product distribution holds, the calculated
value of cy from the slope (log a) should be consistent with that calculated from the
intercept, log(1 - cu>*/(u. The probability of chain growth cx is the ratio of the chain
propagation rate constant to the chain propagation plus the termination rate constants.
The ASF equation has the consequence that methane can be synthesized in 100%
selectivity; all other products have well-defined maxima in allowed selectivities as
shown in Fig. 20 (Mills, 1993).
Numerous FI studies leave no doubt that the ASF equation predicts the proportions
of methane, to gasoline, to diesel, and waxes that will be produced in FT reactions. The
highest selectivities attainable by the ITS are, in wt%, methane 100; ethylene 30;
C,-C, olefins 56; gasoline 48. The large array of products from the FTS, as in the Sasol
operation, includes a wide range of hydrocarbons mixed with some oxygenated com-
pounds, necessitating a considerable number of separation steps.
In wax production in fixed bed reactors, there is a discernible change in the slope of
(Y plots around C ,,,, giving two (Y values, with the second always higher than the first.
The result is an increase in the amount of heavy ends produced. In slurry phase
operations, there is a large reservoir of liquid phase and the two (Y values are clearly
observed in this case (Dry, 1996).
There has been an enormous amount of research on the mechanism of the FI
reaction. The earliest postulation was made by Fischer and Tropsch (1926), who
suggested that carbon deposited from CO as a surface or bulk carbide eventually formed
FT products. Pichler and Schulz (1970) postulated a mechanism whereby CO was
inserted into a metal-alkyl or metal-hydrogen bond. Biloen et al., 1979 presented
evidence for the stepwise insertion of CH, units produced from a syngas mixture.
Ponec (1984) and Rofer-DePoorter (1981) have pointed out that the FT mechanism
includes various pathways with the same intermediate leading to different products,

01 oa 0.3 0.4 05 0.6 0.7 as 03 Lo

Fig. 20. Plots of calculated selectivities (percent carbon atom basis) of carbon number product cuts as a
function of the probability of chain growth.
I. Wender / Fuel Processing Technology 48 (1996) 189-297 235

whereas different intermediates can give the same products. This is probably the best
view, although the work of Biloen et al. mentioned above and that of Brady and Pettit
(19801, Brady and Pettit (1981) support the view that the principal FT mechanism
involves a -CH,- stepwise polymerization.
Many mechanisms have been postulated for the FTS; a plausible one is given in Fig.
21 (Dry, 1996). The first step is the dissociation of CO, perhaps assisted by the
chemisorption of hydrogen. Aldehydes and alcohols result from the insertion of CO into
the chain; the latter is a chain growth termination event.
Fairly simple catalyst and process changes in the FTS can yield large increases in
oxygenated compounds of all chain lengths. In line with the chain termination steps,
alcohol selectivity with iron catalysts is directly proportional to the hydrogen partial
pressure, as is the olefin to paraffin ratio. There is a good correlation between the partial
pressure of CO, and the selectivity to acids.
Graphitic carbon forms by agglomeration of carbon atoms that are formed by CO
dissociation.
The effect of temperature on selectivity is consistent for all FT catalysts. As the
temperature of the synthesis is increased, the methane selectivity increases, the amount
of olefins in the product rises and the selectivities toward oxygenated compounds
decrease. The temperatures used at Sasol bear out these effects and take them into
account in actual operation.
Promoters have been divided into two types according to their mode of action.
Oxides that are difficult to reduce, such as SiO,, Al,O,, MgO, ThO,, La,O, and ZnO,
are called structural promoters. They furnish a large surface area and prevent recrystal-
lization and sintering of the active catalyst. There is evidence that these so-called
textural promoters often interact chemically with various oxidation states of the catalyst
and can exchange oxygen atoms.
Chemical promoters exert their influence by mechanisms not clearly understood.
They may transfer electrons to the catalysts or even block pores. Alkalis and their salts
are the chemical promoters most often used-they are basic salts and catalyst basicity,
especially for iron catalysts, is a key parameter. Promoter effects depend not only on the
type and amount of alkali salt added but also on the interaction of the alkali with the
support, with other promoters, and with impurities. Potassium oxide, K,O, is the
chemical promoter used most often. If the alkali reacts with the support, the basicity will
be decreased and more alkali must be added. Alkalis are considered to be electron
donors.
The effects of alkali promotion on FT catalysts such as iron may be summarized as
follows: (a) suppression of hydrogenation capability, (b) increase in CO dissociation, (c)
increase in formation of long-chain hydrocarbons, and (d) decrease in the conversion
activity of CO.
The sample ratto P,,/Pco adequately represents the selectivity for a fixed bed
reactor operating at a low temperature with an extruded iron catalyst (Dry, 1990). The
partial pressure of CO, and also the total pressure (Dry, 1981) influence the selectivity
for iron catalysts which operate at a higher temperature.
IT catalysts can lose activity for a number of reasons: (a) sintering (loss of active
surface area due to growth of crystals), (b) conversion of the active parts of the catalyst
236 1. Wender/Fuel Processing Technology 48 (1996) 189-297

TNlTlATION AND C COMPOUNDS


I
GRAPHITE
t +c

M M M
I M

1 H2

1
CH2
H
2

LcHq
5
M
M
CHAINGROWTH k
CH
7H3 I2
CH2 Cl-l (A)
(1) C9 - CH - .-__C)

f 5
i
M M M

R
/ 2 H2

(2) CH -=W

f W
M

1: ZHZ, RCH2CH2COOH
I
.c. RCH2CH0

Fig. 2 I. Mechanism of the IT reaction (Dry, 1990).

to inert phases (i.e. a metal to a metal oxide), (c) deposition of carbonaceous substances
on the active surface area (carbon deposition), or (d) chemical processing or poisoning
of the surface (sulfur is the chief culprit here).
Poisoning by sulfur compounds apparently occurs chiefly by adsorption of sulfur
atoms on active metal sites, destroying their catalytic activity. The amount of allowable
I. Wender/ Fuel Processing Technology 48 11996) 189-297 2.17

sulfur in the ET reaction is on the order of a few parts per billion. Coke deposition is
probably the most important mode of catalyst deactivation.
Tungsten and molybdenum catalysts that are more sulfur resistant than the usual FT
catalysts have been found (Murchison and Murdick, 198 1; Bartholomew, 1991). In
general, they have low activity and poor selectivity. An alkalized molybdenum catalyst
that is selective to low molecular weight products has been developed.
Bartholomew (1991) has provided an excellent review of recent developments in
catalysis of the FT reaction.

4.5.4. Circumventing Fischer-Tropsch chain growth kinetics. Use of zeolites


A great deal of work has been done on circumventing the ASF “constraints” but
these effects have usually been shown to be short-lived or the result of errors in analysis
or data interpretation. But interesting and somewhat effective methods have been
examined as a way of circumventing ASF chain growth kinetics. They involve intercep-
tion of FT intermediates by two general approaches: (a) a FT catalyst either supported
on a zeolite or physically admixed with a zeolite or (b) a multi-step (generally two-step)
process involving FlS followed by an upgrading step using a zeolite catalyst. Although
many zeolites have been used, most work has centered on Mobil’s ZSM-5 shape
selective catalyst.
FT catalysts usually operate in the 250-350°C temperature range, whereas zeolites
such as ZSM-5 are generally employed at somewhat higher temperatures, so that
optimum conditions for FT catalysts and zeolites differ. Reactions with combined
catalysts must be carried out at intermediate temperatures.
Haag and Huang (19791, Haag and Huang (1981) demonstrated the advantages of
using iron-potassium or cobalt in the first stage of a dual reactor arrangement with
ZSM-5 in the second conversion stage. In this way, the ZSM-5 catalyst can be operated
at about 355°C and the FT catalyst at normal FT temperatures. It is also simpler to
regenerate each catalyst, as each catalyst requires a different regeneration procedure.
Nijs et al. (1979) were the first to use zeolites as a chain-limiting catalyst in the FE.
The FTS was conducted with the usual Ru/SiO, catalyst and gave an ASF distribution
with 60% of the product above C ,2. A combination of a FT catalyst and a faujasite type,
RuNa zeolite gave a product with less than 1% above C ,2 but with a marked increase in
unwanted selectivity to methane. Ballivet-Tkatchenko and Tkatchenko ( 198 1) prepared
catalysts by thermally decomposing metal carbonyls of Fe, Co and Ru in the cavities of
zeolites; they obtained selective formation of C , -C, hydrocarbons and the usual ASF
plot.
Nijs et al. (19791, Rao and Gormley (1990), Bartholomew (1991) and others have
discussed possible explanations for the above effects. Shape selectivity seems obvious
but metal dispersion effects (Shamsi et al., 1986) and experimental occurrences such as
liquid product holdup in the support material may play a part. It is possible that selective
adsorption of heavier molecules takes place on the zeolite and apparent deviation from
ASF kinetics only occurs in about the first 24 h of reaction.
Mobil has developed a two-stage slurry FT/ZSM-5 process combining slurry-phase
FT technology with a fixed bed ZSM-5 reaction (Kuo, 1983). Based on a pilot plant
study, a conceptual design for a 27000 bpd gasoline plant was drawn up. The product
238 I. Wet&r/ Fuel Processing Technology 48 (1996) 189-297

differs from the usual FI product in having significant amounts of aromatics (Kuo,
1985).

4.6. Fischer-Tropsch Sasol plants

As mentioned earlier, the first commercial FT plant built for profit was the iron
catalyzed, medium pressure (loo- 150 psig) fixed fluidized bed reactor in the Hydrocol
Process at Brownsville, Texas. It used syngas made from natural gas and operated from
1950 to 1953.
At about this time, 1955, in South Africa, the Sasol 1 FI plant, based on the use of
iron catalysts, went into commercial operation. This country has an abundance of coal, a
dearth of petroleum resources, and large centers of population far from the ocean,
together with a political and economic need for resource autonomy. The Sasol acronym
was derived from the South African Coal, Oil and Gas Corporation, Ltd. a private
company established with government funding through the Industrial Development
Corporation to convert coal to liquid fuels and chemicals via the ITS. Sasol 1 produced
about 8000 bpd of products, supplying about 5% of South Africa’s motor fuel needs plus
other fuels and chemicals. Sasol 2, with an output of 50000 bpd, started operation in
1980; Sasol 3 went on stream in 1983. South Africa then had the capability of providing
= 40% of their liquid fuel and petrochemical needs from coal using the FT process.
An excellent account of the PI’S has recently been given by Dry (1996).
Sasol 1 initially had two types of reactors. The fixed bed tubular reactor, termed the
ARGE reactor, used a precipitated iron catalyst promoted with copper and a potassium
salt such as K2C0,. The other type was an entrained fluidized bed reactor, the Synthol
reactor, which used a fused iron catalyst with alkali and other proprietary promoters.
Sasol 2 and 3, however, use only Synthol reactors each consuming about 40000 tons of
coal per day. Sasol 2 is the largest grass roots, single-purpose processing complex in the
world, occupying an area of one mile by one and one-half miles and valued at several
billion US dollars. Sasol 3 is essentially a twin plant located adjacent to Sasol 2.
The Sasol plants are specific applications of the ITS. Coal is gasified in Lurgi
gasifiers which yield significant amounts of methane and large amounts of CO, in the
gaseous products and HJCO ratios that hover around 2. Lurgi gasifiers are generally
best used with lower rank, non-caking coals. The overall methane yield from the gasifier
and the FTS itself can reach as high as 20%. Early on, no pipelines for the distribution
of the methane existed in South Africa and the amount of methane made exceeded the
country’s demand; the Sasol plants were forced to reform the excess methane to make
more syngas, which was then recycled. As a result, the thermal efficiency at Sasol,
defined as the lower heating value (LHV) of the products divided by the LHV of all the
coal used, was only about 40% for the production of motor fuels, including gasoline,
diesel oil and LPG (Eisenlohr and Gaensslen, 1981). The efficiency could rise to about
60% if the methane could be utilized as such.
Research on the FTS decreased abruptly in the 1950s. Only South Africa had a
modest program devoted mostly to practical problems that arose from operation of Sasol
1, while a small group continued research at the US Bureau of Mines (Anderson, 1956,
Anderson, 1984).
1. Wrnder / Fuel Processing Technology 48 (19961 189-297 239

The Arab oil embargo of 1973 furnished great impetus for new research on the Ff’S
with emphasis on improving gasoline and diesel oil yields. Sasol workers had greatly
improved their Lurgi gasifier operation, effectively cleaned the gases that emerged from
the gasifier, removing CO, and almost all the sulfur compounds, and developed
excellent methods for the separation of the multitude of products obtained. They were
able to maximize gasoline and diesel production while separating out many “petroleum”
feedstocks such as ethylene and propylene and also oxygenated chemicals (ethanol,
acetic acid, acetone, etc.>. The then projected distribution of products from Sasol 2, in
tons per year, was: motor fuels 1500000; ethylene 185000; chemicals 8.5 000; tar
products 185 000; ammonia (as N) 100000; and sulfur 90000. Some 2 140000 tons per
year of saleable products were envisioned. The plants have over the more than 15 years
of operation undergone a large number of additions and expansions and trhe Secunda
plants now produce nearly 7 million tons per year of marketable fuels and chemicals
(Geertsema, 1996).
Under certain conditions, synthetic fuels from coal are viable, and improvements,
some quite striking and valuable, are continuously being made in the FTS. Much was
learned from the building of Sasol 2, so that the cost of Sasol 3 was significantly less
than that of Sasol 2.
The Sasol Fischer-Tropsch process for the synthesis of fuels and chemical is outlined
in Fig. 22 (Dry, 1996). More than 100 products are marketed by Sasol today, furnishing
a wide spectrum of fuels and chemicals. A complex refinery system is required to deal
with all the products.
Sasol, originally funded with Government money, was fully privatized in 1979 and its
shares have since been traded on the Johannesburg Stock Exchange. The only protection
that was applied was in regard to gasoline sales; this came to 4.3 cents per gallon to the
end user. It is estimated that this tariff protection saved South Africa $1.6 billion a year
in foreign exchange; it furnished Sasol the difference between $21.40 a barrel of oil and
the world price of $18.77 at the end of 1995. However, the government announced that
it would reduce the tariff protection in 1996 to provide $19 a barrel and cut it further by
mid-1999 to assure $16 a barrel. However, there is no protection for chemicals, which
must compete in the international market.
When the Sasol 2 and 3 plants began operation, synfuels provided about half of South
Africa’s needs. With growth in other areas in the last decade. Sasol now supplies only a
third of the country’s fuels.
It is informative to look at Sasol’s operating profit contributions (Table 6). The
substantial increase in profits from petrochemicals is due largely to capital projects
specifically aimed at increasing the number and variety of output of high-value
chemicals.
Sasol operations have a significant effect on the economy of South Africa. More than
a trillion (10’2> dollars is saved on its foreign exchange and a very large number of jobs
are created, directly or indirectly.

4.6. I. Sasol reactors


Sasol mines a total of about 40 million tons of coal per year. There are three major
plants, Sasol 1 at Sasolburg and Sasol 2 and 3 at Secunda. The total of 97 Lurgi gasifiers
240 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

Fig. 22. Block diagram for Sasol plant processes. (Reprinted with permission of Elsevier Science Publishers
from Dry, 1990).

Table 6
Source of operating profit at Sasol (1994-95 profit R 2805 million = $780 million)
Percentages
93-94 94-95

Synfuels 48 44
Coal 14 11
Refining, fuel gas 21 14
Petrochemicals 17 31
loo loo
I. Wender /Fuel Proce.s.sing Technology 48 (1996) 189-297 241

consumes 27 million tons of coal per year. Gasifier coproducts include ammonia, sulfur,
phenols, cresols, pitches, anode grade coke and, soon, metallurgical grade coke.
Two types of reactors were used at Sasol 1 until recently. They were the ARGE fixed
bed reactors. which are still in use, and the Synthol circulating fluidized bed (CFB)
reactors, which were shut down at Sasol 1 in 1992 but are still in full operation at Sasol
2 and 3. The fixed bed catalyst uses fine metallic iron filings as the raw material; it is a
precipitated and extruded catalyst with alkali promoters. Catalyst changes are made after
70-100 days. Low temperatures, about 225°C at 25 atm (one at 45 atm since 1984) are
used. At higher temperatures, carbon deposits on the catalyst would lead to reactor
plugging. The heat of reaction is removed by circulating water on the outside of the
tubes.
The Synthol catalyst is made from millscale from a nearby steelworks. The ground
catalyst with added promoters is then fused in an open arc furnace. The catalyst leaves
the furnace at about 15OO”C, solidifying as it cools. The ground catalyst ( = 200 mesh) is
reduced with hydrogen and stored in the absence of air.
In the Synthol reactors, which operate at about 340°C and 25 atm, flowing hot
catalyst from the standpipe is entrained by the feed gas into the reactor zone where the
heat of reaction is transferred to heating coils. The turbulent flow of gas and catalyst
passes through heat exchangers to the wide settling hopper above the standpipe; here
catalyst and gas disengage. Hydrocarbon vapors and gas leave the reactor via cyclones
to remove entrained catalyst, returning them to the settling hopper. About a third of the
generated heat is removed by internal heat exchangers; the rest leaves with the recycle
gas. Temperatures of the exit gas are = 320-360°C.
As depicted in Table 7 (Jager et al., 19901, the Synthol reactor produces more light
hydrocarbons, more olefins, more oxygenated compounds, more gasoline and less heavy
oil and waxes than the fixed bed.
Sixteen later versions of the CFB reactors were erected at Sasol 2 and 3 in 1980 and
1983 at Secunda. They had capacities some three times that of the corresponding
reactors at Sasol 1, to an equivalent of 6500 bpd.
The CFB reactor, however, has a number of limitations (Jager et al., 1990)
- It is physically complex and is suspended in a complex structure.
- Circulation of large tonnages of catalyst results in considerable recycle gas compres-
sion with added costs.

Table 7
Product selectivities of Sasol commercial reactors. (Reprinted with permission of Baker A.G. Publishers from
Jager et al., 1990)
Product Fixed bed Synthol (fluidized bed)
CH, 4 7
C, -C, oletins 4 24
C, -C, paraffins 4 6
Gasoline I8 36
Middle Distillate I9 I2
Heavy Oils and Waxes 48 9
Water Soluble Oxygenates 3 6
242 I. Wender / Furl Processing Technology 48 (1996) 189-297

?? Carbon formation increases with increase in temperature and lowering of the HZ/CO
ratio. Carbon deposition reduces the amount of iron catalyst in the reactor, limiting
reaction rates and catalyst life.
* There is considerable erosion at the various bends in the reactor.
+ No further scale-up of the Secunda CFB reactors is deemed feasible.
A development at Sasol led to the successful commissioning of a commercial scale
conventional fixed fluidized reactor as an alternative to the CFB reactor. This reactor
type is now called the SAS (Sasol Advanced Synthol) reactor. A 5 m diameter reactor
operated in Sasol 1 from 1989 to 1992. During this time, important design data were
gathered. With the revamping of the Sasol 1 facility, it was decided to dedicate it to the
production of waxes and waxy products. Therefore, this reactor was transformed into a
new type of reactor for the production of heavy hydrocarbons. A new SAS reactor has
been recently commissioned in Sasol 2, Secunda. This larger reactor, 8 m in diameter
and 38 m high, is coupled to the existing production facility. The operation during the
initial 6 months has been smooth.
In the SAS (Sasol Advanced Synthol) reactor, the gas enters the reactor via a
distributor and bubbles through the catalyst bed. It is referred to as “fixed” since the
bed, although fluidized, is not transported as in the CFB reactor.
The main benefits of the SAS reactor are
* Cost, less than half that of the older reactors.
- Better thermal efficiency.
- Pressure drop over the reactor is lower so that gas compression costs are lower.
- Isothermal behavior.
* Greater flexibility.
- Significant savings in operating and maintenance costs.
- High oil selectivities are achieved and the percentage conversion is higher than in the
CFB reactor.
- Scale-up to about a 16500 bpd SAS unit is envisaged.
A preliminary economic analysis and comparison of the SAS and CFB reactors
indicated that the SAS will be a leading contender for future synthetic fuels from
coal-generated syngas. The ability of the FTS to co-produce chemicals and to produce
fuels with low sulfur and aromatics makes this an attractive option. SAS reactors are less
than half the size of the CFB reactors of the same capacity (Jager et al., 1990).
In May 1993, Sasol brought into operation its fourth and newest type of reactor, a
first-of-its-kind commercial large-scale slurry bed reactor know as the SSBP (Sasol
Slurry Bed Process; Fourie, 1992; Geertsema, 1993). It resembles a SAS reactor except
that the catalyst is suspended in a liquid, usually a FI wax. The 5 m diameter, 22 m high
SSBP has been operating successfully since start-up. Although great potential exists to
vary product yields and selectivities by adjusting operating conditions or catalyst
properties in the SSBP, its first use is in the production of FT wax. The production of
high-quality diesel fuel is a longer term goal, although the wax can be cracked to yield
this type of liquid fuel. The SSBP has a number of advantages including a low pressure
drop, isothermal behavior, good scale-up potential, on-line catalyst removal, improved
catalyst economy and low turndown ratio. Construction of a slurry-phase reactor is about
45% cheaper than that of the same capacity multi-tubular unit (Jager et al., 1994).
I. Wen&r/ Fuel Processing Technology 48 (1996) IN-297 243

ling

Boiler
Feed
Water

Synthesis
Gas

SYNTHOL REACTOR SASOL ADVANCED FLUIDIZED REACTOR

Fig. 23. High temperature IT reactors.

Of the four Sasol commercial reactor options, two are operated at high temperature
(= 34O”C), Fig. 23, and two at low temperatures (220-27O”C), Fig. 24 and Table 8.
The high temperature process (HTFI) was first used at Sasol 1, 2 and 3 for
circulating fluidized reactors (Synthol units). An updated variant, a more classical
fluidized bed with smaller costs and larger benefits (the Sasol Advanced Synthol
process), was operated at Sasol 1 and now at Sasol 2 Secunda. It is likely that any new
commercial venture will use either the SSBP or the SAS reactor, depending on the type
of product that is desired. The high temperature process yields large amounts of olefins,
a lower boiling range and very good gasoline. Diesel fuel can be produced readily by
oligomerization of olefins. Substantial amounts of oxygenates are also produced.
The low temperature FI (LTFI’) has, until recently, been commercially operated only
in fixed bed tubular reactors using an iron catalyst or, as Shell now operates, with a
modified cobalt catalyst. This process yields much more paraffins and linear products
and can be adjusted to very high wax selectivities. The primary diesel cut and wax
cracking products can give excellent diesel fuels. The very linear primary gasoline
fraction needs further treatment to attain a good octane number. Olefin and oxygenate
levels are lower than for high temperature FT reactors.
As mentioned, Sasol has commercialized a low temperature FI process using the
concept of a slurry bed. Iron based catalysts are used in the SSBR. The single slurry bed
unit at Sasol 1 now uses all the syngas previously fed to three decommissioned Synthol
units. Synfuels production has ceased at Sasol 1. The major products from Sasol 1 are
waxes from the new slurry unit and from six tubular ARGE reactors. A range of
244 I. Weruler / Fuel Processing Technology 48 (1996) 189-297

Fresh

Tube Bundles
- Steam

TO
-

Gas Outlet
Synihesis
-km

ARGE REACTOR SLURRY BED REACTOR (SSBP)

Fig. 24. Low temperature FT reactors.

products including very high purity ethanol and also propanol, paraffin cuts, phenol,
o-cresol, ammonia and industrial gas is also produced. Typical product distributions
from three reactor types are given in Table 9.
Of the 40 million tons of coal per year used at Sasol, 65% goes to gasification and
35% to power/steam. The major products (mass %) are gasoline 38, diesel 22,
industrial gas and LPG 13, ethylene and propylene 4.4, polypropylene 2.0, alcohols and
ketones 6.3 and “black” products 5.8 (Geertsema, 1993).
Iron catalysts have been used in all Sasol reactors. However, variously promoted
forms of cobalt have been tested in pilot plants. Potentially, a cobalt catalyst should
work well in the SSBR reactor. Cobalt generally gives more saturated products (less
olefins) and also fewer oxygenated compounds.

Table 8
Sasol Fischer-Tropsch commercial reactors. CFB = circulating fluidized bed; SAS = Sasol advanced synthol;
ARGE = low temperature fixed bed reactor; SSBP = Sasol slurry bed reactor Geertsema, 1993
CFB SAS ARGE SSBP
HTFT HTFI. Lm LTFI

Current capacity 110000 - 3200 2400


Capacity per reactor 6500-7500 3400 operated, 500-700 2400
11300design
Expected scaled-up capacity per reactor - 16500 1200 (design) 10000
Max % gasoline 80 80 30 30
Max % diesel 70 70 85 85(?)
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 245

Table 9
Typical primary Fischer-Tropsch product spectra (Geertsema, 1993)
ARGE SSBR SYNTHOL = SAS

cs-Cl2 cl,-Cl, cs-Cl2 c,,-c,s c5-c,0 c,,-c,4


Paraffins 53 65 29 44 13 15
Oletins 40 28 64 50 70 60
Aromatics 0 0 0 0 5 I5
Oxygenates 7 7 7 6 12 IO
100 100 100 100 100 100
% n-Paraffins in paraffin cut 95 93 96 95 55 60

4.6.2. Petrochemical expansion at Sasol


Sasol is investing about $1 billion in plants commissioned in 1993 and planned to
1995 and beyond. Foreign exchange savings should reach about $1.6 billion per year.
Some of Sasol’s commitments are listed in Table 10.
Product streams from Sasol plants contain a wide variety of terminal linear olefins,
with C,-C, exceeding 750000 tons per year. Of this, more than 250000 tons per year
consists of 1-pentene and more than 170 000 tons per year is I-hexene. Currently these
olefins are utilized in the gasoline pool. Sasol initiated the first phase of a large olefins
project (Chemical Marketing Reporter, 1992) with a goal of a capacity of 450000 tons
per year of I pentene and 1-hexene. The Sasol plants employ a newly developed, low
cost recovery and purification process which will produce copolymer grade I-hexene
and I-pentene at a combined capacity of 100000 tons per year. The first phase was
commissioned in 1994. 1-Pentene copolymers are expected to provide interesting
advantages in improving the performance of I-butene copolymers while meeting food
grade restrictions on higher olefin copolymers. It also could be an excellent comonomer
for ultra low density polyethylenes and elastomeric polyethylenes.
In addition to the utilization of these olefins in polymer production or perhaps in their
conversion to alcohols via the hydroformylation (0x0) reaction, Sasol has a number of

Table IO
Current Sasol expansion (1994)
Proiect Commissioned $ million

Sasol I upgrade 5/ I993 215


Acrylic fibres 6/1993 120
Anode coke 6/ I993 100
Open cast mine, Sasol I 4/ I993 40
Cresylic acids 4/ 1993 27
a-Olefins 3/1994 122
Increased industrial gas + 0, 7/ I994 100
New gas pipeline 7/ I994 47
Alkylamines 9/ I994 I6
Acrylonitrile l/1995 102
SAS reactor Secunda 9/ I995 41
990
246 I. Wender/Fuel Processing Technology48 (1996) 189-297

further products under consideration. These include acetic acid and acetates recovery,
methanol synthesis, MTBE and ETBE, bisphenol A, natural gas utilization as a source of
syngas, and exploration for oil.

4.63. Natural gas to FT products


Natural gas is well known as a clean, efficient hydrocarbon source, World proven
natural gas resources amount to some 4000 trillion (4000 X lO’2) cubic feet, but about
half is found in areas far from markets and it is not economic to transport it as a gas for
long distances. The world is dependent on liquids for transportation fuels and the
conversion of remote natural gas to clean liquids is a highly desirable goal. As an
example of the possible use of this remote resource, gas conversion technology could
furnish about 100000 bpd of premium transportation fuels for some 20 years from only
six trillion cubic feet of gas reserves (Velocci, 1991).
Some driving forces for the utilization of natural gas, especially of remote gas, are (a)
long pipelines are not only costly but often simply not feasible, (b) liquified natural gas
(LNG) markets are limited, (c) there is promise in utilization via the synthesis of
methanol or a ET process, and (d) political and environmental issues are driving the
development of processes for conversion of natural gas to fuels and chemicals.
There is considerable interest in cobalt IT catalysts and a large number of patents
have been issued to major oil companies, including Gulf, Exxon, Shell and Statoil. The
cobalt catalysts under development by the various companies can be grouped as follows
(Goodwin, 1991); they generally contain about 20 wt% of cobalt (Table 11).
These companies were or are focusing on the production of heavy ends via the ITS;
the products can then be hydrocracked to yield desirable middle distillate fractions.
Although Fe, Co, Ni and Ru are the most active metals for the FTS (Vannice, 1982),
only Fe and Co are feasible catalysts. When the source of syngas is coal gasification, the
high CO/H, ratio so produced favors the use of an iron based catalyst. Iron has high
WGS activity so that less hydrogen is needed. However, there is one CO, molecule
produced for each -CH,- in the products when iron is used and this may be a
consideration in regard to the so-called greenhouse effect. On the other hand, cobalt is
not a good WGS catalyst and oxygen exits mostly as water.
Addition of small amounts of a noble metal can lower the reduction temperature
needed for IT cobalt catalysts (Raab et al., 1990). Fiato et al., 1989 have suggested that
the function of the second metal is to improve the regenerability of the catalyst. The
noble metal probably acts as a source of spillover hydrogen at lower temperatures,

Table 11
Typical catalyst constituents
Company Primary Secondary 1 Secondary 2 support
Gulf co Ru oxides alumina
Exxon co Re/Ru oxides titania
Shell co with or without noble metal zfl2 silica
Statoil co Re oxides alumina
1. Wender / Fuel Processing Technology 48 (I 996) 189-297 247

resulting in better reducibility or removal of carbonaceous material from the catalyst by


hydrogenation.
Important catalyst parameters given in the patents for improving catalytic properties
include the method of preparation, impurities in the support, support porosities, Fe or Co
dispersion, metal-support interaction and phase transformations.

4.6.4. The Mossgas gas to fuels process


In 1987, a project to produce liquid fuels from offshore natural gas in South Africa
was announced. The government-supported project, known as the Mossgas project, had
the Industrial Development Corporation as a major shareholder. The project is located in
Mossel Bay in the Eastern Cape area of South Africa (van Rensberg, 1990).
The natural gas is 72 miles offshore from Mossel Bay. An underwater pipeline was
built to transport about 168 million cubic feet of gas per day and 50 metric tons per hour
of gas condensate to the town of Mossel Bay, the offshore gas to be recovered with
separation of the associated condensate. The total was then to be converted to transporta-
tion fuels. Sasol licensed the use of three Synthol (CFB) reactors to the Mossgas project.
This first use of the Sasol Synthol process on a natural gas feed was commissioned in
1993 but the current project is now catalogued as a “commercial mistake.” The gas
field would be depleted by 1997 and the $3.3 billion that it cost would not be
recuperated. The South African government has given the go-ahead for Mossgas to
invest a further R 360 m ($100 million) in satellite gas fields. There is the possibility to
convert the facility into a fully fledged petrochemical producer. It is hoped that a buyer
for the plant would be found (Cohen and Ensor, 1996).

4.6.5. The Shell middle distillate process


The Royal Dutch/Shell companies have been carrying out research and development
since the late 1940s on converting coal and natural gas to liquid fuels. Impetus to their
work was furnished by the 1973 oil crisis. They developed the Shell Coal Gasification
process (SGP) for power generation via syngas; a 250 MW unit, operated by the Dutch
electricity generation board, began operation in 1993.
In 1989, Shell, together with its partners Petronas, the Sarawak State Government and
the Mitsubishi Corporation, announced the construction of a commercial Shell Middle
Distillate Synthesis (SMDS) plant in Bintulu, Sarawak, Malaysia to convert natural gas
to liquid transportation fuels (van der Burgt et al., 1990; Sie et al., 1991; Eilers et al.,
1990; Ansorge and Hoek, 1992; Tijm et al., 1993). The aim is to convert 100 million
cubic feet of natural gas from offshore fields to about 500000 tons per year of
hydrocarbons. Natural gas reserves are large and growing, rivaling the level of crude oil
reserves. Capital expenditure for a syngas plant based on coal is about twice that of a
plant based on natural gas.
The SMDS process involves three stages: syngas manufacture, heavy paraffin
synthesis (HPS) via the FTS, and heavy paraffin conversion (HPC). The products,
mainly kerosene, gas oil and some naphtha, are finally separated by distillation. The
high quality of the products, with no sulfur or aromatics (typical of FT products),
coupled with an anticipated growing demand for middle distillate fractions, especially in
the developing countries, would seem to make the SMDS a desirable route to environ-
248 I. Wender / Fuel Processing Technology 48 (1996) 189-297

mentally acceptable liquid transportation fuels. The SGP gasification to syngas uses a
non-catalytic autothermal partial oxidation of methane operating at 1300-1500°C at up
to 70 atm with a carbon efficiency of over 95%. Little adjustment of the desired 2:1
HJCO ratio is required.
CH, + l/20, --) 2H, + CO + - (CH,) - -t H,O (4.36)
Steam reforming of the C ,-C, FI products yields additional hydrogen.
In the next step, the heavy paraffin synthesis (HPS), the reaction mechanism follows
ASF polymerization kinetics, characterized by the probability of chain growth (Y versus
chain termination. A high (Y(Fig. 25) corresponds to a high average molecular weight,
highly linear paraffinic product. Shell has a proprietary catalyst based on cobalt,
probably plus a noble metal such as ruthenium or rhenium. The desired product is a long
chain hydrocarbon wax. The reaction is carried out in a fixed bed tubular reactor with
efficient energy recovery. The catalyst is expected to have a useful life of over a year
and is regenerable. The FT catalyst and operating conditions were selected so as to yield
a heavy product with a high cx value; the formation of light hydrocarbons is minimized.
The HPG product is fractionated and the fraction boiling above the gas oil range is
recycled to the HPC reactor. Selectivities are influenced by varying process severity or
conversion per pass. Products can be about 50% kerosene on total liquid product or, for
the gas oil mode of operation, some 60% gas oil.
The combination of chain length independent FTS of heavy paraffins and their
selective cracking is the basis of the SMDS process. Both the kerosene and the gas oil

0.85 0.90 0.95

a = PROBABILITY OF CHAIN GROVi-lH

co (cLAssICAL)
Fe (CLASSICAL)

_------a-------_---_

Fig. 25. Molecular mass distribution in raw product (Tijm et al., 1993).
I. Wender/ Fuel Processing Technology 48 (19961189-297 249

Table 12
Variation in product range and some leading properties of SMDS products (Sic et al., 1991)

roduct Gas oil mode Kerosene mode

Tops/naphtha wt% I5 25
Kerosene wt% 25 50
Gas oil wt% 60 25
Property Gas oil (diesel) Kerosene
Boiling range “C 250-360 150-250
Density kg/m3 780 750
Pour point “C -10
Cetane number 75
Smoke point mm > 50
Freezing point “C -47

meet all relevant specifications, as shown in Table 12 for typical product data (Sie et al.,
1991). The smoke point of the kerosene and the cetane number of the gas oil are
excellent; they can be used to upgrade low-quality liquids derived from thermal or
catalytic cracking reactions.
The theoretical maximum thermal efficiency for methane conversion to linear hydro-
carbons (based on lower heating values) is 78%. The SMDS plant is heat-balanced so
that no extra natural gas is used for utilities. The SMDS plant, if all goes well, will
operate at 65% thermal efficiency, which is 80% of the theoretical maximum.
An outline of the SMDS process configuration is shown in Fig. 26. Operations began
in late 1993.

HUU -1
c4

HZ
Natural @Naphlha
gas
HPC +
d SGP Distillation -------Kerosene
synlm
9= facifilies

FGasoil

SGP: Shell Gsslllcallon Process

HPS: Heavy Pmfffnprooess


HMU: Hydrogen
Manvfacfullng
Unit
HPC: Heavy ParefftnconversIon
HUG: Hydrogenatlcm
U~II
b%‘Pu: Wax ProductIon Uni1

Fig. 26. Simplified process flow scheme for SMDS.


250 I. Weruler/ Fuel Processing Techmlo~y 48 (1996) 189-297

The first commercial SMDS plant, partly aided by its desirable location in Bintulu,
Malaysia, is expected to cost about $660 million for a 12000 bpd output. It is expected
that future plants will be up to 50000 bpd, with advantages of scale-up. As is the usual
case, syngas manufacture constitutes over 50% of the total SMDS capital process cost.
Future work will be directed to lowering this cost, improving the catalyst and the design
of the synthesis reactors and general process integration. With further development
leading to larger plants, it is expected that specific capital costs for remote gas areas
could be in the range $25000-$30000 per daily barrel (Tijm et al., 1993).
4.6.6. Exxon advanced gas conversion technology
Fully aware of the advantages of natural gas and of the promise of the use of
available and remote natural gas, the Exxon Research and Engineering Company started
work in 1980 on development of an advanced technology for converting the gas to
high-quality refinery feedstock. They have developed the Exxon AGC-21 Advanced Gas
Conversion Technology, consisting of an integrated three-step sequence of fluid bed
syngas generation, a newly developed slurry-phase hydrocarbon synthesis and fixed bed
upgrading of products (Eisenberg et al., 1993, Eisenberg et al., 1994).
The syngas generation step incorporates a novel fluid bed reactor in which partial
oxidation and steam reforming take place simultaneously in one reactor containing
fluidized catalyst particles (FBSG). IT hydrocarbon synthesis (HPS) uses a proprietary
high performance catalyst system (cobalt plus small amounts of a noble metal) com-
bined with an advanced multiphase slurry reactor. The system presents significant
challenges which have been met: improved heat removal, scale-up of the multiphase
fluid reactor dynamics, maintenance of catalyst performance and separation of the liquid
wax from the catalyst slurry. The final step involves mild isomerization and upgrading
to water-white refinery feedstock, readily transported by pipeline or tanker.
As in usual FT syntheses, the AGC-21 process yields products free of sulfur,
nitrogen, vanadium, nickel, asphaltenes and polycyclic aromatics. Product options
include premium-quality diesel and jet fuel, lube oils, waxes, and chemicals.
The process provides flexibility in varying the nature of the FI product (Table 13). A
simplified flow diagram of the Exxon AGC-21 process is shown in Fig. 27 (Eisenberg et
al., 1993).
4.6.7. The Koelbel-Englehardt synthesis
The Koelbel-Englehardt (K-E) synthesis is an offshoot or variant of the IT
synthesis (Koelbel and Ralek, 1980). It is mentioned here because of its contribution to
the enhancement of a growing interest in slurry-phase FT reactors. The process is suited

Table 13
Exxon process produces a flexible petroleum product slate (Eisenberg et al., 1994)
Product Max. cat feed Max. diesel/jet
Naphtha 15 30
Diesel/jet 50 70
Cat feed 35 0
Total 100 100
1. Wet&r/Fuel Processing Technology 48 t/996) 189-297 251

Fig. 27. Exxon AGC-21 gas conversion process. Process development units: flow diagram (Eisenberg et al.,
1993).

for operation with low hydrogen, CO-rich gas, as produced from industrial gases such as
producer gas, blast furnace gas and gas from second generation coal gasifiers.
The K-E process operated in 19X- 1953 at Meerbeck, Germany in the liquid
three-phase column reactors found most suitable for this synthesis. The pilot unit had a
production of 11.5 tonne per day of synthesis products which were essentially the same
as those obtained in FT reactions. The reaction enthalpies of the ET and K-E reactions
differ only by single or multiple values of the WGS reaction.
It is worthwhile pointing out that Fischer and his co-workers were the first to use
slurry reactors in the FT synthesis (Fischer et al., 1932). In general, this type of reactor
uses finely divided catalysts suspended in a high-boiling oil (which may be a high-boil-
ing Ff fraction). Hall et al. (1952) in England, Schlesinger et al. (1951) at the US
Bureau of Mines, Sakai and Kunugi (1974) in Japan and Kuo (1983) at Mobil have all
carried out work on slurry FT reactions.
As mentioned earlier, Sasol is now operating a slurry bed FT reactor. On February
23, 1996, Sasol Limited and Haldor Tops&z of Denmark announced a technology
cooperation agreement between the two companies. The agreement calls for combined
promotion of Sasol’s slurry-phase distillate technology and Haldor Topsiie’s natural gas
conversion technology (Oil and Gas Journal, 1996). Using this slurry-phase process, it
was estimated by Sasol that a plant to produce a million barrels per day of diesel fuel
would need about 85 billion cubic meters per year of natural gas feedstock; it could be
built for $300-400 million. Combining Topsoe’s experience in natural gas conversion to
syngas with Sasol’s slurry technology would allow economic conversion of any viable
252 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

gas source worldwide to high-quality diesel fuel and potentially to petrochemicals.


Plants would be built close to the gas field and converted straight from gas to middle
distillates. Sasol envisaged that natural gas/diesel field plants could be build in modular
fashion; additional units would be added as demand increased.
Exxon has built and operated a large demonstration slurry unit in Baton Rouge,
Louisiana. A number of firms, including Exxon and Shell, have participated with the
DOE in the development of a slurry-phase IT reactor using precipitated iron catalysts at
LaPorte, Texas (Rae and Gormley, 1990).

4.6.8. Rentech, Inc.


Rentech, Inc. is a publicly traded Colorado corporation, which was operating a small,
= 300-500 bpd, Fischer-Tropsch unit in Pueblo, Colorado. The plant used an iron
based catalyst in a slurry reactor to produce naphtha, diesel fuel and wax. In March
1996, Rentech sold this plant to its Indian licensee, Donyi Polo Petrochemicals, Limited,
Bombay, India. The unit was dismantled and shipped to Calcutta, India. It will be
installed in Arunachal Pradesh, which is located in northeastern India. It will use as the
feedstock approximately 100000 cubic meters per day of presently flared gas. It should
be in operation using Rentech technology in early 1998 (Yakobson, 1996).

4.7. The Isosynthesis

The Isosynthesis is part of the more generalized reaction systems associated with the
FI process. This syngas reaction was developed during World War II by Pichler and
Ziesecke (1949). Details of the project, actually started in 1941, were kept secret
because its primary goal was the catalytic production of isobutane and isobutene, raw
materials for high octane gasoline syntheses (Pichler, 1952). The Isosynthesis has been
reviewed by Cohn (1956) and by Shah and Perrotta (1976).
Syngas is catalytically converted predominantly to branched hydrocarbons using
certain difficultly reducible oxides as catalysts. Development of the process was rapid
but its commercial use was cut off by the successful development of new catalysts for
the production of high octane gasoline from readily available petroleum.
Although both use syngas as the feed, the Isosynthesis differs from the FI’S in several
ways
* The Isosynthesis gives high yields of isoparaffins rather than normal paraffins.
* The catalysts are difficultly reducible oxides such as ThO, or ZIG, rather than
reduced transition metals.
* Isosynthesis temperatures and pressures are considerably higher than those used in
the FIX.
* Isosynthesis catalysts are not poisoned by sulfur to any great extent.
* The distribution of products in the Isosynthesis differs considerably from those
predicted by the ASF distribution (Table 14).
- Optimum pressures for the production of liquid hydrocarbons in the Isosynthesis are
between 30 and 60 MPa. At higher pressures, methane and dimethyl ether predomi-
nate.
Thorium oxide (ThO,) is a good catalyst for the Isosynthesis reaction; the catalyst
could be regenerated with air after several weeks so that it had considerable activity over
1. Wender/ Fuel Processing Technology 48 (1996) 189-297 253

Table 14
Comparison of product distribution for isosynthesis with that predicted for ASF kinetics
Product selectivity/W%

Isosynthesis ” Predicted by ASF


c, +cz 17 17
C, +C, 43 22
C 5+ 37 61
CH,OCH, 3

‘I Conditions: 45O”C, 600 atm, ThOz catalyst.

long times. Syngas is consumed in a 1.2CO/lH, ratio in a single pass. Although not
affected by sulfur compounds, the synthesis should be carried out in chrome alloy steel
or copper-lined vessels which exclude FTS catalysts (Shah and Perrotta, 1976). In
addition to ThO,, both ZrO, and, to a lesser extent, CeO, are active Isosynthesis
catalysts at 450°C and 150 atm.
At temperatures below 375°C alcohols and dimethyl ether become the main prod-
ucts; above 5OO”C, low molecular weight hydrocarbons predominate. At 450°C 25%
iso-C, compounds, chiefly isobutane, are formed, with about 16% of methane and 46%
of liquid products containing mostly branched aliphatics plus some aromatics and
naphthenes.
The activity of ThO, is greatly increased by the addition of Al,O,; the best results
were obtained with 20% of Al,O, added to the ThO, catalyst. Whereas Al,O, addition
increased the formation of isobutane, addition of ZrO, to ThO, increased the formation
of liquid products.
Very little has been reported about the Isosynthesis process since the early work by
Pichler and Ziesecke (19491, but interest in this reaction has been revived, chiefly
because of the growing demand for isobutene and other branched hydrocarbons.
Sofianos (1992) has reviewed the synthesis. The existing literature seems to reveal that
the main products of the Isosynthesis reaction, namely isobutene and isobutane, can be
obtained in sufficiently high yields only at high temperatures and pressures. The space
time yields for the Isosynthesis are less than in the case of the synthesis of higher
alcohols. The Isosynthesis reaction is not possible at low pressures as the formation of
DME, lower alcohols and isobutanol predominates under these conditions.
Isobutanol was one of the main products of Pichler and Ziesecke’s Isosynthesis
reaction; this indicates a relationship between the Isosynthesis and the higher alcohol
synthesis. Isobutanol plus other higher alcohols can be produced using a number of
catalyst systems under milder conditions with greater yields than iso-C, compounds
from the Isosynthesis. Large amounts of methanol are always present and there has
come into play a driving force to realize a direct reaction between it and isobutanol or
isobutene derived from isobutanol, to form MTBE.

4.8. Direct combustion for generation of electrici

A technology that shows great promise for development involves pressurized


oxygen-blown entrained gasification of coal to produce syngas and its use in the
254 I. Wender/Fuel Processing Technology 48 (1996) 189-297

Integrated Gasification Combined Cycle (IGCC) for the generation of electricity (Alpert,
1991). The US Department of Energy Clean Coal Technology program has been
carrying out a program in this area using federal funds plus added private sector funds.
A 100 MWe Cool Water Coal Gasification Combined Cycle Power test program,
based on Texaco Coal Gasification, was completed in 1989. A team that included the
Southern California Edison Company, Texaco Inc., The General Electric Company,
Bechtel, a Japanese consortium led by Tokyo Electric Power Company, and EPRI as
principals (ESEERCO and SOHIO Alternate Energy also were contributors) funded,
built and operated the first integrated, commercial scale, coal gasification combined
cycle power plant in the world (Spencer et al., 1982).
Cool Water’s IGCC unit achieved net system heat rates of 9000 BTU per kWh with
very low SO,r emissions of about 120 ppm and NO., emissions of about 25 ppm, with no
particulate emissions.
The Dow Chemical Company has been operating a second IGCC project at Plaquem-
ine, Louisiana with a single gasifier producing fuel sufficient for 160 MW. The plant
was originally fueled by natural gas; gasification of subbituminous coal to syngas was
started in 1987. The Dow technology exceeded its rated design capacity and acceptable
capacity by 1990. A commercial plant using this technology is planned by the Public
Service Indiana Company.
Several IGCC-type units are being operated or planned by US utilities and IGCC is
being actively pursued or seriously considered in many countries. There is a great
continuing interest in the commercialization of IGCC plants around the world and the
reasons are not hard to find. Allowable emissions from electricity generation plants are
becoming increasingly stringent and IGCC plants have superior environmental pet-for-
mance. In IGCC operations, well over 99% of the sulfur is removed and NO, emissions
are well below existing new source performance standard (NSPS) emission standards for
new plants. The H,S and NH, in the gasification products are removed to a greater
extent than is practical from combustion gas from conventional power plants. In IGCC,
sulfur and nitrogen, and also particulates, are removed before combustion, rather than
after combustion as in conventional pulverized coal combustion plants. The efficiency of
IGCC plants is greater than that achieved in pulverized coal-fired plants. With further
improvements in gas turbines, generation efficiencies using clean syngas could approach
or exceed 45-55%. This is thus a conservation project, as more electricity is produced
per ton of coal so that CO, emissions are reduced. It may be possible to capture and
sequester the CO, produced. IGCC plants can be used in small or in large utilities, as
they can be built in standardized modules. A flow diagram of an IGCC unit is shown in
Fig. 28.
There is a continuing development of improved gasifiers for IGCC. The principal
gasifiers now in use are the Texaco, Dow (DESTEC), Royal Dutch/Shell and British
Gas/Lurgi gasifiers, and coal gasification efficiency will be improved with ongoing
developmental efforts.
Another important proposal that expands the IGCC outlook is the possibility of
co-producing electricity and chemicals derived from syngas. The gasifier can operate
continuously and, if the requirement for generation of electricity falls, the syngas can be
used, for example, for the synthesis of methanol, dimethyl ether, methyl formate or
1. Wemlrr/ Fuel Processing Technology 48 11996) 189-297 255

Fig. 28. Block flow diagram of Cool Water coal gasification program (Spencer et al., 1986). The boiler
feedwater heat exchange is not shown.

some other stable organic fluid which could be stored and used for peak shaving
electricity generation. It may be feasible to synthesize methanol in a once-through
operation (avoiding the cost of recycling), remove the methanol and use the unreacted
syngas for electricity generation. The co-production of methanol and DME has been
discussed earlier in this report (Hansen and Joensen, 1991; Brown et al., 1991).
As mentioned earlier, Air Products and Chemicals, Inc. and Eastman Chemical are
constructing an advanced coal-to-methanol plant to be located at the Eastman Chemical
Company facility in Kingsport, TN. It is planned to demonstrate on a commercial scale
the production of methanol from coal-derived syngas using the Liquid Phase Methanol,
LPMEOH’” technology. Production of dimethyl ether (DME) as a mixed co-product
with methanol for use as a storable fuel is also being considered.
DME has a number of commercial uses. In a storable blend with methanol, the
mixture can be used as a peaking fuel in IGCC electric power generating facilities; there
are indications that the cost of energy so stored can be cheaper than using methanol
alone. Also, in a storable blend with methanol, DME can be used to increase the vapor
pressure of the mixture. The resulting higher volatility is expected to provide beneficial
“cold start” properties to methanol being used as a diesel engine fuel. As indicated
earlier, blends of DME and methanol can be used as a chemical feedstock or for the
synthesis of newer oxygenated fuel additives. DME is also being viewed as an
environmentally benign aerosol to replace Freon.
Utilities are in the business of generating electricity. It is conceivable that, perhaps in
cooperation with other manufacturers, utilities may eventually be able to broaden their
horizons by entering or cooperating with the fuel and chemical markets via syngas.
Routes from syngas to hydrogen, methanol, DME, ammonia, gasoline, diesel, jet fuels
256 I. Weruler/ Fuel Processing Technology 48 (1996) 189-297

and other markets are well-established. IGCC electricity-generating plants may help
satisfy the growing need for hydrogen gas, as aromatics in gasoline are restricted by the
CAAA of 1990 and its amendments. Over five million tons of H, were produced in the
United States in 1989 (McLamon and Cairns, 1989) and the demand for H, is growing
rapidly.
There is little doubt that the future of IGCC is bright throughout the world. The price
of natural gas, which can be used in IGCC generation of electricity, is projected to rise.
Coal is found throughout the world and plans for IGCC operation continue to be made.
A group of Dutch utilities (SEP) is building a 250 MW IGCC unit in the Netherlands,
scheduled to begin operation in 1993; it will use Shell gasification technology. The
process is closely integrated and promises to have a high efficiency. Coal gasifiers
generally yield syngas with high CO/H, ratios; interestingly, CO has a usable heating
value about 18% higher than that of H, (Camell, 1977).
Additional IGCC projects are planned to be built outside of the United States.
Elcogas, a newly formed company headquartered in Spain, was created in 1992 with
backing from the Commission of the European Communities THERMIE Programme.
Major European electricity suppliers from Spain, France, Portugal, Italy, Britain and
Germany will be involved in the construction and operation of this industrial plant. It
will be situated at Puertollano, Spain and will be the largest IGCC plant in the world.
Coal and petroleum coke will be gasified in this plant. Net performance is projected to
be 45% and 335 MW of electrical power will be generated. The plant is scheduled to
begin operation early in 1997.
In the US, a 260 MWe IGCC plant is planned by the Tampa Electric Company. It
will operate a Texaco pressurized, oxygen-blown, entrained-flow coal gasifier using
2000 tons of coal per day.
The management, disposal and use of by-products from IGCC power generation have
been discussed by Clarke (1991).
A decade ago, when the Cool Water project was just getting started, IGCC was
envisioned primarily as a coal-fired technology. Surprisingly enough, oil-fired IGCC
seems perhaps closer to commercialization than coal-fired IGCC. Petroleum coke, by
itself or mixed with coal, is also coming along for use in IGCC units. Oil refiners are
finding that investment in IGCC technology could be a good capital investment because
it uses low-value, high-sulfur fuels, such as bottom-of-the-barrel heavy oils, and
petroleum coke as feedstocks while generating the hydrogen (from syngas) needed for
desulfurization. IGCC also generates more than enough electricity for the production of
steam for use in distillation (Lamarre, 1994).
Italy is Europe’s largest consumer of oil for the generation of power. Most of the oil
is high in sulfur (about 3%), for world petroleum reserves are becoming heavier and
therefore higher in sulfur. A new European Community directive requires that the sulfur
content of fuel oil be limited to 1% by 1998 and 0.25% by 2003. Currently, four Italian
refineries plan to build IGCC units to begin operation between 1997 and 1999. This is a
classic case of market forces (oil-based technology) overtaking the original plans based
on coal based systems.
The largest hurdle for IGCC technology use in US refineries is competition from
natural gas, which is clean burning and could just as easily supply the electricity and
I. Weruler/ Furl Prowssing Technology 48 (1996) 189-297 257

hydrogen that coal, oil or coke gasification could supply. But IGCC remains an
attractive option because it uses the growing supply of petroleum resid and petroleum
coke for productive use. It is difficult to sign a long-term contract at a set price for
natural gas; there is little doubt that its price will rise in the coming years.
Mexico is also looking to the use of petroleum coke in IGCC units. China is
interested in IGCC for the production of power from coal; concern about air quality is a
driving force (Lamarre, 1994).
Some IGCC advantages are listed below.
- A clean environment.
- Fuel flexibility.
- High efficiency.
+ Low capital costs.
- Modularity.
* Phased construction.
- Marketable by-products.
* Co-products.
- Ability to use “dirty” fuels.
- Public acceptability.

4.9. Methanation, syngas to substitute natural gas (SNG)

The formation of methane from syngas was first reported by Sabatier and Senderens
in 1902. A nickel catalyst is used almost exclusively to convert syngas to methane,
although other transition metals may be used. The usual temperature range is 700-
1000°C. The dominant mechanism is dissociation of CO followed by rehydrogenation of
the surface carbon atoms to methane. Adsorbed oxygen is removed from the catalyst
surface as CO, by reaction with another CO molecule (Somorjai, 1981).
Natural gas is a clean fuel in growing use throughout the world. Immense amounts of
natural gas exist in remote areas of the world and new sources of methane are being
found in all parts of the world. Lay (1993) estimated the remaining recoverable natural
gas resource base of the lower 48 United States to be about 1.3 trillion (1.3 X IO’*)
cubic feet at the start of 1993. Natural gas consumption is forecast to increase steadily at
least through 2010 and gas supplies could be available at lower prices than in previous
years. Increased demand comes from increasing use in electricity generation and also in
industrial markets. Economics and environmental concerns are the main driving forces
for increased use of methane; technological progress in locating and bringing natural gas
to market are fueling these increased uses.
There is only one plant in the United States that now converts syngas to substitute
natural gas (SNG): the Great Plains Gasification facility in North Dakota. This plant was
built with a loan guarantee of about $1.8 billion from the US Department of Energy. A
consortium of four companies had contracted to purchase and distribute the SNG to their
customers. The project has been a success from a technical point of view but oil and gas
prices have not proved high enough to make the project economic.
The SNG facility was turned over to the Basin Electric Cooperative to avoid US
government losses. The facility is operated at a profit at this time; the consortium of gas
258 I. Wender / Fuel Processing Technology 48 (I 996) 189-297

companies pays about $2.75-3.00 per million BTU of SNG, and the market price is
approaching $2.00 for gas.
The Dakota Gasification Company sponsored a working paper by MITRE to perform
a techno-economic analysis of possible modifications of the Great Plains Gasification
facility (Gray et al., 1991). The purpose of this study was to find ways of using all or
part of the syngas available at the plant to produce liquid petroleum substitutes having a
market value greater than the SNG currently produced, if current marketing agreements
for SNG expire or otherwise become ineffective. Seven possible modifications were
considered but there is no interest in implementation of any of them.
The Great Plains Gasification facility is located at Beulah, North Dakota and
produces SNG from North Dakota lignite using Lurgi dry-ash gasification technology
(Sasol gave valuable assistance). The plant is a technical success and currently produces
SNG at some 120% of design capacity. About 165 million standard cubic feet (SCF) of
pipeline-quality gas are produced from = 17 000 tons of coal (lignite) per day. Current
market prices of natural gas are lower than SNG production costs but the gas purchase
contracts in effect enable the Dakota Gasification Company, owner of the facility, to
operate the plant economically at the time of writing.

4.10. Fuel cells

Fuels cells are included in this report because the primary reactants are hydrogen,
syngas and/or carbon monoxide. A review of fuel cell technology has been given by
Kinoshita and Cairns (1984). A thorough overview of fuel cell technology is available in
the Proceedings of the International Conference on Fuel Cells (Wedaa, 1994). Although
usage of these syngas components is small at present, fuel cell development is in a
growth stage with a promising future.
Fuels cells have the advantage of converting energy from a chemical reaction directly
to electricity. Although fuel cells have been investigated for well over a century, NASA,
in the 1960s furnished a new emphasis for their development for space applications.
They worked well in space but were too costly for general use on earth.
Fuel cells generate electricity without combustion. They differ from batteries, which
are energy storage devices; fuel cells are energy conversion devices. Reactants are
supplied continuously in a fuel cell. The great advantage of the fuel cell is that it is free
of &not-type thermodynamic limitations.
Some advantages of the production of electricity from fuel cells are: no moving parts,
no frictional wear, virtual absence of noise, minimal environmental pollutants, efficiency
independent of size, and modularity that allows installation in almost any area so that
power line investment and transmission losses are decreased. But there are disadvan-
tages, including reliability, overall high costs, actual efficiency, sluggish reaction
kinetics and ohmic losses (Prentice, 1984).
Recent research has focussed on overcoming the two primary impediments to the
widespread use of fuel cells: high initial cost and short operational lifetimes.
The variety of fuel cells may be classified by the type of electrolyte used. In order of
increasing temperature, they are: the polymer electrolyte fuel cell (PEFC) at 80°C the
alkaline fuel cell (AFC) at 80-9O”C, the phosphoric acid fuel cell (PAFC) at 260°C the
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 259

Y
GENERATION WITHOUT COMBUSTION
Fuel cells are like batteries that run
on fuel. Oxidation of hydrogen and carbon
i’l
Fuel (CH, )
Jntema1 refomling
3H1+ CO

monoxide at the anode releases


electrons to an external circuit, which
conducts them to the cathode, where
they combine with oxygen atoms in a
reduction reaction. The negatively
charged oxygen ions thus produced
are transported through an electrolyte
to the anode. Solid-oxide fuel cells
have a ceramic electrolyte and operate
at temperatures high enough for internal
reforming of methane fuel into the hydrogen
and carbon monoxide needed for the CathCdC
oxidation reaction.

Fig. 29. Generation of electricity without combustion: operation of a fuel cell (Douglas, 1994).

molten carbonate fuel cell (MCFC) at 650°C and the solid oxide fuel cell (SOFC) at
1000°C. All are of interest, with the PAFC, MCFC and SOFC gaining most attention.
SOFC units do not have the nuisance of using liquid electrolytes and are of
comparatively simple design, involving inexpensive ceramic or metallic components.
They have the potential of long service lives: nearly 10 years, or almost twice that
projected for cells with liquid electrolytes. EPRI, Westinghouse and other companies are
actively supporting research on SOFCs with the aim of achieving commercial viability
for dispersed electrical power generation by the year 2000 (Douglas, 1994).
The mode of generation of a fuel cell is shown in Fig. 29. Electrons are released to an
external circuit by oxidation of H, and CO at the anode. They are conducted via an
external circuit to the cathode where they combine with oxygen atoms in a reduction
reaction. The negatively charged oxygen ions are transported through an electrolyte to
the anode. SOFCs with a ceramic electrolyte operate at high temperatures (1000°C) so
that internal reforming of methane to H, and CO for the oxidation reaction can take
place (Douglas, 1994).
Challenges facing the commercialization of fuel cells have been outlined by delegates
from 23 countries who met recently at the Commonwealth Institute, London (Fourth
Grove Fuel Cell Symposium, 1995).

5. Category 3. The hydroformylation (0x0) reaction

5.1. Introduction

The hydroformylation (0x0) reaction consists of the reaction of an olefin with syngas
to form aldehydes or alcohols with one more carbon that the starting olefin. The reaction
260 I. Weruler / Fuel Processing Technology 48 (1996) 189-297

was discovered by Roelen (1938), Roelen (1943) while he was studying the effect of
added olefins on the Fischer-Tropsch reaction using a heterogeneous cobalt catalyst.
The hydroformylation reaction was first conducted in Germany as if it were a heteroge-
neously catalyzed reaction, but it was found necessary to continuously add cobalt salts to
the reactor. Roelen recognized that the reaction was catalyzed by soluble cobalt
carbonyls and that the hydroformylation is a homogeneously catalyzed reaction. It was
the first industrially important reaction catalyzed by soluble complexes; this oldest
homogeneously catalyzed reaction is still the most important of its kind. Hydroformyla-
tion of olefins forms the basis of a worldwide industry which accounts for over 7 million
tons of products per year (Papp and Baerns, 1991).
There are a number of reviews of the hydroformylation reaction (Pino et al., 1977a;
Pruett, 1981; Keim, 1983; Henrici-Olive and Olive, 1984; Whyman, 1985; Parshall and
Ittel, 1992). The chief uses of syngas are based on the use of traditional heterogeneous
catalysts: the manufacture of hydrogen gas, the synthesis of methanol and the FT
reaction. Hydroformylation is the fourth largest use of syngas, utilizing over 200 billion
SCF per year. The simplest hydroformylation reaction takes place with ethylene
Co*(CO),
CH,=CH,+CO+H2 + CH,CH,CHO (5.1)
or Rh complexes

Hydroformylation and many other reactions catalyzed by soluble transition metal


complexes are known; they yield high-value chemicals. In these syntheses, the reactants,
catalysts and products are present in the same phase, usually the liquid phase. Heteroge-
neously catalyzed reactions take place on metals usually supported on various oxides;
homogeneous transition metal catalysts involve soluble complexes at low temperatures
(< 250°C) with high selectivity. In contrast to the complex reactions that occur on
surfaces in heterogeneously catalyzed reactions, the mechanisms of homogeneously
catalyzed reactions are reasonably well understood. Such catalysts have significant
industrial applications and will find increased use in the future.
The hydroformylation of olefins with syngas is a rapid reaction catalyzed chiefly by
soluble complexes of cobalt or rhodium. The reactivity of various transition metal
catalysts, compared with cobalt, follows the general pattern
Rh > Co > Ru, Ir > Mn > Fe
103- 104 1 10-Z 1O-4 10-h

Although the hydrogenation of olefins and of carbon monoxide to methane and water
is thermodynamically more favored than hydroformylation, these reactions are sup-
pressed or diminished owing to the high selectivity of the 0x0 catalytic system.
The hydroformylation of propylene to form C, aldehydes and alcohols is the basis for
an important industrial application of the hydroformylation reaction.

CH,CH = CH, + CO + H, “zRhCH,CH,CH,CHO + CH,CH(CH,)CHO

The aldehydes can be converted to alcohols during the reaction or separately over a
hydrogenation catalyst. The synthesis of n-butyraldehyde from propylene accounts for
over 4 million tons per year of 0x0 products.
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 261

The most important application of the 0x0 reaction is in the synthesis of Z-ethyl-
hexanol, which is widely used as a plasticizer alcohol

base -H,O
2CH,CH,CH,CHO + CH,CH,CH,CH(OH)CH(CH,CH,)CH)CHO -+ (5.3)

CH,CH,CH,CH = C(CH,CH,)CHO?

CH,CH,CH,CH,CH(CH,CHs)CH,OH (5.4)

The hydroformylation reaction is applied to a large number of olefins, including


unsaturated oils, fats, polymers, terpenes, carbohydrates and even to steroids. In addition
to plasticizers, products include solvents, synthetic detergents, flavorings, perfumes, and
products for the fine chemical industry in general (Pino et al., 1977b).

5.2. Catalysis by cobalt

Cobalt metal and most cobalt salts form cobalt carbonyl, Co,(CO),, under hydro-
formylation conditions (200-300 atm, 1: 1 HJCO, and 1 IO-200°C). Catalyst concen-
trations are about 0.1 - 1% of metal to olefin. Ratios of n-butyraldehyde to isobutyralde-
lyde (n/iso ratios) of about 4: 1 or higher can be achieved. The straight-chain compound
is the most desirable product. About 2% of olefin hydrogenation occurs and some 5% of
high-boiling products are generally found.
The cobalt catalyst functions as both a double-bond isomerization and an 0x0
catalyst. Although internal olefins are thermodynamically favored over terminal olefins,
good yields of straight-chain aldehydes and alcohols can be obtained from internal
olefins with cobalt catalysts; the double bond migrates to the terminal position during
the reaction. The normal isomer is favored at lower temperatures and higher pressures of
carbon monoxide. The cobalt-catalyzed hydrofotmylation reaction was operated success-
fully for over 20 years with little technical change, but the need to improve the
cobalt-catalyzed process finally became obvious. There were apparent demands to
conduct the reaction at lower pressures (the high pressure is necessary to stabilize the
cobalt carbonyl intermediates), to obtain higher n/iso ratios. to improve catalyst
stability and recovery and to cut down on side reactions.
In the late 1950s it was found that replacement of the carbonyl groups in Co,(CO),
by organophosphines and similar ligands formed complexes of higher thermal stability.
This led to two new hydrofotmylation processes, the Shell phosphine-modified cobalt
process (Slaugh and Mullineaux, 1968; Tucci, 1970) and the Union Carbide phosphine-
modified rhodium process (Fowler et al., 1976).
In the Shell process, modified cobalt carbonyls such as HCo(CO),(PBu,) are used.
This and similar cobalt carbonyl complexes, although less active than Co,(CO), even at
180°C are much more active for hydrogenation to yield n/iso alcohol ratios of = 7: 1
(Falbe, 1970). The tributylphosphine-cobalt catalyst is more stable and can be used at
about 100 atm rather than 200-300 atm with the cobalt carbonyl system (Parshall and
Ittel, 1992). The alcohols can be obtained by distillation and the catalyst recovered.
262 I. Wender / Fuel Processing Technology 48 (1996) 189-297

5.3. Catalysis by rhodium

Although rhodium carbonyls are 103-lo4 times more active than cobalt carbonyls,
they give mostly branched aldehydes, i.e., the n/iso ratios are very low. But active
rhodium catalysts with excellent selectively to linear aldehydes (high n/iso ratios) were
obtained with PBu, or P(C,H,), ligands attached to rhodium (Pruett and Smith, 1969,
1970). Union Carbide started commercial production of these modified rhodium cata-
lysts in 1976 (Fowler et al., 1976, Brewester, 1976); n-butanol yields of about 85% can
be obtained.
A comparison of the phosphine-modified cobalt process and the phosphine-modified
rhodium process is given in Table 15.
Hydroformylation is a versatile reaction in which a -CHO or -CH,OH group is
added to the carbon chain of an olefin. It has a myriad of applications, a few of which
are given below. With large amounts of industrial olefins or if a particular chemical is
desired, there is the choice of three complementary hydroformylation routes which
compete for use. The choice, use of straight cobalt or phosphine-modified cobalt or
rhodium, depends on the particular olefin and the desired product. Phosphine-modified
rhodium is the clear choice to produce 2-ethylhexanol from propylene plus syngas. For
some applications, the latter choice is not versatile enough and the non-liganded cobalt
would be used for hydroformylation of mixed olefins of higher carbon number.
Phosphine-modified cobalt is best for the production of higher alcohols rather than
aldehydes. The choice is not always clear-cut. It is worth noting that many hydroformy-
lation plants worldwide still use the original cobalt catalyst with no added ligands, but
the use of rhodium complexes as hydroformylation catalysts is growing rapidly.
Because this technology is so versatile and has such different applications, a few
more examples of specific commercial hydroformylation uses will be given. Vitamin A
is produced on a relatively large scale and competition is keen in this area. Both
Hoffmann-La Roche and BASF employ hydroformylation as a key step in the commer-
cial production of vitamin A (Parshall and Ittel, 1992). A commercial route to butanediol
based upon the hydroformylation of ally1 alcohol has been developed. Shell manufac-
tures a number of unsaturated aldehydes based on the selective hydrofotmylation of
unsaturated intermediates derived from their metathesis plant (Parshall and Ittel, 1992).
The products are used in perfumes and in other commercial products. Ajinomoto
manufactured monosodium glutamate by the hydroformylation of acrylonitrile for a

Table 15
Comparison of phosphine-modified cobalt and rhodium processes
Catalyst precursor Co,(CO), + n-Bu,P HRh(COXPPh 3)3 + PPh 3
Phosphine/metal ratio 2:1 50-loo:1
PIEssure 50-100 abn l-25 atm
Temperature 160-200°c 60- 120°C
Catalyst concentration (% metal/olefm) 0.6 0.01-0.1
n/is0 ratio 7:1 8:1-16:l
Olefin hydrogenation 10% 5-10%
High boiling products 1% low
I. Wet&r/ Fuel Processing Technology 48 (1996) 189-297 263

decade ending in 1973 (Yoshida, 1978). Shell carried out the hydroformylation of
various internal olefins to convert them to terminal alcohols There are numerous
examples of the use of the hydroformylation reaction.
There are continuing advances in hydrofotmylation catalysis and the area is still the
subject of active investigation. Ruhrchemie/Rh&te-Poulenc has commercialized a
propylene hydroformylation process using a trisulfonated triphenylphosphine rhodium
ligand (Kuntz, 19871. This complex is soluble in water but insoluble in the organic
phase, so that essentially no losses of rhodium occur to the organic phase. With
water-soluble, ligand-modified rhodium catalysts, products remain in the organic phase
and modified catalysts are immobilized in the water phase. The metal catalyst is
prevented from leaching into the process products. The catalysts are separated by
decantation from unreacted olefins and products, greatly simplifying the process as
compared with other 0x0 process routes. This variant of the 0x0 process is in
commercial use (Comils and Weibus, 1995) and has great potential for extension to
other carbonylation reactions.
It is necessary, of course, to recover the rhodium essentially completely in hydro-
formylation reactions. Even cobalt should be recovered.

5.4. Chiral catalysis

An sp3 carbon atom with four different groups attached to it can exist in two
enantiomeric forms. Each enantiomer is a mirror image of the other. Most biologically
active compounds contain at least one optically active carbon; sugars and proteins in
living organisms are found in only one of the possible optically active forms. The
addition of chiral ligands to cobalt or rhodium catalysts in the hydroformylation reaction
can yield products in which one enantiomer exists in excess (Consiglio et al., 1973;
Watanabe et al., 1979; Botteghi et al., 1980). This is an exciting field of research that
promises to grow in importance because of its environmental and conservation benefits.
As examples, most pesticides are optically active but often only one of the emu-
tiomers is biologically active. We can synthesize the active enantiomer and spray fields
with half the amount of pesticide now in use. The inactive enantiomer at present has no
function but to contaminate the area. The drug thalidomide, whose use was found to be
catastrophic, exists as a racemate, an equal mixture of two enantiomers. Only one of the
two enantiomers is toxic; the other appears to be useful in other safe treatments.
These optically active products are usually made using selective homogeneous
catalysts, although heterogeneous catalysts are under development. Chiral syntheses can
be used to manufacture high-value pharmaceuticals, herbicides, insecticides, fungicides,
etc. A well-known example, not made via the hydroformylation reaction, is the
manufacture of L-dopa, which is used in the treatment of Parkinson’s disease, via a
homogeneous catalytic asymmetric hydrogenation using a chiral rhodium complex
(Knowles, 1983).

5.5. Mechanism of the hydroformylation reaction

Although there are differences in the mechanistic routes involved in the cobalt,
cobalt-phosphine and rhodium-phosphine processes, the steps in the generally
264 1. Wender/Fuel Processing Technology 48 (1996) 189-297

“accepted” reaction routes are similar. It is important that the first step involves the
removal of CO from the added catalyst complex to give a coordinatively unsaturated
species (16-electron species). The olefin then adds to the vacated orbital, followed by
insertion of the coordinated olefin into the Co-H or Rh-H bond (Orchin and Rupilius,
1972).

cc+ (CO), + H, + ZHCo(CO)~

HCo(CO), + HCo(CO), + CO

R’CH=CH, + HCo(CO), + RICH+ CH,


HCo(C0) 3

R’CH? CH, e RCo(CO), +m


T-- RCo(CO),
HCo(CO),

RCo(CO), P RCOCo(CO),

RCOCo(CO), wcoco(co),
-

HZ
RCHO + HCo(CO),
RCOCo(CO),
I.--
__ RCHO + Co,(CO),
HCo(CO),

R = R’CH,CH, or R’CHCH,

The aldehydes can be reduced to alcohols in the same system or they may be
hydrogenated to alcohols in a second step.
Similar mechanistic sequences can be written for the rhodium-catalyzed hydroformy-
lation of olefins.

6. Category 4. The Mobil methanol to gasoline (MTG) and related processes

6.1. Introduction: zeolites

Although invented by scientists at Mobil over 15 years ago, the Mobil MTG process
is still considered a breakthrough in catalytic science (Meisel, 1981; Chang, 1983; Chen
et al., 1989; Tabak and Yurchak, 1990). It has been hailed as the first new route for the
production of gasoline in over 40 years. The process is based on the conversion of
syngas to methanol or dimethyl ether (formed by dehydration of two molecules of
methanol) over a synthetic pentasil zeolite family, ZSM-5, invented by Mobil in the
1960s (Argauer and Landolt, 1972).
ZSM-5, the key element in the MTG process, has a unique pore structure and unusual
catalytic properties. It has two sets of intersecting channels with openings of = 6 A.
One set consists of elliptical lo-membered ring straight channels; the other is a set of
tortuous sinusoidal channels (Fig. 30). The channel opening is just large enough to allow
I. Wendrr/ Fuel Processing Technology 48 (1996) 189-297 265

Fig. 30. Structure of ZSM-5.

molecules with 10 or fewer carbons to wend their way through the crystalline structure.
If the molecule is small, it moves rapidly through the zeolite and experiences less
cracking. Molecules containing more than 10 carbon atoms are unable to leave the
catalyst pores and eventually reform until they are of the right size to emerge from the
catalyst.
A word about zeolites, which have become possibly the most useful catalysts in
chemical processing with a rapidly growing potential for new applications (Chen et al.,
1989, Haag et al., 1987). Zeolites are crystalline, three-dimensional aluminosilicates in
which the building blocks are silica and alumina tetrahedra. The framework will have a
net negative charge because every oxygen in the “infinite” lattice is shared by two
tetrahedra, with silicon being tetravalent but aluminum only trivalent. The negative
charge is balanced by exchangeable cations: a general structure may be written as
(Chang, 1983)

M,>,,(AlO,1,WO, >y. x&O


where n is the charge on the cation and x is the water of hydration. The SiOJAlO,
ratio varies in different zeolites; it is high in ZSM-5. The mouth of a particular zeolite
channel consists of a ring of a fixed number of tetrahedra, which determines the
diameter of the mouth or the pore size.
Over 30 natural zeolites are known and many more than that number have been
synthesized in the laboratory. Although the central atoms of all natural zeolites are
dominated by Si and Al, chemically related atoms such as B, P, Ge, Ga, etc. can be
incorporated into zeolites.

6.2. The MTG reaction

Molecular shape selectivity in these aluminosilicates was first reported by Weisz and
Frillette (1960). Weisz et al. (1962) defined two types of shape selectivity: reactant
selectivity, where certain molecules enter into the zeolite and others are excluded by
virtue of their shape and size, and product selectivity, where some of the products
formed within the pores are too bulky to diffuse out. These latter molecules are either
266 I. Wender / Fuel Processing Technology 48 (1996) 189-297

converted to smaller molecules or eventually deactivate the catalyst by blocking the


pores. Csicsery (1976) has reviewed the modes of action of shape selective catalysts.
The Mobil (MTG) process may be represented by the following equations
2CH,OH e CH,OCH, + H,O (6.1)
-H,O
CH ,OCH s + C 2 - C s olefins (6.2)

C, - C, olefins + paraffins, cycloparaffins, aromatics (6.3)


Selectivity to gasoline-range hydrocarbons is more than 85% with virtually no
compounds heavier than C,, being formed. Liquefied petroleum gas (LPG) makes up
most of the remainder of the product.
One hundred tons of methanol is converted to nearly 44 tons of hydrocarbons and 56
tons of water. The hydrocarbons produced contain 95% of the energy in the methanol
feedstock; the exothermic heat of reaction contains the remaining 5% of energy.
The first (and only) commercial MTG plant came on stream in 1985 in New Zealand,
originally producing 14500 bpd of gasoline from natural gas.
Approximate distributions of products obtained from Sasol’s fixed bed ARGE and
fluid bed (Synthol) reactors are compared with those obtained from Mobil’s MTG
process in Table 16 (Dry, 1981; Mills, 1977; Roper, 1983).
A normalized distribution of aromatics obtained in the MTG process is, in wt%:
benzene 4.1, toluene 25.6, ethylbenzene 1.9, o-xylene 9.0, m-xylene 22.8, p-xylene
10.8, trimethyl substituted benzenes 14.1, other aromatics 12.4 (Walker, 1985).
Essentially three steps are involved in the MTG process: conversion of methanol to
dimethyl ether (DME), initial formation of a carbon-carbon bond, and finally aromatiza-
tion with hydrogen transfer. Crude methanol containing 17% water may be used in this
process.
Commercial development of the MTG process is carried out in a fixed bed reactor in
New Zealand. Chang (1983) has reviewed the chemistry and operating characteristics of
the process. In 1984, a 100 barrel per day demonstration plant was built in Wesseling,

Table 16
Product distributions from Sasol (FT) and Mobil (MT@ processes
ARGE (fixed bed) FT Synthol (fluid bed) m Mobil (fixed bed) MTG
Temperature/K 490-520 633-685 63-685
Pressure/atm 26 22 14-24
Feed 1.7H,:lCO 3H,:lCO CH,OH
Product distribution
Light gas C, -C, 11.0 20.1 1.3
LPG c,-c, 11.0 23.0 17.8
Gasoline C,-C,, 25.4 39.0 80.9
c,3-Cl9 14.0 5.0 0
Heavy oil, C ,9+ 37.0 6.0 0
Oxygenated compounds 2.3 7.0 0
Aromatics, % of gasoline 0 5 38.6
I. Wrnder / Furl Procr.wn~ Technology 48 (1996) 189-297 267

near Bonn, Germany (Tabak and Yurchak, 19901. The major concern in both types of
reactor is in the control and dissipation of heat generated by the exothermic reactions
involved.
A process flow diagram of the New Zealand MTG plant is shown in Fig. 31 (Tabak
and Yurchak, 1990). Methanol is first dehydrated to an equilibrium mixture of
methanol/DME/water which releases about 20% of the overall heat of reaction. The
mixture is then fed into reactors containing the ZSM-5 catalyst to convert both the
methanol and the DME to gasoline and water. The raw gasoline is distilled to reduce the
content of durene (1,2,4,5tetramethylbenzene) to a satisfactory level. The gasoline
produced is of high quality with a RON (research octane number) of 92.0-92.5.
The mechanism of the conversion of methanol and DME to gasoline is still
controversial and is discussed elsewhere (Chang, 1988; Hutchings and Hunter, 1990).
The 100 barrel per day second generation MTG fluid bed reactor built and operated
in Germany gave higher yields of gasoline than the fixed bed MTG reactor. With
improved energy utilization and advantages in yield and quality of the products, there is
a cost advantage of at least 10% for the fluid bed semiworks (Mills, 1993). This
successful demonstration of the MTG fluid bed operation is available for commercializa-
tion but, with the availability of cheap petroleum, there is no interest in further pursuit of
the MTG process at this time.
The MTG plant was the result of an innovative discovery and concept. However, the
cost of producing high-quality gasoline from natural gas via syngas has proved too
expensive for the New Zealand government, which originally planned to have an
“indigenous” source of gasoline and to lower their balance of payments in this way.
The MTG plant in New Zealand has been sold to a private company which sells both
methanol and high octane gasoline, the relative amounts changing with market prices
and demand.
Compared with related processes, Sasol would seem to be at a disadvantage for
several reasons: coal is the syngas source, a significant amount of methane produced has
to be reconverted to syngas, and the IT product consists of a multiplicity of hydrocar-
bons and oxygenates. Sasol built new and improved reactors and developed sophisti-
cated and complicated processes to separate the daunting mixture of products so as to
market everything possible. In 1994-95, fuels accounted for 44% of Sasol’s production
with chemicals accounting for another 31% (mining, fertilizers and explosives constitute

Pro&Xx to
Trertmonr
Fig. 3 I. Schematic of New Zealand gas to gasoline complex.
268 I. Wender / Furl Processing Technology 48 I 1996) 189-297

the remainder). Sale of FT waxes has turned out to be a profitable business. Sasol has
not hesitated to acquire technology from abroad and competes on the world market. The
MTG process in New Zealand produces only methanol and gasoline; the methanol is
traded on the world market.

6.3. MTE, MT0 and MOGD reactions

Mobil has also developed several processes based on methanol as a source of


ethylene (MTE), of olefins (MTO) and of gasoline plus diesel (MOGD) (Tabak and
Yurchak, 1990). These are promising catalytic routes to valuable products from methanol.
By reducing the zeolite acidity and raising the operating temperature to > 5OO”C,Mobil
succeeded in modifying the rate of olefin formation so that over 80% of the product
consists of C,-C, olefins. This is known as the MT0 process. Oligomerization,
disproportionation and aromatization of the C 2-C l0 olefins formed in the MT0 process
are the basis for Mobil’s MOGD (olefin to gasoline and distillate) process (Garwood,
1983). Premium-quality distillate and low-pour-point stable jet fuels that meet all
commercial and military specifications are produced. These processes have not come
into commercial practice as yet but they have considerable promise for the future.

6.4. Topside integrated gasoline synthesis (TIGAS)

The TIGAS process, developed by Haldor Topsae A/S, has the unique feature of
integrating the methanol synthesis with Mobil’s MTG process into a single loop without
isolation of methanol as an intermediate. The process combines steam reforming and
autothermal reforming for the production of syngas (Topp-Jorgensen and Rostrup-Niel-
sen, 1986, Topp-Jorgensen, 1988). The MTG process takes place in three sequential
steps: production of syngas, synthesis of methanol, and conversion of methanol to
hydrocarbons over a ZSM-5 catalyst. However, these syntheses are preferably carried
out at different pressures. Syngas is typically manufactured at 15-20 atm; the methanol
loop normally operates at 50-100 atm and the MTG fixed bed process is carried out at
15-25 atm. The TIGAS process aims to modify operating catalysts and conditions so
that the pressures level out. This process, as shown in Fig. 32, combines steam
reforming with autothermal reforming using oxygen in a secondary reforming step for
syngas production. This allows syngas production at a pressure level near that of the
methanol synthesis.

Recycle Gas
7

Natural GassA
Oxygenate Gasoline
Steam * sWG??s & -
Production Synthesis Synthesis

Fig. 32. Gasoline from natural gas with the TEAS process (Topp-Jorgensen, 1988).
I. Wender/ Fuel Processiq Technology 48 (19961 189-297 269

TIGAS uses a multifunctional catalyst based on a combination of the synthesis of


methanol, which is an equilibrium limited reaction, and the dehydration of methanol to
DME, which is an equilibrium unlimited reaction (Lee et al., 1992). The methanol
synthesis takes place over a copper-containing catalyst and the dehydration takes place
over an a-A1203 catalyst. The use of a combined methanol/DME synthesis in the
TIGAS process using a multifunctional catalyst system helps to allow a low operating
pressure for the integrated process. A relatively large amount of hydrogen in the
gasoline synthesis significantly reduces the olefin content of the gasoline product. A
semi-industrial pilot plant using the TIGAS process was built and operated in Houston,
Texas. The capacity of the plant was 400 Nm3 hh ‘, corresponding to a production of
MTPD of gasoline. A process unit was operated for IO 000 hours during the l984- 1987
period. The main advantages of the TIGAS process are process flexibility and lower
investment (Mills, 1993).

7. Category 5. Methanol plus syngas or CO for synthesis of chemicals

7. I. Introduction

A classification of syngas chemistry was given earlier. Category 1 dealt with syngas
as a source of hydrogen, which can be considered the largest use of syngas. Category 2
dealt with the direct conversion of syngas to fuels and chemicals; this encompassed the
direct use of syngas for conversion to methanol, in the Fischer-Tropsch reaction, in the
direct combustion of syngas for the generation of heat and electricity, in the synthesis of
methane, and in the direct synthesis of higher alcohols.
The fourth largest use of syngas after hydrogen, methanol, and FT products is in the
hydroformylation (0x0) reaction, the conversion of olefins and syngas to aldehydes and
alcohols containing one more -CHO or -CH,OH group than the starting olefins. This
use of syngas is dealt with early because of its importance and because it does not fit
well into other categories.
Category 4 dealt with the conversion of methanol to fuels and chemicals, including
the Mobil MTG, MTE, MTO, and MOGD reactions.
It seems surprising that there is no industrially practiced direct conversion of syngas
to a major chemical, hence there is no category designation for chemicals made directly
from syngas.
This absence of such a category is not due to a lack of efforts to synthesize major
chemicals, such as ethylene glycol. acetic acid, acetic anhydride, acetaldehyde and
ethanol, among others, directly from syngas. The theoretical reaction stoichiometry for
the conversion of some of these chemicals and the Hz/CO ratio of the syngas required
are listed in Table 17.

7.2. On the direct synthesis of chemicals from syngas

It would seem highly desirable to synthesize chemicals directly from syngas, espe-
cially ethylene glycol, acetic acid and acetic anhydride. The chemical industry is
270 I. Wender / Fuel Processing Technology 48 (1996) 189-297

Table 17
Theoretical reaction stoichiometry for some important chemicals if made directly from syngas (Aquil6 et al.,
1983) a
Chemical Reaction stoichiometry H, /CO ratio in syngas lb feed per lb product
acetaldehyde 3H, +2CO+CH,CHO+H,O 1.5 1.4
acetic acid 2H 2 + 2C0 --f CH ,COOH I 1
acetic anhydride CH ,OAc + CO + Ac,O 0 I
ethylene glycol 3H, +2CO+ HOCH,CH,OH 0 1.5
ethanol 4H, +2CO+CH,CH,OH+H,O 2 1.4
ethylene 4H, +2CO+H,C=CH, 2 2.3

AC is CH,CO. a Reprinted by permission from Hydrocarbon Processin g, March 1983, p. 57, copyright 1983
by Gulf Publishing Co., all rights reserved.

increasingly being subjected to pressures to minimize or eliminate waste in the manufac-


ture of products (Sheldon, 1992, Sheldon, 1993). Although efforts to synthesize several
chemicals directly from syngas have not been successful as yet, their direct synthesis is
an extremely worthwhile goal.
It is not well perceived that the main source of waste (defined as everything but the
desired product) in the synthesis of organic chemicals is inorganic salts usually formed
in acid-base neutralizations, i.e. sodium chloride, sodium sulfate, ammonium sulfate
and other salts.
We shall see later that some direct carbonylation reactions, including the catalytic
carbonylation of methanol to acetic acid, the carbonylation of propyne (methylacetylene)
in the manufacture of methyl methacrylate and the synthesis of the drug ibuprofen via a
catalytic carbonylation, are elegant examples of processes with high atom utilization
(Sheldon, 1994).
It is revealing to mention, as examples, efforts made by the Japanese to synthesize
four of these chemicals directly from syngas; similar efforts have been made in
laboratories throughout the world with the same goals. The Japanese C-l Chemistry
National Project started in December 1980 and was officially concluded in July 1988
(Nakamura, 1990). Fourteen chemical companies, one private research institute and the
National Chemical Laboratory were involved in the pursuit of the synthesis of four
chemicals directly from syngas: acetic acid, ethylene glycol, ethanol and ethylene
(Research Association for Cl Chemistry, 1989). After a thorough search for catalysts,
bench-scale units were developed and pilot tests were designed but were subsequently
cancelled. None of the direct syngas routes to these chemicals is now used industrially,
although engineering data are available for the resumption of efforts toward commercial-
ization.
The Union Carbide Corporation patented a process for the direct production of C,
oxygenated compounds from syngas using rhodium catalysts (Bhasin, 1975, Bhasin et
al., 1978). However, these and other researchers in several laboratories found that acetic
acid could not be produced selectively from syngas at the desired rate with rhodium
catalysts or modified rhodium catalysts (Ichikawa, 1982; Watson and Somorjai, 1981;
Orita et al., 1984; Arakawa, 1984; van der Lee et al., 1986). Other Group VIII catalysts
I. Wender/Fuel Processing Technology 48 (19961189-297 271

such as cobalt, ruthenium or palladium gave poor conversions and selectivities for acetic
acid.
The direct synthesis of ethylene glycol from syngas is highly desirable, with
seemingly complete conversion of feed to product. The Union Carbide Corporation
pioneered work in this synthesis, discovering a homogeneous rhodium catalyst that
seemed effective for this reaction (Dombeck, 1983). However, catalyst activity was low,
the required syngas pressure was too high and selectivity was too low. Further work
showed that bulky alkylphosphine-modified rhodium catalysts improved the results but
conversions and selectivity were too low under the severe reaction conditions. Detailed
engineering of a three ton per day pilot plant for the direct conversion of syngas to
ethylene glycol was completed (Nakamura, 1990). But ethylene glycol is today not made
directly from syngas, although potentially promisin g routes to the glycol involving the
use of syngas are given in Fig. 33.
Attempts to synthesize ethanol directly from syngas were also uneconomical but
yielded data for future work (Pruett and Walker, 1974; Rathke and Feder, 1978;
Williamson and Kobylinski, 1979; Kiennemann et al., 1987).
The reaction of methanol with syngas to yield ethanol (the homologation of methanol)
will be discussed later.

7.3. Scope of methanol-syngas chemistry

Although the introduction of a one-step synthesis of chemicals from syngas is an


attractive concept, it has been demonstrated that many industrial chemicals are produced

Ml3TllANOl.

Cfl3C0011
Methyl formate
ffCOOCHs

Fig. 33. Methanol based route to chemicals via syngas.


272 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

more economically by multi-step (sequential) reactions; each step is carried out at


conditions that are favorable thermodynamically and kinetically. But limitations of the
selectivity or activity of direct routes to chemicals from syngas have been circumvented
by the use of methanol as a feedstock. A look at the special nature of methanol is in
order (Wender, 1984). It is made commercially and economically from syngas in over
99% selectivity. It is especially susceptible to nucleophilic attack in the presence of
suitable acids because of the lack of steric hindrance on a one-carbon alcohol; methanol
is over 100 times more reactive than ethanol in this type of reaction.
The carbon-oxygen bond in methanol is the strongest C-O bond of any alcohol, with
only ethanol having a comparable bond energy. Methanol and transition metal com-
plexes react with HI or I, to form CH,I as a reactive intermediate in many syntheses.
Methyl-metal bonds are readily formed with transition metal complexes. The CH,
group readily undergoes a migratory insertion reaction to form acyl-metal (CH,CO-M)
bonds.
Methyl alcohol is a stronger acid (loses a proton more easily) than any other alcohol.
The methoxide ion CH,O- is strong enough to react with the weakly basic CO to form
methyl formate, which has a chemistry of its own.
There are excellent reviews of the industrial synthesis of major chemicals from
methanol (Parshall and Ittel, 1992; Agreda and Zoeller, 1993; Fahey, 1987; Keim, 1983;
Papp and Baems, 1991; Herman, 1984).
Some examples of products potentially synthesized from either methanol or formal-
dehyde are shown in Fig. 33. Formaldehyde was the main chemical derived from
methanol, although the amount of methanol devoted to the synthesis of methyl t-butyl
ether has surpassed formaldehyde use. The reactions of methanol may be divided into
simple carbonylation (methanol plus CO>, reductive carbonylation (methanol plus
syngas) or oxidative carbonylation (methanol plus CO and oxygen). Syntheses from
formaldehyde are divided similarly: carbonylation (formaldehyde plus CO plus either
H,O or ROH) or reductive carbonylation (formaldehyde + syngas).
A number of chemicals, including acetic acid, methyl acetate, acetic anhydride and
vinyl acetate, are of great commercial importance and will be discussed here. They all
contain the acetyl group (CH,CO, often designated as AC).
7.3.1. Formaldehyde
Formaldehye (HCHO) production, worldwide, has historically been the largest single
application of methanol. Methanol constitutes 35.3 wt% of MTBE and its use in the
manufacture of this and other methyl ethers as octane enhancers and fuels became, in
1944, the chief outlet for methanol.
Formaldehyde is used in a large variety of resins including urea-formaldehyde
polymers, in addition to its use in the manufacture of pentaerythritol, hexamethylenete-
tramine and other chemicals (Calkins, 1984).
Formaldehyde is manufactured commercially by air oxidation, mostly by oxidation of
methanol over a silver catalyst at atmospheric pressure; a 99% conversion of methanol is
achieved with a molybdate catalyst at about 400°C and one atmosphere. A selectivity of
about 95% is achieved at a 99% conversion of methanol (Machiels et al., 1984).
CH,OH + l/20, + HCHO + H,O (7.1)
I. Wender/ Furl Processing Technology 48 (1996) 189-297 273

7.3.2. Acetic acid


Acetic acid has been produced in large quantities for more than 100 years. At present,
more than five million tons per year worldwide are manufactured (Sheldon, 1994). Its
main uses are in the manufacture of cellulose acetate, vinyl acetate monomer, acetic
anhydride, a number of acetates, pharmaceuticals, dyes, pesticides, the dimethyl ester of
terephthalic acid, as a solvent in xylene oxidation, and in other applications.
Changes in the various methods of manufacture of acetic acid reflect underlying
trends in the chemical industry toward cheaper, more available feedstocks, more energy
efficient processes and the discovery of new, more active catalysts. Excellent reviews of
the synthesis of acetic acid are available (Whyman, 1985; Papp and Baems, 1991;
Parshall and Ittel, 1992; Agreda and Zoeller, 1993). In some countries, small amounts of
acetic acid are still obtained by fermentation of ethanol or by the distillation of wood.
The first large synthetic process for manufacture of acetic acid (AcOH, where AC is
CH,CO-) was based on the reaction of the expensive and energy-rich compound
acetylene to acetaldehyde followed by oxidation (Eq. (7.2)) (Aquilo et al., 1983).

HC = CH + H,OHz+CH,COOH (7.2)

There are now three major processes for the production of acetic acid: (a) ethylene
oxidation via the Wacker process (Eq. (7.3)) (Jira et al., 1976); (b) the oxidation of
butane or naphtha to acetic acid in 30-60% yields (Parshall and Ittel, 1992) (Eq. (7.4));
and the carbonylation of methanol to acetic acid (Eqs. (7.5) and (7.6)).

CH, = CH, + [0] ?CH,COOH (7.3)

Butane or naphtha + [0] + mixed acids (7.4)

CH,OH + CO$CH,COOH (7.5)

CH,OH + COzCH,COOH (7.6)

BASF discovered and commercialized the cobalt iodide-catalyzed carbonylation of


methanol in the mid-1960s (Reppe et al., 1961; Hohenschutz et al., 1966). However, the
cobalt-catalyzed carbonylation of methanol to acetic acid operated under severe condi-
tions of temperature and pressure and it was quickly replaced by a rhodium iodide
catalyst system (Paulik and Roth, 1968). This process, known as the Monsanto process,
operated under milder conditions with greater selectivity than the cobalt based process.
Rh is recovered in over 99% yield. A comparison of the rhodium and cobalt systems is
given in Table 18.
The Monsanto process (Roth et al., 1971; Paulik, 1973) can produce acetic acid at 1
atm, but the commercial synthesis is carried out at 30-40 atm and 180°C with Rh as
catalyst and HI ( = 10-l M) as promoter. The catalytic system is very corrosive,
requiring expensive steels for construction; complete recycle of Rh and HI is necessary.
Iodine is used to convert methanol to the more electrophilic methyl iodide. A catalytic
cycle for the rhodium-catalyzed carbonylation of methanol is shown in Fig. 34 and a
214 1. Wender / Fuel Processing Technology 48 (1996) 189-297

Table 18
Comparison of rhodium- and cobalt-catalyzed syntheses of acetic acid
Rhodium Cobalt
Metal concentration = IO-’ M = lo- ’M
Temperature = 180°C = 220°C
Pressure 30-40 atm 700 atm
Selectivity (on CH,OH) 0.99 0.9
By-products 0 CH,, CH,CHO, C,H,OH, CO,
C, H=,CO,H, acetates, 2-ethylbutanol

conceptual flowsheet for the carbonylation is given in Fig. 35. The reaction is zero order
with respect to methanol and carbon monoxide and first order in methyl iodide (Roth et
al., 1971). The active species has been identified as cis-Rh(CO),I; by infrared analysis.
Large excesses of water are used, supposedly to suppress formation of methyl acetate. It
has been found that the addition of lithium iodide allows for a great reduction in the
amount of water required while maintaining high rates (Zoeller, 1993).
At present, about half of acetic acid production worldwide uses the Monsanto
rhodium-iodide catalyst. This route to acetic acid has very favorable economics and
essentially all new plants will use the Rh-catalyzed methanol carbonylation. Indeed,
carbonylation of methanol to acetic acid is one of the most successful chemical
processes that is based solely on the use of syngas. At present, 59% of worldwide acetic
acid output is made by this process.
In 1986, Monsanto sold the rights to the methanol carbonylation process for acetic
acid production to BP Chemicals (Howard et al., 1993). In 1996, BP introduced a new
catalyst system based on the use of iridium acetate enhanced with various promoters. It
is claimed that the iridium-based route is more efficient, cutting energy and purification
costs compared with the rhodium-catalyzed methanol carbonylation (Chem. Eng. News,
1996). It is of interest that earlier work showed that iridium is an active catalyst for
methanol carbonylation but is not as selective as rhodium.
There have been attempts to use a cheaper nickel based catalytic system for the
carbonylation of methanol (Rizkalla, 1987). This is still a possible catalytic route but the
hazards of nickel tetracarbonyl and the complex nature of the nickel catalyst, in addition

H.0

Fig. 34. Mechanism for the rhodium-catalyzed methanol carbonylation (Zoeller, 1993).
I. Wender/Fuel Procrssing Technology 48 (1996) 189-297 27s

Fig. 35. Conceptual flowsheet for the carbonylation of methanol to acetic acid (Zoeller, 1993).

to the incidence of side reactions such as hydrocarbon formation, will probably delay or
possibly prevent commercialization of a nickel based methanol carbonylation process.
Methanol can be converted to acetic acid via methyl formate in a two-step process
shown by the following reactions (Lee et al., 1990).
base
CH,OH+CO-+HCOOCH, U-7)

HCOOCH, + CH,COOH (T-8)


Many homogeneous and heterogeneous catalysts have been used for this conversion
to acetic acid (Torrence et al., 1991); all involve addition of methyl iodide. Most of the
chemistry is similar to that already proposed for the direct carbonylation of methanol
and is related to the commercial acetic acid process. It is not likely that this route to
acetic acid through methyl formate will be used industrially; it converts methanol to
acetic acid in two somewhat difficult steps.

7.3.3. Acetic anhydride


About a million tons of acetic anhydride are produced each year (Cook, 1993). It is
used to acetylate the hydroxyl group of a number of substrates to yield products such as
cellulose acetate (its chief use>, textile fibers, cigarette filters, various polymers,
coatings, aspirin, acetaminophen, flavors, fragrances, sweeteners and other uses includ-
ing as a general reagent for making acetates as a replacement for acetyl chloride (which
releases the corrosive HCl).
276 I. Weruler / Fuel Processing Technology 48 (1996) 189-297

The old process for production of Ac,O used the energy intensive ketene route
shown in the following equations for over 65 years to manufacture acetic anhydride. The
ketene is made from either acetone (Fife and Zhang, 1986) or acetic acid (Cook, 1993).
750°C
CH,COCH, + CH, = C = 0 + CH, (7.9)
Et.&‘%
CHJOOH + CH, = C = 0 + H,O (7.10)
700°C

CH, = C = 0 + CH,COOH + (CH,CO),O (7.11)


The latest and seemingly most viable synthesis uses a homogeneous catalysis system
to convert methyl acetate to Ac,O

CH,COOCH, + CO cat~st(CH3CO),0 (7.12)

Here a carbonyl group is apparently inserted between the 0-CH, bond of methyl
acetate. The mechanism for this insertion is complex, involving a noble metal, methyl
halide/ionic halide and an acetate salt. This synthesis uses the Tennessee-Halcon
process to convert methyl acetate to acetic anhydride.
In 1983, the Eastman Chemical Company in Kingsport, Tennessee became the first
major chemical company to begin again to use coal as a raw material for the
manufacture of industrial chemicals on a large scale (Agreda et al., 1992). They began
manufacturing AC*O from bituminous coal.
The Eastman Chemical Company uses two Texaco gasifiers to produce syngas from
coal. The process is based on technology developed independently by Halcon SD and by
Eastman in the 1970s (Worsham, 1994). As shown by the following equations, the entire
synthesis of Ac,O is developed from a syngas feed.
CH,OH + CO --) CH,COOH (7.13)
CH,OH + CH,COOH + CHJOOCH, + H,O (7.14)
CH,COOCH, + CO --) (CH,CO),O (7.15)
There are important differences between the aqueous acetic acid synthesis and the
anhydrous Ac,O system (Zoeller et al., 1992; Worsham, 1994). It is necessary to have
an iodide salt and a reducing agent (H,) present for the commercial synthesis of Ac,O.
A significant process implication is that the carbonylation of methyl acetate has a lower
heat of reaction; this lower thermodynamic driving force results in equilibrium limita-
tions for Ac,O conversion. The halolysis of methyl acetate by LiI is the key for the
increase in activity brought about by addition of inorganic promoters (Luft and Schrod,
1983).
CH,COOCH, + LiI -+ CH,I + LiOOCCH, (7.16)
Eastman will produce about 1.2 billion pounds per year of Ac,O from 1100 tons of
bituminous coal per day, thus becoming the world’s largest manufacturer of acetyl
products and eliminating the equivalent use of about 1.23 million barrels of oil per year.
1. Wrnder / Fuel Processing Technology 48 (1996) 189-297 277

To sum up, the Eastman process proceeds according to the following overall reaction
sequence.
Cu/ZnO
2H, + CO + CH,OH (7.17)

Rh.I-
CH,OH + CO + CH,COOH (7.18)
I6OT.25 am

CH,OH + CH,COOH --) CH,COOCH, (7.19)

CH,COOCH, + CO Rhc~p’rx(CH~CO),O (7.20)

(CH,CO),O + cellulose -+ cellulose acetate + CH,COOH (7.21)


(recycled)

Eastman has undergone a recent expansion and is well on its way to establishing
economic routes to the production of a series of industrial oxygenated chemicals with
this background.
An overall flow diagram for Eastman’s acetic anhydride complex is shown in Fig. 36.
There are upcoming developments in this area of chemicals production (Papp and
Baems, 1991). Hoechst has patented a process to co-produce acetic acid and Ac,O by
carbonylation of acetic acid and methyl acetate in a single reactor. Rhodium salts plus
ammonium or phosphonium iodide make up the catalyst system. BP has announced a
similar process. The different processes are compared in Table 19.

7.3.4. Vinyl acetate


The largest single use of acetic acid, almost six billion pounds per year, is in the
production of vinyl acetate (Summer and Zoeller, 1993). Vinyl acetate-derived polymers
are found everywhere in our modem society. Vinyl acetate gives us polyvinyl acetate,

COal
Gasification
Plant

Acenc Acid
suthr

Fig. 36. Block diagram of main steps required to produce acetic anhydride from coal (Cook, 1993).
278 I. Wemler / Fuel Processing Technology 48 (1996) 189-297

Table 19
Reaction conditions for the carbonylation of methanol and methyl acetate
Acetic acid Acetic anhydride Acetic acid, Acetic acid,
Monsanto Eastman anhydride Hoechst anhydride BP

Temperature 190- 195°C 190°C > 150°C 183°C


Pressure 30-35 bar 50 bar > IO bar 30 bar
Catalyst Rh Rh Rh Rh

used in adhesives, coatings, latex paints and finishes; polyvinyl alcohol, used in textile
sizing, paper-coating, pastes, plywood adhesives and cement additives; ethylene-vinyl
acetate polymer for packaging film, hot-melt adhesives and cable coverings; ethyl-vinyl
alcohol copolymer for flexible containers and bottles; and polyvinylbutyral for laminated
safety glass, automotive and architectural applications. Polyvinyl acetate formed for
magnet/wire insulation and vinyl chloride-vinyl acetate copolymer for protective
surface coatings are important uses of vinyl acetate in our surroundings.
Since 1960, most vinyl acetate is produced by the palladium-catalyzed reaction of
ethylene and acetic acid, involving the subsequent reoxidation of the reduced palladium
(Robinson, 1965; Daniels, 1978).

CH, = CH, + CH,COOH + l/20, ZCH, = CHOCOCH, + H,O (7.22)


vinylacetate
As acetic acid is made from methanol and CO, this process is based on syngas for
70% of its weight, giving overall yields of 90% and 95% based on ethylene and acetic
acid respectively (Whyman, 198.5).
Halcon has obtained patents describing a syngas based process leading to vinyl
acetate (Parshall and Ittel, 1992): methyl acetate (or DME) reacts with syngas to form
CH,CH(OAc), (EDA), which loses acetic acid to give vinyl acetate.
H,/co - AcOH
CHJOOCH, --, EDA + CH, =CHOCOCH, (7.23)
metal oxide

Eastman and Halcon have piloted routes from acetic anhydride to vinyl acetate but
the processes are not economical at present. These syntheses of vinyl acetate based on
syngas chemistries are still under development. A process operated by Celanese reacts
acetaldehyde and acetic anhydride (Ac,O) to form ethylidene diacetate, which is then
cracked to vinyl acetate over a sulfonic acid catalyst (Rizkalla and Goliaszewski, 1987)
CH,CHO + Ac,O + (AcO),CHCH, (7.24)

(Ac,O),CHCH,zAcOCH = CH, + CH,COOH (7.25)


heat vinyl acetate

7.3.5. Homologation (reductive carbonylation) of methanol to ethanol


Wender et al. (1949, 1951) showed that Co,(CO), catalyzed the reaction of methanol
with syngas at 185°C and 270 atm to yield ethanol in 39% selectivity and 70% methanol
conversion.
CH,OH + 2H, + CO --, CH,CH,OH + H,O (7.26)
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 279

The reaction is really a hydrocarbonylation of methanol but the term “homologation”


has come to mean the reaction of an organic molecule with syngas resulting in the
extension of the carbon chain by one methylene (-CH,-1 unit.
The homologation of methanol to ethanol received little attention following its
discovery. In the 197Os, interest was rekindled as a result of the oil crises. Numerous
reviews of further work in this area have since appeared (Albanesi, 1973; Wender, 1976;
Piacenti and Bianchi, 1977; Koermer and Slinkard, 1978; Pretzer and Kobylinski, 1980;
Chen et al., 1982; Fakley and Head, 1983; Jenner, 1989; Parshall and Ittel, 1992).
The homologation of methanol is thermodynamically favored as the pressure is
raised, but the chief problem is kinetic control over the number of reaction products,
including methane, acetates, acetic acid, acetals and CO,.
Cobalt complexes, first used, are good catalysts in their activity and selectivity for the
conversion of methanol and syngas to ethanol and acetaldehyde. Rhodium, usually more
active than cobalt in homogeneously catalyzed reactions involving methanol and syngas.
produces acids and esters, with ethanol a significant product at high partial pressures of
hydrogen (Dumas et al., 1979).
Keim (1989) has been studying the conversion of methanol to acetaldehyde
H,,‘CO
CH,OH -+ CH,CHO (7.27)

In the presence of a polar solvent (dioxane, sulfolane), a catalyst system [(Ph,P),N]


[Co(CO),]/12 gave a methanol conversion of 97% with 80% selectivity to the aldehyde.
A pilot plant was operated for over three years in cooperation with Union Rheinische
Braunkohle AG Wesseling. Keim feels that an acetaldehyde plant based on the homolo-
gation of methanol may be built in the near future. An engineering evaluation concluded
that this route to the aldehyde is more economical than the Wacker process (Jira et al.,
1976). However, additional capacity for the aldehyde is not presently needed because its
principal use, oxidation to acetic acid, has been largely replaced by the Monsanto
methanol-to-acetic acid process.
A strong increase in activity to ethanol has been obtained by adding iodide promoters
such as I, or CH,I (Gauthier-Lafaye et al., 1982). The homologation of methanol to
ethanol using phosphine ligands and Ru compounds has been described by Fiato (1980)
and by Bozik et al. (1980). With a catalyst containing Co, Ru, I and a phosphine ligand,
methanol conversion was 50-60% (Papp and Baems, 1991).
The relative rates of homologation of several alcohols with cobalt carbonyl and
iodine was studied by Berty et al. (1956). Ethanol reacts 42 times more slowly than
methanol (Jenner, 19891, consistent with an S, 2 displacement of iodide by Co(CO),.
Propanol and isopropanol are even less active. However, t-butyl alcohol, which readily
forms carbenium ions, reacts very rapidly, evidently by an S,l mechanism, to form
(CH,),CCH,OH.
There is no commercial plant that now uses the reductive carbonylation of methanol
to obtain either acetaldehyde or ethanol. Poor selectivity and severe operating conditions
are obstacles to the use of this route to ethanol. Although the conversion of starch to
ethanol by fermentation has taken over, there are those who feel that the catalyzed
homologation of methanol to ethanol has a future.
280 I. Wender/Fuel Processing Technology 48 (1996) 189-297

7.3.6. Homologation of ethers, esters and carboxylic acids


Braca et al. (1978), Braca et al. (1981) found that dimethyl ether and methyl acetate,
which are formed in the homologation of methanol, could be carbonylated further
200°C
CH,OCH, + 2H, + 2C0 --, CH,COOC,H, + H,O (7.28)
I50 atm

CHJOOCH, + 2H, + CO --) CH,COOC,H, + H,O (7.29)


These reactions were achieved with Ru/I- catalysts. Ethyl acetate selectivities
reached 80%. Cobalt and rhodium did not carbonylate these substrates (Sheldon, 1983).
The same type of catalyst, Ru/I-, is used in the conversion of carboxylic acids.
Acetic acid, for instance, yields propionic acid plus small amounts of n-butyric and
pentanoic acids (Knifton, 1981a, Knifton, 1981b, Knifton, 1981~).
Hz/co Hz/co
CH,COOH --$ CH,CH,COOH + CH,CH,CH,COOH,etc. (7.30)
RU/I_

7.3.7. Methyl formate


Methyl formate is an intriguing chemical that has been proposed as a building block
in C, chemistry (Keim, 1983, Keim, 1987). The paper by Lee et al. (1990) provides an
excellent review of the synthesis, properties, and uses of methyl formate. At present,
however, methyl formate does not qualify as a major building block in C , chemistry, but
an examination of the various methods of its synthesis and of its possible uses as a
feedstock for conversions to a number of important chemicals warrants discussion.
Methyl formate is convenient to handle, store, and transport as a stable liquid; it boils at
31.5”C and is easily separated from methanol, b.p. 64.7”C. It could be handled, stored
and transported in a manner resembling that for liquefied petroleum gas (LPG).
Methyl formate is synthesized commercially by the alkali methoxide-catalyzed
reaction of methanol and CO (Christiansen, 1919). Methoxide ion (OCH,)- is a strong
base and nucleophile and it attacks the electrophilic CO molecule
CH,O- + C = 0- + (CH,OCO) - (7.31)
(CH,OCO)- + CH,OH + HCOOCH, + CH,O- (7.32)
NaOCH,
CH ,OH + CO + HCOOCH, (7.33)

A drawback, however, is that dry methanol must be used in this synthesis. The CO
stream containing at least 50% CO and not more than 10 ppm of water or 50- 100 ppm
of CO, should be available at about 600 psig. Conversion of methanol and CO is
typically 30% and 95%, respectively; selectivity to methyl formate is = 99%. The
presence of hydrogen or nitrogen in the feed gas does not interfere.
Methyl formate itself has limited application as a solvent and as an insect control
agent. It is mainly used as an intermediate in the production of formic acid and
formamides.
There are a number of alternate ways of synthesizing methyl formate that are of
interest. The Mitsubishi Gas Chemicals Co. had a unit that produced over 20000 tons
1. Weder/ Fuel Processing Technology 48 (1996) 189-297 281

per year (tpy) of methyl formate by the dehydrodimerization of methanol (Komatsu and
Yoneoka, 1988; Nojiri and Misono, 1993). The plant may have been abandoned because
of the rising price of methanol.

2CH,OH + HCOOCH, + 2H, (7.34)


250 - 300°C
3-5atm
Cu/Zr/Al,O,

Perhaps the most promising route to methyl formate is by its direct synthesis from
syngas by the Brookhaven and related processes or by the concurrent synthesis of
methanol and methyl formate. In each case, methyl formate can be the principal product,
separated from by-product methanol by distillation.

2H, + 2C0 -+ HCOOCH, (7.35)

Good yields of methyl formate and methanol can be achieved at 50-100°C and
= 600 psig. Significant improvements in these processes may be achieved in the future.
Methyl formate has a number of potential industrial uses. It is sometimes necessary to
separate H, and CO in syngas or to adjust their ratio. Under certain conditions CO
and/or H? must be recovered from non-conventional sources such as off-gases from the
production of steel or from vent streams from chemical processes.
The commercial carbonylation route to methyl formate can be operated with as little
as 50% CO in the feed gas, as mentioned earlier. The syngas can be enriched in CO in a
once-through methanol synthesis unit or by removal of H, by a membrane, by diffusion
or by some other means (Halcon SD, 1985). The CO-rich syngas is carbonylated to
methyl formate, which could be stored for further utilization including decomposition to
pure CO. The hydrogen-rich vent stream has many uses (Lee et al., 1990).
Methyl formate can be decomposed to give methanol and CO

HCOOCH, -+ CH,OH + CO (7.36)

Under these conditions methyl formate decomposes selectively even in the presence
of excess methanol. With a KCl/activated carbon catalyst, 99.5% of methyl formate is
converted to methanol and CO with a selectivity exceeding 99%. Methyl formate is thus
a good CO carrier.A 1: 1 syngas can be obtained from methyl formate by treating it at
300-350°C over alkali or alkaline earth oxides

HCOOCH, --+ 2H, + 2C0 (7.37)

This syngas has the stoichiometry and purity that makes it a suitable feed for
chemical syntheses based on a temporarily out-of-commission coal gasifier or for a
hydroformylation unit.
Methyl formate has been reported to isomerize selectively to acetic acid (Pruett and
Kacmarik, 1982).

HCOOCH,TCH,COOH (7.38)

No CO is consumed in this reaction but it only takes place under CO pressure.


Although many transition metals catalyze this isomerization, it appears that a Rh-LiI
282 1. Wender/Fuel Processing Technology 48 (19961 189-297

catalyst is highly efficient for the reaction at 180°C. The methyl formate conversion and
the molar selectivity to acetic acid are both over 99%. Lithium iodide aids the reaction
by cleaving methyl formate and the intermediate methyl acetate to CH,I (Schreck et al.,
1988). Methyl iodide is an intermediate both in the direct synthesis of acetic acid from
methanol and CO and in the isomerization of methyl formate to acetic acid. The
difference is that the methyl formate synthesis involves insertion of CO into an
oxygen-hydrogen bond whereas acetic acid is formed by insertion of CO between the
carbon-oxygen bond in methanol.
This isomerization can find use when methyl formate is an undesirable by-product (as
in the oxidation of butane) or if the formate could be produced more cheaply. In some
locations, dehydrogenation of methanol to methyl formate, with subsequent isomeriza-
tion, could be an economic route to acetic acid.
It is unlikely, however, that this isomerization, which has been known for many
decades (Dreyfus, 1929; Roper et al., 1985), can compete with the Monsanto process for
the synthesis of acetic acid from methanol and carbon monoxide.

7.3.8. Formic acid


Until recently, most US formic acid production was a by-product from the liquid-phase
oxidation of butane to acetic acid. With the advent of cheaper acetic acid made by
carbonylation of methanol, the importance of this by-product route from hydrocarbons
has declined. The present method for the synthesis of formic acid is by hydrolysis of
methyl formate but, as retroesterification occurs readily, HCOOH is generally made by
first synthesizing formamide (HCONH,) (Peltzman, 1984)
80- IOO”C
HCOOCH, + NH, + HCONH, + CH,OH (7.39)
5 atm

The formamide is then continuously hydrolyzed to HCOOH

HCONH, + H,O + HCOOH + NH, (7-40)


About 400000 tons per year of formic acid are produced worldwide. Its major
industrial uses are in the textile and leather industries. It has been used in Europe for
many years as a silage preservative.

7.3.9. Ethylene glycol


The direct synthesis of ethylene glycol from syngas is a most attractive route from a
raw material point of view (Table 171, but this has been an elusive target. At present,
ethylene glycol is manufactured commercially by the reaction of ethylene with oxygen
over a silver catalyst. The ethylene oxide so formed is then hydrolyzed to ethylene
glycol. Over five billion pounds of ethylene glycol were produced in the United States in
1993.
The direct production of ethylene glycol from syngas was discovered by DuPont
using cobalt as the catalyst (Aquild et al., 1983). More recently, Union Carbide
synthesized ethylene glycol directly from syngas with a rhodium catalyst, but the
reaction rate is too low, requiring severe conditions and difficult separation of the glycol
1. Wender / Furl Proc’rssin~ Technology 48 (I 996) 189-297 I?83

from co-produced methanol, propylene glycol, glycerol and other by-products. The
process has been carried to a semi-works but is not economic at present.
Kh
3H,+CO -+ HOCH,CH,OH (7.41)
210- 250°C
500 - 3400 atm

Many companies are endeavoring to find a new commercial process for the synthesis
of ethylene glycol. As shown in Fig. 33, oxidative carbonylation of methanol to
dimethyl oxalate (itself useful in agriculture as a compound which releases nitrogen
slowly, in the food industry and in pharmaceuticals) followed by hydrogenation yields
ethylene glycol.
Perhaps the most promising process involves the synthesis of dimethyl oxalate by
oxidation of methanol with oxygen and nitric acid, a reaction applied commercially
since 1978 by Ube Industries/Japan (Roper, 1991; Forster, 1976; Wallet-, 1985).
2CH,OH + 2N0 + l/20, -+ 2CH,ONO + H,O (7.42)

Anhydrous methyl nitrite is carbonylated in a second stage over a Pd/Fe catalyst,


furnishing a 97% yield of dimethyl oxalate, which is then hydrogenated over a
ruthenium catalyst to ethylene glycol in 90% yield.
2CH,ONO + 2C0 + H,COOC - COOCH, + 2N0 (7.43)
CH,OOC - COOCH, + 4H, + HOCH,CH,OH + 2CH,OH (7.44)
The nitric oxide (NO) is converted to methyl nitrite for recycle
NO + CH,OH + l/20, --) CH,ONO + 1/2H,O (7.45)
This is an attractive route to ethylene glycol involving methanol, CO and HZ in
different stages.
Formaldehyde derived from methanol can be converted to glycolic acid derivatives
(Fig. 33) which can be hydrogenated to ethylene glycol. Other routes to the glycol by
reaction of formaldehyde have been studied by Celanese (Kollar, 1982), by Monsanto
and by Exxon (Chem. Eng. News, 1983).

7.3. IO. Dimethyl carbonate


Dimethyl carbonate is a non-toxic chemical that is being promoted as a replacement
for the toxic intermediates phosgene and dimethyl sulfate. This follows the current trend
in the chemical industry to reduce the use of highly toxic substances.
A glance at Fig. 33 indicates that dimethyl carbonate can be synthesized by the
oxidative carbonylation of methanol. Roman0 et al. (1980) synthesized dimethyl carbon-
ate from methanol, carbon monoxide and oxygen in the liquid phase in a slurry reactor
using a cuprous chloride catalyst
2CH,OH + CO -I- l/20, -+ (CH,O),CO + HZ0 (7.46)

EniChem has produced dimethyl carbonate in a 5000 tons per year plant since 1983
(Mauri et al., 1985). The plant output was expanded to 8800 tons per year in 1988 to
meet the growing interest in the carbonate. The chief captive uses are as a phosgene
substitute in monomers for optical resins and in the manufacture of resins for polycar-
284 I. Wender/ Fuel Proce.ssin,g Technology 48 (1996) 189-297

bonates and carbamates; it is also used in the synthesis of synthetic lubricants and in
polyurethane manufacture. Dimethyl carbonate is used in place of the toxic dimethyl
sulfate in methylation reactions. It is being considered in fuel reformulation as an
oxygen-rich octane enhancer in motor fuels.
Dow has also developed a process for the synthesis of dimethyl carbonate (Chem.
Eng. News, 1987). The carbonate is produced via a vapor-phase carbonylation of
methanol using a solid catalyst of activated carbon impregnated with cupric chloride. It
is claimed that this process reduces corrosion problems and simplifies product recovery.

7.3.11. Methyl methacrylate


Methyl methacrylate polymers have many end uses and about one million tons per
year are manufactured worldwide. The o-methyl group of polymethylacrylates imparts
the stability, hardness and stiffness of methacrylate polymers. The current process uses
an acetone cyanhydrin route which utilizes methanol in its manufacture (Porcelli and
Juran, 1986). Environmental pressures have resulted in on-site synthesis of HCN to
avoid shipment of by-product HCN (from manufacture of acrylonitrile).
To avoid the use of HCN, methyl methacrylate can be manufactured from propylene
C,H, + l/20, + CO + CH,OH 4 CH, = C(CH,)COOCH, + H,O (7.47)
or
C,H, + 0, -I-CO + CH,OH --, MMA + 2H,O (7.48)
Carbon monoxide is expensive to transport and the amount needed for a reasonably
sized methyl methacrylate plant is much less than the economic size for low cost CO
production. However, a propylene carbonylation route to MMA would make sense as
part of a complex that produced other C , derivatives.
One plant in Germany manufactures methyl methacrylate from propionaldehyde,
which is produced from ethylene via the hydroformylation of ethylene
C,H, + CO + HCHO + CH,OH + MMA + H,O (7.49)
A disclosure from SRI International reports that Shell will produce MMA from
propyne (CH, = CH) at a cost lower than that of the acetone cyanohydrin and other
processes (Chemical Marketing Reporter, 1993). Although Shell has not yet announced
any such plans, it is indicated that the process is described in US and European patents,
has been piloted and is ready to go into commercial production. Evidently, the MMA
would be produced according to the following equation
CH,C = CH + CO + CH,OH --) CH, = C(CH,)COOCH, (7.50)
The propyne (methylacetylene) is obtained from the C, product stream from an
ethylene plant.

7.3.12. Chlorinated hydrocarbons


Methyl chloride, methylene chloride, chloroform and carbon tetrachloride are derived
mainly by reaction of methanol and chlorine. The two principal processes for the
commercial production of methyl chloride are the chlorination of methane and the
I. Wmdrr/ Furl Processing Technology 48 (19961 189-297 285

reaction of methanol and hydrogen chloride. With methanol, the process produces and
recycles the HCl. In the chlorination of methane, I mol of HCl has to be disposed of,
generally as 3 1 wt% NC1 (Calkins, 1984).
In the US, the liquid-phase reaction of methanol and HCl has the widest use. In
Europe and Japan, the gas-phase methanol hydrochlorination process is used more than
in the US (Holbrook, 1993).
With methanol as starting material, the reactions that occur are
CH,OH + HCl -+ CH,CI + H,O (7.51)
CH,Cl + Cl, -+ CH,Cl, + HCI (7.52)

CH,CI + 2C1, + CHCl, + 2HCl (7.53)

CH,Cl + 3C1, --) Ccl, + 3HCI (7.54)

7.3.13. Me~hylamines
The methylamines have a number of applications, with dimethylamine used the most,
monomethylamine the next and trimethylamine the least. They are produced by the
reaction of methanol and ammonia in the vapor phase using a dehydration catalyst at
about 450°C and l-20 atm (Calkins, 1984).
CH,OH + NH, + CH,NH, + H,O (7.55)

2CH,OH + NH, -+ (CH,),NH + 2H,O (7.56)

3CH,OH + NH, -+ (CH,),N + 3H,O (7.57)

Depending on the amine product desired, a 2: 1 to 6: 1 ratio of ammonia to methanol is


employed. The reaction reaches approximate thermodynamic equilibrium as a function
of the ammonia to methanol ratio. The more utilized dimethylamine is produced in only
30-40% of the total amine yield. The other amines, plus additional ammonia, are
recycled to the reactor. Yields based on methanol can reach 99%. Silver/alumina
catalysts with silver phosphate, molybdenum sulfide or cobalt sulfide promoters are the
catalysts of choice.

8. Miscellaneous reactions of syngas or carbon monoxidekarbonylation)

8.1. Introduction

The major reactions of syngas and of CO (carbonylation) have been discussed but
there are other reactions of these gases which are too numerous to detail. Some of these
are used commercially today, hundreds of others appear interesting but have not been
exploited (Colquhoun et al., 1991). A number of these reactions were used in industrial
processes that have been improved or replaced by cheaper or environmentally superior
syntheses. Many of these reactions of syngas or of CO are proprietary and it is difficult
to obtain detailed information on their commercial viability. There is sometimes a fine
line between commercial and pre-commercial processes.
286 1. Wender/ Fuel Processing Technology 48 (1996) 189-297

8.2. Reppe carbonylation reactions

Reppe and his coworkers, between 1938 and 1945, invented a series of reactions of
acetylenes and olefins with CO and protic reactants such as water, alcohols or amines to
give carboxylic acids and their derivatives (Reppe, 1953; Reppe and Vetter, 1953; Falbe,
1970). The applicability of these syntheses was amazingly general. At one time, some
600000 tons of Reppe carbonylation products were made per year (Mullen, 1980). The
following are examples.

HC = CH + CO + H,O --) CH 2 = CHCOOH (acrylic acid) (8.1)

HC = CH + CO + ROH + CH, = CHCOOR (8.2)

HC = CH + CO + R,NH + CH, = CHCONR, (8.3)


The same types of reactions could be carried out with olefins

CH, = CH, + CO + ROH + CH,CH,COOR (8.4)


CH, = CH, + CO + R,NH + CH,CH,CONH,,etc. (8.5)
The carbonylation of acetylenes and olefins could be carried out stoichiometrically
with metal carbonyls at atmospheric pressure with the carbonyl acting as both a supplier
of CO and a catalyst. Commercially, however, the syntheses were carried out at high CO
partial pressures with catalytic amounts of transition metal carbonyls produced in situ
from various metal salts; nickel compounds were most effective. Acetylenes react more
easily than olefins, which require higher temperatures and partial pressures than
acetylenic feedstocks.
Newer plants for the synthesis of acrylic acid use a process based on the oxidation of
propylene

H,C = CH - CH, + 0, tc? H,C = CHCHO (8.6)


-H,O

The aldehyde is then oxidized in air with a molybdenum catalyst to acrylic acid,
CH, =CHCOOH (Weissermel and Arpe, 1993).
The syntheses of acrylic acid and acrylate esters are examples of Reppe reactions.
These compounds are used in the manufacture of polymers and are made from
acetylene, CO and water or an alcohol. Acrylic acid is formed in a catalytic process by
treating acetylene in tetrahydrofuran with CO, water and NiBr, at about 200°C and 60
atm (Toepel, 1964). Although this process has been displaced by one based on the
oxidation of propylene in the US, it is still used elsewhere to a considerable extent
(Weissermel and Arpe, 1993). Reppe and coworkers accumulated enough experience to
be able to carry out reactions of acetylene itself under elevated pressures without
incident. There have been several instances in which explosions have occurred when
acetylene reactions were carried out at elevated pressures and temperatures. Worldwide
capacity for acrylic acid is about two million tons per year.
I. Wet&r /Furl Proctwing Technokqy 48 (1996) 189-297 287

8.3. The Koch reaction: olefins to carboxylic acids

In the Koch reaction, olefins are converted to highly branched carboxylic acids using
strong acids such as H,SO, or H,PO,/BF, as catalysts. The olefin adds a proton to
form a carbenium ion, which is then isomerized

RCH,CH = CH,H-;RCH,;HCH, + RdHCH,CH j -+ &(CH,)* (8.7)

RC(Cw,), + CO + H,O 2o-Jo”c RC( CH,),COOH


20- lOOatm

Esters of these branched so-called neo-acids exhibit unusual thermal and oxidative
stability and are difficult to saponify. These properties are largely the result of alkyl
branching LYto the carboxyl group. Based on their great stability, the acids are suitable
components of synthetic oils. Several companies now operate industrial processes based
on this reaction.

8.4. Homologation of carboxylic acids

Knifton (1981a), Knifton (1981b) has discovered a useful way of lengthening the
carbon chain of carboxylic acids by adding one carbon at a time. The reaction, using
acetic acid as an example, follows
RuO,, HI
CH,COOH+CO+2H, + CH,CH,COOH + H,O (8.8)
1OOatm.22OT

With propionic acid, two isomeric products are obtained


CH,CH,COOH + CO + 2H, + CH,CH,CH,COOH + (CH,),CHCOOH
(8.9)
The normal to iso acid ratio varies from 4 to 8. Since acetic acid is itself synthesized
entirely from syngas, all these acids can be built exclusively from syngas with no need
for petrochemical feedstocks.

8.5. Carbonylation of nitro compounds to isocyanates

Nitroaromatic compounds react with carbon monoxide to give the corresponding


isocyanates in the presence of palladium catalysts (Parshall and Ittel, 1992). With
nitrobenzene, the reaction may be written

C,H,NO, + 3C0 2 C,H,N = C = 0 + 2C0, (8.10)


190°C

This is a reductive carbonylation reaction and it has advantages over the usually
practiced two-step syntheses of isocyanates from the toxic compound phosgene (COCI, >

C,H,NO, %6H$JH~C~‘2C6H5NC0 + 2HCI (8.11)


288 I. Wender/ Fuel Processing Technokqy 48 (1996) 189-297

The reaction given in Eq. (8.10) is an environmentally better way to prepare


diisocyanates such as toluene diisocyanate (TDI) and methylene diphenylisocyanate
(MDI), which are used to manufacture polyurethane for fibers, coatings and adhesives;
their largest use is in the preparation of foams.
The structures of MD1 and TDI are given below (Scheme 1)
Most modem polyurethanes are prepared from diisocyanates such as TDI or MD1 in
combination with diols such as ethylene or propylene glycol, polyether diols, aliphatic
polyester diols or siliconediols.

8.6. Amidocarbonylation reactions

In general, many carbonylation reactions are being used for the production of
high-value specialty chemicals. As an example, the synthesis of amido acid precursors
from olefins or aldehydes plus syngas and amides has been under study by Texaco (Lin
and Knifton, 1992; Knifton et al., 1993). Wakamatsu et al. (1971) discovered the
amidocarbonylation reaction but this potentially useful reaction was not thoroughly
investigated until recently. Knifton and coworkers have studied the chemistry involved
to synthesize a variety of specialty chemicals by tailoring cobalt and rhodium catalysts
to make particular products selectively. These amidocarbonylation products can be
formed from a wide variety of olefins and aldehydes which react with syngas to give
useful products, including specialty surfactants, surface active agents (C ,4-C ,6 alkyl
amido acids), food additives, chelating agents, and intermediates for sweeteners such as
aspartame.
With available inexpensive C ,2-C 14 straight-chain terminal olefin feedstocks, amido
acids with up to 95% linearity can be synthesized
Co,Rh, 100°C
CH,(CH,),,CH = CH, + CH,CONH, + H, + CO +
50- 130atm
CH,(CH,) ,,CH(COOH)NHCOCH, (8.12)

8.7. Esters from epoxides

An example of an interesting carbonylation reaction which occurs from easily


available compounds under comparatively mild conditions (6X, 130 atm) is the

Scheme I.
I. Wender/ Fuel Processing Technology 48 (1996) 189-297 289

synthesis of B-hydroxyesters from epoxides. Carbonylation of ethylene oxide or propy-


lene oxides in methanol to the ester proceeds as follows (Heck, 1963; Eisenmann et al.,
19611

OH
CofCO), I
RHF-,C”2 -!- CO + CH,OH + RCHCH,COOCH,
(8.131
0

This synthesis does not seem to have been generally exploited.

8.8. Esters from ethers

Ethers may be carbonylated to esters with both homogeneous and heterogeneous


transition metal catalysts (Piacenti and Bianchi, 1962).
ROR + CO --) RCOOR (8.14)
Catalysts that have been used include Ni metal, Co/SiO, or NiI,/SiO,. Hydrogen
halides such as HI promote the reaction, as in the conversion of methanol to acetic acid;
the ether is probably cleaved by the acid promoter to yield alkyl halides and an alcohol.
The alkyl halide is catalytically carbonylated to give an acyl halide that reacts with the
alcohol to give the ester. This reaction may be applied to the conversion of dimethyl
ether to esters
ROR+HI*RRI+ROH (8.15)
RI + CO + RCOI (8.16)
RCOI + ROH + RCOOR + HX (8.17)

Acknowledgements

The support of the Electric Power Research Institute (EPRI) is gratefully acknowl-
edged. Thanks are due to G.A. Mills for invaluable help and to J. Inga in preparation of
the manuscript.

References

Abbott, P.E.J., Conduit, M.R. and Mansfield, K., 1989. Presented at the 1989 World Methanol Conference,
Houston, TX, December 5-7.
Agreda, V.H. and Zoeller, J.R., 1993 (Ed%). Acetic Acid and its Derivatives, Marcel Dekker, New York.
Agreda, V.H., Pond, D.M. and Zoeller, J.R., 1992. Chemtech, 22 (March): 172.
Aker Engineering, 1982. Petrole Informations, May 13.
Albanesi, G., 1973. Chim. Ind. (Milan), 55: 3 18.
Alpert, S.B., 1991. Annu. Rev. Energy Environ., 16: I.
Anderson, R.B., 1956. In Emmett, P.H. (Ed.), Catalysis, Vol. IV, Reinhold Publishing Corp., New York, p. 1.
290 I. Wender/ Fuel Processing Technology 48 (1996) 189-297

Anderson, R.B., 1984. The Fischer-Tropsch Synthesis. Academic Press, New York.
Anderson, R.B., Feldman, J.B. and Starch, H.H., 1952. Ind. Eng. Chem., 441 2418.
Ansorge, J. and Hoek, D., 1992. Conversion of Natural Gas to Clean Transportation Fuels: The Shell Middle
Distillate Synthesis Process (SMDS). Am. Chem. Sot. Prepr., Div. Pet. Chem., 37: 833.
Aquilo, A., Alder, J.S., Freeman, D.N. and Voorhoeve, R.J.H., 1983. Hydrocarbon Process., (March): 57.
Arakawa, H., 1984. Chem. Lett., 1607.
Argauer, R.J. and Landolt, G.R., 1972. US Patent 3,702,886.
Baglin, E.G., Atkinson, G.B. and Nicks, L.J., 1981. Ind. Eng. Chem. Prod. Res. Dev., 20: 87.
Ballivet-Tkatchenko, D. and Tkatchenko, I., 1981. J. Mol. Catal., 13: 1.
Bart, J.C.J. and Sneeden, R.P.A., 1987. Catal. Today, 2: 1.
Bartholomew, C.H., 1988. In Paal, Z. and Menon, P.G. @is.), Hydrogen Effects in Catalysis, Marcel Decker,
New York, Chap. 20, p. 543.
Bartholomew, C.H., 1991. Recent developments in Fischer-Tropsch catalysis, in Guczi, L. (Ed.), New Trends
in CO Activation. Elsevier, Amsterdam, p. 159.
BASF, 1923. US Patents 1,558,559 and 1,569,775.
Berty, J., Marko, L. and Kallo, D., 1956. Chem. Tech. (Berlin), 8: 260.
Betty, J.M., Krishman, C. and Elliot, J.R., Jr., 1990. Chemtech, 20 (October): 624.
Bhasin, M.M., 1975. US Patent 4,014,913.
Bhasin, M.M., Bartley, W.J., Ellgen, PC. and Wilson, T.P., 1978. J. Catal., 54: 120.
Biloen, P. and Sachtler, W.M.H., 1981. Adv. Catal., 30: 165.
Biloen, P., Helle, J.N. and Sachtler, W.M.H., 1979. J. Catal., 58: 95.
Botteghi, C., Branca, M. and Saba, A., 1980. J. Organomet. Chem., 184: C17.
Bozik, J.E., Kobylinski, T.P. and Pretzer, R.W., 1980. US Patent 4,239,924.
Braca, G., Sbrana, G., Valentini, G. and Andrich, G., 1978. J. Am. Chem. Sot., 100: 5238.
Braca, G., Paladini, L., Sbrana, G., Valentini, G., Andrich, G.; Gregori, V., 1981. lnd. Eng. Chem. Prod. Res.
Dev., 20: 115.
Brady, R.C., 111; Pettit, R., 1980. J. Am. Chem. Sot., 102: 6181.
Brady, R.C., Ill and Pettit, R., 1981. J. Am. Chem. See., 103: 1287.
Brewester, E.A.V., 1976. Chem. Eng., 83: 90.
Bridger, G.W., Spencer, MS., 1989. In Twigg, M.V. (Ed.), Catalysis Handbook, 2nd edn., London, Chap. 9.
Broden, G., Rhodin, T.N., Bruckner, C., Benbow, R. and Hurych, A., 1976. Surf. Sci., 59: 593.
Brown, D.M., Bhatt, B.L., Hsiung, T.H., Lewnard, J.J. and Waller, F.J., 1991. Catal. Today, 8: 279.
Burch, R., Golunski, S.E. and Spencer, MS., 1990. Catal. Lett., 5: 55.
Calkins, W.H., 1984. Catal. Rev. - Sci. Eng., 26: 347.
Camell, D.W., 1977. Medium-Btu Gas as Feedstock for Chemicals and Synthetic Fuels, Presented at the
Gorham International - Institute of Gas Technology Conference, Low BTU Gas: Its Future, Dundee,
Illinois.
Chang. C.D., 1983. Catal. Rev. - Sci. Eng., 25: I.
Chang, C.D., 1988. Stud. Surf. Sci. Catal., 36: 127.
Chem. Eng. News, 1983. April 11, p. 41.
Chem. Eng. News, 1987. September 28, p. 26.
Chem. Eng. News, 1996. July 1, p. 7.
Chemical Marketing Reporter, 1992.
Chemical Marketing Reporter, 1993. February 15.
Chen, M.J., Feder, H.M. and Rathke, J.W., 1982. J. Am. Chem. Sot., 104: 7346.
Chen, N.Y., Garwood, W.E. and Dwyer, F.G., 1989. Shape Selective Catalysis in Industrial Applications,
Marcel Decker, New York, p. 221.
Chinchen, G.C. and Spencer, M.S., 1991. Catal. Today, 10: 293.
Chinchen, G.C., Denny, P.J., Parker, P.G., Spencer, MS. and Whan, D.A., 1987. Appl. Catal., 30: 333.
Chinchen, G.C., Denny, P.J., Jennings, J.R., Spencer, M.S. and Waugh, K.C., 1988. Appl. Catal., 36: I.
Chinchen, G.C., Mansfield, K., Spencer, M.S., 1990. Chemtech, 20 (November): 692.
Christiansen, J.R., 1919. US Patent 1,302.Oll.
Clarke, L.B., 1991. Management of By-products from IGCC Power Generation, IEACR/38, May 1991, IEA
Coal Research, London.
I. Wender/Fuel Processing Technology 48 (1996) 189-297 291

Cohen, T. and Ensor, L, 1996. Nywerheidsberigte Industrial Reports, July.


Cohn, E.M., 19.56. The Isosynthesis. In Emmett, P.H.,(Ed.J, Catalysis, Vol. IV, Reinhold Publishing Corp.,
New York, p. 443.
Colquhoun, H.M.. Thompson, D.J. and Twigg, M.V., 1991. Carbonylation, Plenum Press, New York, N.Y.
Consiglio, G., Botteghi, C., Salomon, C. and Pino, P., 1973. Angew. Chem. Int. Ed, Engl., 12: 669.
Cook, S.L., 1993. In Agreda, V.H. and Zoeller, J.R. (Eds.). Acetic Acid and Its Derivatives, Marcel Dekker.
New York, Chap. 9, p. 145.
Cook, S.L., 1995. Prepr. Div. Fuel Chem. Am. Chem. kc., 40 (April 2): 124; Chem. Eng. News, (April 17):
26.
Comils, B. and Weibus, E., 1995. Chemtech, 25 (January): 33.
Courty, P.E., Chaumette, P., Rainbauh, C. and Travers, P.H., 1990. Rev. Inst. Fr. Pet., 45 (July-A~gd: 561.
Csicsery, S.M., 1976. In Rabo, J.A. (Ed.), Zeolite Chemistry and Catalysis, ACS Monogr., No. 171, p. 680.
Daly, F.P.. 1984. J. Catal., 89: 13 I.
Daniels, W., 1978. Kirk-Othmer Encyclopdia of Chemical Technology, Vol. 23, 3rd Edn., John Wiley & Sons,
p. 817.
Davies, P.R., Snowden, F.F., Bridger, G.W., Hughes, D.O. and Youn,,0 D.W., 1966. US Patent 1.010,871.
Dombeck, B.D., 1983. Adv. Catal., 32: 325.
Douglas. J., 1994. EPRI J., (I9 March): 6.
Dreyfus, H., 1929. US Patent 1,697,109.
Dry, ME., I98 I. In Anderson, J.R. and Boudart, M (Ed%), The Fischer-Tropsch Synthesis, Catalysis Science
and Technology, Vol. I, Springer-Verlag, New York, Chap. 4.
Dry, ME., 1990. Catal. Today, 6: 183.
Dry, M.E., 1996. Appl. Catal., A, General, 138: 3 19.
Dumas, H., Levisalles, J. and Rudler, H., 1979. Organomet. Chem.. 117: 239.
Eilers, J., Posthuma, S.A. and Sie, S.T., 1990. Catal. Lett., 7: 253.
Eisenberg, B., Lahn, G.C., Fiato, R.A. and Say, G.R., 1993. Preprint, presented at the Alternate Energy ‘93,
Colorado Springs, April 1993.
Eisenberg, B., Ansell, L.L., Fiato, R.A. and Bauman, R.F.. 1994. Preprint, presented at the 73rd Annual GPA
Convention, New Orleans, LA, March 7.
Eisenmann, J.L., Yamartino, R.L. and Howard, J.F., 1961. J. Org. Chem., 26: 102.
Eisenlohr, K.H. and Gaensslen, H., 1981. Fuel Processing Technol., 4: 43.
El Sawy, A.H., 1990. Contractor’s Report, US Department of Energy, SAND 89-7 I5 I. UC- 125.
Fahey, D.R., 1987. Industrial Chemicals via C, Processes, ACS Symp. Ser., No. 328.
Fakley, M.E. and Head, R.H., 1983. Appl. Catal., 5: 3.
Falbe, J., 1970. Carbon Monoxide in Organic Synthesis, Springer-Verlag, Berlin.
Fiato, R.A., 1980. US Patent 4.233.466.
Fiato, R.A., Iglesia, E. and Soled, S.L., 1989. US Patents 4,738,948 and 4,822,824.
Fife, W.K. and Zhang, Z.-d., 1986. Tetrahedron Lett., 27: 4933.
Fischer, F. and Tropsch, H., 1926. Brennst-Chem., 7: 97.
Fischer, F., Roelen. 0. and Feisst, W., 1932. Brennst.-Chem., 1.3: 461.
Flory, P.J., 1950. J. Am. Chem. Sot.. 58: 1877.
Forster, D.. 1976. J. Am. Chem. Sot., 98: 846.
Forzatti, P., Tronconi, E. and Dasquon, I., 1991. Catal. Rev. - Sci. Eng., 33: 109.
Fourie, J.H., 1992. Presentation at the Alternate Energy I992 Conference, Kiawak Islands, April.
Fourth Grove Fuel Cell Symposium, 1995. Platinum Met. Rev., 39: 160.
Fowler, R., Connor, H. and Baehl, R.A., 1976. Chemtech, 6 (December): 772.
Friedel, R.A. and Anderson, R.B.. 1950. J. Am. Chem. Sot., 72: 1212.
Frohning, C.D., Koelbel, H., Ralek, M., Rotti g, W., Schnur. F. and Shulz, H., 1982. Fischer-Tropsch Process,
Chemical Feedstocks from Coal, John Wiley & Sons, New York: 300.
Garwood, W.E., 1983. ACS Symp. Ser., No. 218, p. 383.
Gauthier-Lafaye, Perron, R. and Colleuille, Y., 1982. J. Mol. Catal, 17: 339.
Geerts, J.W.M.H., Hoebink, J.H.B.J. and van der Wiele, K., 1990. Catal. Today, 6: 613.
Geertsema, A., 1993. Plenary Lecture, Presented at the Tenth Annual International Meeting, Pittsburgh Coal
Conference, Pittsburgh, PA, September.
292 I. Wet&r /Fuel Processing Technology 48 f 1996) 189-297

Geertsema, A., 1996. Private communication.


Gogate, M., Kulib, C.J. and Lee, S., 1993. Prepr. Div. Fuel Chem. Am. Chem. Sot., Vol. 38 (31, Chicago,
August, p. 1100.
Goodwin, J.G., Jr., 1991. Prepr. Div. of Pet. Chem. Am. Chem. Sot., Atlanta, GA, April, p. 156.
Gray, D., Tomlinson, G. and El Sawy, A., 1991. Hybrid Plant Coal Liquefaction Concept at the Great Plains
Synfuels Plant, The Mitre Corporation, McLean, VA.
Haag, W.O. and Huang, T.J., 1979. US Patent 4,159,995
Haag, W.O. and Huang, T.J., 1981. US Patent 4,279,830.
Haag, W.O., Kuo, J.C. and Wender, I., 1987. Gasification for Synthesis of Fuels and Chemicals. Chap. 5 in
Coal Gasification: Direct Application and Syntheses of Chemicals and Fuels, DOE/ER-0326, p. 117.
Halcon SD, 1985. US Patent 4,524,581.
Haldor Tops&e A.S.. 1985. US Patent 4,536,485.
Hall, C.C., Gall, D. and Smith, S.L., 1952. J. Inst. Pet. Technot., 38: 845.
Hansen, J.B. and Joensen, F., 1991. In Holman, A., Jens, J.J. and Kolboe, S. (Eds.), Natural Gas Conversion
Symposium Proceedings, Elsevier, Amsterdam, p. 457.
Heck, R.F., 1963. J. Am. Chem. Sot., 85: 1460.
Henrici-Olive, G. and Olive S., 1984. Catalyzed Hydrogenation of Carbon Monoxide, Springer-Verlag, Berlin.
Herington, E.F.G., 1946. Chem. Ind. (London), 347.
Herman, R.G. (Ed.), 1984. Catalytic Conversion of Synthesis Gas and Alcohols to Chemicals, Plenum Press,
New York.
Herman, R.G., 1991. In Guczi, L. (Ed.), In: New Trends in CO Activation, Elsevier, Chap. 7, p. 265.
Herman, R.G., Klier, K, Feeley, O.C. and Johansson, M.A., 1994. Prepr. Div. Fuel Chem. Am. Chem. Sot., 39
(4): 1141.
Hohenschutz, H., von Kutepow, N. and Himmle, W., 1966. Hydrocarbon Process., 45: 141.
Holbrook, M.T., 1993. In Kirk-Othmer Encyclopedia of Chemical Technology, Vol. 5, p. 1017.
Howard, M.J., Jones, M.D., Roberts, MS. and Taylor, S.A., 1993. Catal. Today, 325.
Hutchings, G.J. and Hunter, R., 1990. Catal. Today, 6: 279.
Hydrocarbon Processing, 1992. May, p. 35.
Hydrocarbon Processing, 1994. April, p. 112.
Ichikawa, M., 1982. Shokubai, 24: 212.
Jager, B., Dry, M.E., Shingles, T. and Steynber,, m A.P., 1990. Catal. Lett., 7: 293.
Jager, B., Kelfkens, R.C. and Steynberg, A.P., 1994. Natural Gas Conversion, Vol. II, p. 419.
Janaf, 1971. Thermodynamic Tables, US Department of Commerce, NBS, NSRDS-NBS, p. 37.
Jenner, G., 1989. Appl. Catal., 50: 99.
Jira C., Blau, W. and Grimm, D., 1976. Hydrocarbon Process., 97 (March).
Kagan, Yu.B., Liberov, LG., Slivinsky, E.V., Lockev, S.M., Lim, G.I., Rozovsky, A.Ya. and Bashkirov,
A.N., 1975. Dokl. Akad. Nauk SSSR, 222: 1093.
Katzer, J.R., Sleight, A.W., Gajardo, P., Michel, J.B., Gleason, E.F. and McMillan, S., 1981. Faraday Discuss.
R. Sot. Chem., p. 72.
Keim, W. (Ed.), 1983. In Catalysis in C, Chemistry, D. Reidel Publishing Co., Dordrecht, p. 103.
Keim, W., 1987. In Fahey, D.H. (ed.), Industrial Chemicals via C, Processes, ACS Symp. Ser., No. 328,
Chap. 1, p. 1.
Keim, W., 1989. I. Oganomet. Chem., 372: 15.
Keith, P.C., 1946. Oil Gas J., 45: 102.
Kellner, C.S. and Bell, A.T., 1981. J. Catal., 71: 288.
Kiennemann, A., Hinderman, J.P., Breault, R. and Idriss, H., 1987. In Fahey, D.H. (Ed.), industrial Chemicals
via C, Processes, ACS Symp. Ser., No. 328, p. 237.
Kinoshita, K. and Cairns, E.J., 1984. Kirk-Othmer Encyclopedia of Chemical Technology, Vol. 1 I. 4th Edn.,
p. 1098.
Klier, K., 1982. Adv. Catal., 3 1: 243.
Klier, K., Chatikavanij, V., Herman, R.G. and Simmons, G.W., 1982. J. Catat., 74: 343.
Klier, K., Young, C.W., Himelfarb, P.B. and Herman, R.G., 1986. J. Chem. Sot. Chem. Commun., 193.
Knifton, J.R., 1981a. US Patent 4,270,015.
Knifton, J.R., 1981b. Chemtech, I1 (October): 609.
I. Wrnder / Fuel Prowssing 7’whnology 48 (1996) 189-297 293

Knifton, J.R., 1981~. J. Mol. Catal., I I: 91.


Knifton, J.R., Lin, J.J., Storm, D.A. and Won& SF., 1993. Catal. Today, 18: 355.
Knowles, W.S., 1983. Act. Chem. Res., 16: 106.
Koelbel, H., 1957. Chem;big.-Tech., 29: 505.
Koelbel, H. and Englehardt, F., 1951. Brennst-Chem., 32: 150.
Koelbel, H. and Ralek, M., 1980. Catal. Rev. - Set. Eng., 21: 225.
Koermer, G.S. and Slinkard, W.E., 1978. Ind. Eng. Chem. Prod. Res. Dev., 17: 231.
Kollar, J., 1982. US Patent 4337,371.
Komatsu, M. and Yoneoka, M., 1988. Kogaku To Kogyo, 41: I 134.
Kung, H.H., 1980. Catal. Rev. - Sci. Eng., 31: 243.
Kuntz, E.G., 1987. Chemtech, I7 (September): 570.
Kuo, J.C.W., 1983. Slurry Fischer-Tropsch/Mobil two stage process of converting syngas to high octane
gasoline, Final Report, DOE Contract DE-AC22-80PC30022, June.
Kuo. J.C.W., 1984. In Cooper, B.R. and Ellingson, W.A. (Eds.), Gasification and Indirect Liquefaction
Science and Technology of Coal and Coal Utilization, Plenum Press, New York.
Kuo, J.C.W.. 1985. Two stage process for conversion of synthesis gas to high quality transportation fuels,
Final Report, DOE Contract DE-AC22-83PC60019, October.
Lamarre, L.. 1994. EPRI J.. 19 (July/August): 6.
Lay, K., 199.3. Hydrocarbon Process., (April): 35.
LeBlanc. J.R., Jr. and Rovner, J.M., 1990. Hydrocarbon Process., (March): 51.
Lee, S., 1990. Methanol Synthesis Technology, CRC Press, Boca Raton, FL.
Lee, J.S., Kim, J.C. and Kim, Y.G., 1990. Appl. Catal., 57: I.
Lee, S., Gogate, M.R. and Vijayaraghavan, P., 1992. Proceedings, 16th Annual EPRI Conference on Fuel
Science, 7-I, April.
Lewnard, J.J., Hsiung, T.H., White, J.F. and Brown, D.M., 1990. Chem. Eng. Sci., 45: 2735.
Lin, J.J. and Knifton, J.R., 1992. Chemtech, 22 (April): 248.
Liu, 2.. Tiemey, J.W.. Shah, Y.T. and Wender, I., 1988. Fuel Processing Technol., 18: 185.
Liu, Z., Tiemey, J.W., Shah, Y.T. and Wender, I., 1989. Fuel Processing Technol., 23: 149.
Lortnand, C., 1925. Ind. Eng. Chem., 17: 430.
Luft, G. and Schrod. M., 1983. J. Mol. Catal., 20: 175.
Machiels. C.J., Chowdhry, V., Taley, R.H., Ohuchi, R. and Sleight, A.W., 1984. In Herman, R.G. (Ed.),
Catalytic Conversions of Synthesis Gas and Alcohols to Chemicals, Plenum Press, New York, p. 413.
Mahajan, D. and Mattas, L., 1992. Proceedings of the 16th Annual EPRI Conference on Fuel Science, Palo
Alto, CA, April.
Marchionna, M., Lami, M., Ancilloti, F. and Ricci, R., 1988. Italian Patent 200,281A.
Marchionna, M.. Basini, L., Aragno, A., Lami, M. and Ancilloti. F., 1992. J. Mol. Catal., 75: 147.
Mauri, M.M., Romano, U. and Rivetti, F., 1985. Quad. Ing. Chim. Ital. 21: I.
McLamon. F.R. and Cairns, E.J., 1989. Annu. Rev. Energy, 14: 241.
Meisel, S.L.. I98 I. Philos. Trans. R. Sot. London, A300: 157.
Mills, G.A., 1977. Chemtech, 7 (July): 418.
Mills, G.A., 1988. Catalysts for Fuels from Syngas-New Directions for Research, IEA COM Research,
London, p. 109.
Mills, G.A., 1993. Status and future opportunities for conversion of synthesis gas to liquid energy fuels, Final
Report, National Renewable Energy Laboratory (NREL), US Department of Energy, Sub-Contract No.
DE-AC02-83CHl0093.
Mittasch, T. and Schneider, T., 1913. Gemran Patent 2,955,787.
Mobil Oil, 1975. US Patent X,894,102.
Mullen. A.. 1980. In Falbe, J. (Ed.), New Syntheses with Carbon Monoxide, Springer-Verlag, Berlin, p. 243.
Murchison, C.B. and Murdick, D.A., 1981. Hydrocarbon Process., (January): 156.
Nakamura, S., 1990. Chemtech, 20 (September): 556.
Natta, G., Columbo, U. and Pasquon, I., 1957. In Emmett, P.H. (Ed.), Catalysis, Vol. V.. Reinhold Publishing
Corp., New York, Chap. 3, p. 131.
Nijs, N.H., Jacobs, P.A. and Uytterhoeven, J.B., 1979. J. Chem. Sot. Chem. Commun., 180.
Nojtri, M. and Misono. M., 1993. Appl. Catal. A: Gen., 9.1: 103.
294 1. Wrndet/ Fuel Processing Technology 48 (1996) 189-297

Nonneman, L.E.Y. and Ponec, V., 1990. Catal. Len, 7: 213.


Notari, B., 1991. C-l Chemistry, A Critical Review, Catalytic Science and Technology, Vol. 1, Kodansha,
Ltd., Tokyo, p. 55.
Nunan, J.G., Bogdan, C.E., Klier, K., Smith K. J., Young, C.W. and Homan, R.G., 1988. J. Catal., 113: 410.
Oil and Gas Journal, 1996. (March 4): 43.
Onsager, O.T., 1984. Canadian Patent 1,175,798.
Orchin. M. and Rupilius, W., 1972. Catal. Rev., 85.
Orchin, M. and Wender, I., 1957. In Emmett, P.H. fed.), Reactions of Carbon Monoxide in Catalysis, Vol. 5,
Reinhold Publishing Corp., New York, Chap. 1, p. 1.
Orita, H., Naito, S. and Tamaru, K., 1984. J. Catal., 90: 183.
Owen, G., Hawkes, G.M., Lloyd, D., Lambert, R.M. and Nix, R.M., 1987. Appl. Catal., 33: 405.
Palekar, V.M., Jung, H., Tiemey, J.W. and Wender, I., 1993a. Appl. Catal. A: Gen., 102: 13.
Palekar, V.M., Tiemey, J.W. and Wender, I., 1993b. Appl. Catal. A: Gen., 103: 105.
Papp, H. and Baems, M., 1991. In Guczi, L. (Ed.), New Trends in CO Activation, Elsevier, Amsterdam, Chap.
10.
Parshall, G.W. and and Ittel, SD., 1992. Homogeneous Catalysis, 2nd Edn., John Wiley & Sons, New York,
p. 106.
Patart, M., 1921. French Patent 540,343.
Paul&, F.E., 1973. US Patent 3,769,329.
Paul& F.E. and Roth, J.F., 1968. J. Chem. Sot. Chem. Commun., 1578.
Peeples, J.E., 1991. Fuel Reformulation, 1: 27.
Peltzman, A., 1984. In Herman, R.G. (Ed.), Catalytic Conversion of Synthesis Gas and Alcohols to Chemicals,
Plenum Press, New York, p. 249.
Piacenti, F. and Bianchi, M., 1962. Ref. 2b, p. 18 in Bhattacharyya, S.K. and Pali, S.K., J. Appl. Chem., 12:
174.
Piacenti, F. and Bianchi, M., 1977. In Wender, 1. and Pino, P. (Eds.), Organic Syntheses via Metal Carbonyls,
Vol. 2, John Wiley & Sons, New York, p. 13.
Pichler, H., 1952. Adv. Catal., 4: 272.
Pichler, H. and Schulz, H., 1970. Chem.-lng.-Tech., 42: 162.
Pichler, H. and Ziesecb, K.H., 1949. Brennst-Chem., 30: 13.
Piel, W.J., 1993. Presented at the First Biomass Conference of the Americas, Burlington, VT, August.
Piel, W.J., 1994. Fuel Reformulation, 4 (2, March/April): 28.
Pierantozzi, R., 1993. Kirk-Othmer Encyclopedia of Chemical Technology, Vol. 5, 4th Edn., John Wiley &
Sons, New York, p. 97.
Pino, P., Piacenti, F. and Bianchi, M., 1977a. In Wender, 1. and Pino, P. (Eds.), Organic Syntheses via Metal
Carbonyls, Vol. 2, Wiley-Interscience, New York, p. 43.
Pino, P., Piacenti, F. and Bianchi, M., 1977b. In Wender, 1. and Pino, P. (Eds.), Organic Syntheses via Metal
Carbonyls, Vol. 2, John Wiley & Sons, New York, p. 233.
Pinto, A. and Rogerson, P.L., 1977. Chem. Eng., 102.
Ponec. V., 1984. In Gorbaty. M.L., Larsen, J.W. and Wender, 1. (Eds.), Coal Science, Vol. 3, Academic Press,
Orlando, p. 1.
Porcelli, R. and Juran, B.. 1986. Hydrocarbon Process., (March): 37.
Poutsma, M.L., Elek, L.F., Ibarbia, P.A., Risch, A.P. and Rabo, J.A., 1978. J. Catal., 52: 157.
Prentice, G., 1984. Chemtech, 14 (November): 684.
Pretzer, W.R. and Kobylinski, T.P., 1980. Ann. N. Y. Acad. Sci., 333: 58.
Pruett, R.L., 1981. Science, 211: II.
Pruett, R.L. and Kacmarik, R.T., 1982. Organometallics, I: 1693.
Pruett, R.L. and Smith, J.A., 1969. J. Org. Chem., 34: 327.
Pruett, R.L. and Smith, J.A., 1970. US Patent 3.527.809.
Pruett, R.L. and Walker, W.E., 1974. US Patent 3.833.634.
Raab, C., Lercher. J., Goodwin, J. and Shyu, J., 1990. J. Catal., 122: 406.
Rao, U.V.S. and Gormley, R.J., 1990. Catal. Today, 6: 207.
Rathke, J.W. and Feder, H.M., 1978. J. Am. Chem Sot., 100: 3623.
Reppe, W., 1953. Justus Liebigs Ann. Chem., 582: 1.
I. Wender / Furl Processin,g Technology 48 (I 996) 189-297 2%

Reppe, W. and Vetter, H., 1953. Justus Liebigs Ann. Chem., 582: 133.
Reppe, W., von Kutepow, N. and Bille, A., 1961. US Patent 3,014,962.
Research Association for Cl Chemistry, 1989. Progress in Cl Chemistry in Japan, Kodansha, Tokyo-Elsevier.
Amsterdam.
Rizkalla, N., 1987. In Fahey, D.R. (Ed.), ACS Symp. Ser.. No. 328, p. 61.
Rizkalla, N. and Goliaszewski, A., 1987. In Fahey, D.R. (Ed.), Am. Chem. Sot. Symp. Ser., No, 328, p. 136.
Roberts, R.W., Lim, P.K., McCutcheon, M.S. and Mawson, S., 1992. Preprint Am. Chem. Sot.. Division of
Petroleum Chemistry, Inc., 37 (I): 225.
Robinson, R.C., 1965. US Patent 3,190,912.
Roelen, O., 1938. German Patent 103,362.
Roelen, 0.. 1943. US Patent 2327,066.
Rofer-DePoorter. C.K., 1981. Chem. Rev., 8 I: 447.
Romano, U., Tesei, R., Marcello, M.M. and Pierluigi, R., 1980. Ind. Eng. Chem. Prod. Res. Dev., 19: 395.
Roper, M., 1983. In Keim, W. (Ed.), Catalysis in C, Chemistry, Reidel Publishing Co., Boston, MA, p. 41.
Roper, M., 1991. In Guczi, L., (Ed.), New Trends in CO Activation, Elsevier, Amsterdam, Chap. 9, p. 382.
Roper, M., Elvevoll, E.O. and Lutgendorf, M., 1985. Erdoel Kohle, Erdgas, Petrochem., 1: 38.
Rostrop-Nielsen, J.R., 1994. Catal. Today, 21: 257 and references therein.
Roth. J.R., Craddock, J.H., Hershman, A. and Paul&, F.E., 1971. Chemtech, 1 (October): 600.
Rozovskii, A. Ya., 1980. Kinet. Katal., 21: I.
Sabatier, P. and Senderens, J.B., 1902. Nouvelles syntheses du methane. C. R. Hebd. Seances Acad. Sci., 134:
514.
Sakai, T. and Kunugi, T.. 1974. Sekiyu Gokkai Shi, 17: 853.
Sapienza, R.A., Slegeir, W.A., O’Hare, T.W. and Mahajan, D., 1986. US Patent 4,614,749.
Satterfield, C.N., 1991. Heterogeneous Catalysis in Practice, 2nd Edn., McGraw-Hill, New York, p. 446.
Schlesinger, M.D., Crowell, J.H., Leva, M. and Starch, H.H., 1951. Ind. Eng. Chem.. 43: 1474.
Schreck, D.J., Busby, D.C. and Wegman, R.W., 1988. J. Mol. Catal., 47: 117.
Schulz, G.V., 1935. 2. Phys. Chem., 30: 379.
Schulz, H., 1985. C, Mol Chem., 1: 23 1.
Seddon, D., 1992. Catal. Today, 15: 1.
Shah, Y.T. and Perrotta, A.J., 1976. hid. Eng. Chem. Prod. Res. Dev., 15: 123.
Shamsi, A., Rao, V.US, Gormley, R.J., Obermeyer, R.T., Schehl, R.R. and Stencel, J.M., 1986. Appl. Catal.,
27: 55.
Sheldon, R.A., 1983. Chemicals from Synthesis Gas, D. Reidel Publishing Co., Dordrecht, p. 156.
Sheldon, R.A., 1992. In Sawyer, D.T. and Martell, A.E. (I%.), Industrial Environmental Chemistry, Plenum
Press, New York, p. 99.
Sheldon, R.A.. 1993. In Weijnen. M.P. and Drinkenburg, A.A.H. (Eds.1, Precision Process Technology,
Kluwer, Amsterdam, p. 125
Sheldon, R.A., 1994. Chemtech, 24 (March): 38.
Sherwin, MS. and Frank, M.E., 1976. Hydrocarbon Process., 55: 122.
Shibata, M., Ohbayashi, Y., Kawata, N., Masumoto, T. and Aoki, K., 1984. J. Catal., 95: 296.
Sic, ST., Senden. M.M.G. and Wechem, H.M.H., 1991. Catal. Today, 8: 371.
Sinor, J.E., Consultants Inc., 1990. The Clean Fuels Report.
Slaugh. L.H., 1983. US Patent 4.375,424.
Slaugh, L.H. and Mullineaux, R.D., 1968. J. Organomet. Chem., 13: 469.
Slegier, W., Sapienza, R.. O’Hare, T., Mahajan, D., Foran, M. and Skaperdas, G., 1984. 9th EPRI Contractors
Meeting, Palo Alto, CA, May.
Smith, K.J. and Anderson, R.B., 1983. Can. J. Chem. Eng., 61: 40.
Solianos, A., 1992. Catal. Today. 15: 149.
Somorjai, G.A., 1981. Catal. Rev. Sci. Eng., 23: 189.
Spencer. D.F., Gluckman, M.J. and Alpert, S.B., 1982. Science, 215: 1571.
Spencer, D.F., Alpert, S.G. and Gilman, H.H., 1986. Science, 232: 609.
Stiles, A.B., 1977. AlChE J., 23: 362.
Stiles, A.B., Chen, F., Harrison, J.B., Hu, X.-D., Storm, D.A. and Yang, H.X., 1991. Ind. Eng. Chem. Res..
30: 811.
296 I. Wmrukr/Fuel Processing Technology 48 (1996) 189-297

Starch, H.H., Golumbic, N. and Anderson, R.B., 195 I. The Fischer-Tropsch and Related Syntheses, John
Wiley & Sons, New York
Stull, D.R., Westrum, E.F., Jr. and Sinke, G.C., 1969. The Chemical Thermodynamics of Organic Compounds,
John Wiley & Sons, New York.
Summer, C.E. and Zoeller, J.R., 1993. In Agreda, V.H. and Zoeller, J.R. (Eds.), Acetic Acid and Its
Derivatives, Marcel Dekker, New York, Chap. 12, p. 225.
Sunset, T., Sogge, G. and Strom, T., 1994. Catal. Today, 21: 269.
Supp, E., 1981. Hydrocarbon Process., (March): 71.
Tabak, S.A and Yurchak S., 1990. Catal. Today, 6: 307.
Tijm, P.J.A., Wechem, H.M.H. and Senden, M.M.G., 1993. The Shell middle distillate synthesis project. New
opportunities for marketing natural gas, presented at Alternate Energy ‘93, Colorado Springs, April 27-30.
Toepel, T., 1964. Chim. Ind. (Paris), 91: 139.
Tonner, S.P., Wainwright, M.S., Trimm, D.L. and Cant, N.W., 1983. J. Mol. Catal., 18: 215.
Topp-Jorgensen, J., 1988. In Bibby, D.M., Chang, C.D., Howe, R.F. and Yurchak, S. @Is.), Topside
Integrated Gasoline Synthesis-The TIGAS Process, Methane Conversion, Elsevier, Amsterdam, p. 293.
Topp-Jorgensen, J. and Rostrup-Nielsen, J.R., 1986. Oil Gas J., (May): 68.
Torrence, G.P., Hendricks, J.D. and Aquilb, A., 1991. US Patent 5,026,908.
Trimm, D.L. and Wainwright, M.S., 1990. Catal. Today, 6: 261.
Tucci, E.R., 1970. Ind. Eng. Chem. Prod. Res. Dev., 9: 516.
Unzelman, G.H., 1992. Fuel Reformulation, 2: 16.
van der Bugt, M.J., van Klinken, J. and Sie, ST., 1990. The Shell middle distillate synthesis process. Inst.
Pet. Rev., (April): 45.
van der Lee, G., Schuller, B., Post, H., Favre, T.L. and Ponec, V., 1986. J. Catal., 98: 522.
Vannice, M.A., 1975. J. Catal., 37: 449.
Vannice, M.A., 1977. J. Catal., 50: 228.
Vannice, M.A., 1982. In Anderson, J.R. and Boudart, M. (Eds.), Catalysis: Science and Technology, Vol. 3,
Springer-Verlag, Berlin, p 139.
van Rensberg, S.T.J., 1990. The Mossgas to fuels process, CHEMSA, a SAICHE paper, February.
Velocci, A.L., 1991. Lamp, 73: 14.
Wade, L.E., Gengelbach, R.B., Trumbley, J.L. and Hallbauer, W.L., 1981. Kirk-Othmer Encyclopedia of
Chemical Technology, Vol. 15, 3rd Edn., John Wiley & Sons, New York, p. 398.
Wagmann, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halow, I., Bailey, S.M., Chumey, K.L. and
Nuttal, R.L., 1982. NBS tables of chemical thermodynamic properties. Selected values for inorganic and
C, and C, organic substances in S.1 units. J. Phys. Chem. Ref. Data, 1 I (Suppl. 2): 84 and 92.
Wakamatsu, H., Uda, J. and Yamakami, N.J., 1971. Chem. Sot. Chem. Commun., 1540.
Walker, B.V., 1985. Synthetic and alternate fuels production in New Zealand. Proc. EPRI Conf. on Coal
Gasification Systems and Synthetic Fuels for Power Generation, Vol. 2, Palo Alto, CA, p. 36-1.
Waller, F.J., 1985. J. Mol. Catal., 31: 123.
Watanabe, H., Mitsuda, T.. Yasunori, J., Kiguchi, J. and Takegami, Y., 1979. Bull. Chem. Sot. Jpn., 52: 2735.
Watson, P.R. and Somorjai, G.A., 1981. J. Catal., 72: 347.
Waugh, K.C., 1992. Catal. Today, 15: 51.
Wedaa, H.W., 1994. Proc. hit. Conf. on Fuel Cells, South Coast Air Quality Management District (Hosts),
Long Beach, CA, February 23-24.
Weisz, P.B. and Frillette, V.J., 1960. J. Phys. Chem., 64: 382.
Weisz, P.B., Frillette, V.J., Maatman, R.W. and Mower, E.B., 1962. J. Catal., I: 307.
Weissermel, K. and Arpe, H.-H., 1993. In Industrial Organic Chemistry, Vol. 1, 2nd Edn., Verlag Chemie,
Weinheim, p. 254.
Weitkamp, A.W. and Frye, C.G., 1953. hid. Eng. Chem., 45: 363.
Wender, I., 1976. Catal. Rev. - Sci. Eng., 14: 97.
Wender, 1.. 1984. Catal. Rev. Sci. Eng., 26: 363.
Wender, I. and Klier, K., 1989. Review of indirect liquefaction. Chap. 5 in Coal Liquefaction-a Research
Needs Assessment, DOE/ER-0400, Vol. II, Contract DE-ACOI87ER301 IO, Washington, D.C., US
Department of Energy.
1. Wen&r/ Fuel Processing Technology 48 (19961 189-297 297

Wender, I. and Seshadri, K.S., 1984. In Coal liquefaction, Tech. Progress Rep., US Department of Energy
Grant PC/6000 54/74.
Wender, I., Levine, R. and Orchin, M., 1949. J. Am. Chem. Sot., 71: 4160.
Wender, I., Freidel, R.A. and Orchin, M., 195 I. Science, I 13: 206.
Westerterp, K.R. and Kuczynski, M., 1986. Hydrocarbon Process., p. 80.
Westerterp, K.R., Bodewes, T.N., Vrijand, M.S.A. and Kuczynski. M., 1988. Hydrocarbon Process.. p. 69.
Westerterp, K.R., Kuczynski, M. and Kamphius, C.H., 1989. Ind. Eng. Chem. Res., 28: 763.
Whyman, R., 1985. In Jennings, J.R. (Ed.), Industrial Applications of Homogeneous Catalysts, Selected
Developments in Catalysis, Blackwell Scientific Publications, Oxford, p. 128.
Williamson, R.C. and Kobylinski, T.P., 1979. US Patent 4,170,605.
Worsham, P.R., 1994. Kirk-Othmer Encyclopedia of Chemical Technology, Vol. IO, 4th Edn., John Wiley &
Sons, New York, p. 325.
Xiaogding, X., Doegburg, E.B.M. and Scholten, J.J.F., 1987. Catal. Today, 2: 125.
Yakobson, D.L., 1996. Personal communication.
Yoshida, T., 1978. _%Monosodium glutamate, Kirk-Othmer Encyclopedia of Chemical Technology, Vol. 2.
3rd Edn., John Wiley & Sons, New York. p. 410.
Zoeller, J.R., 1993. In Agreda, V.H. and Zoeller, J.R. (Eds.), Acetic Acid and Its Derivatives, Marcel Dekker,
New York, Chap. 4, p, 35.
Zoeller, J.R., Agreda, V.H., Cook, S.L., Lafferty, N.L., Polichnowski, S.W. and Pond, D.M., 1992. Catal.
Today, 13: 17.

You might also like