Sie sind auf Seite 1von 35

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258685985

Fischer–Tropsch Synthesis: Catalysts and Chemistry

Chapter · August 2013


DOI: 10.1016/B978-0-08-097774-4.00729-4

CITATIONS READS

56 7,910

10 authors, including:

Jan van de Loosdrecht F.G. Botes


Sasol Suid Afrikaanse Steenkool en Olie
61 PUBLICATIONS   1,700 CITATIONS    19 PUBLICATIONS   532 CITATIONS   

SEE PROFILE SEE PROFILE

Ionel Mugurel Ciobîcă Alta C Ferreira

34 PUBLICATIONS   1,205 CITATIONS   
Suid Afrikaanse Steenkool en Olie
9 PUBLICATIONS   129 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Cobalt FTS catalyst deactivation and regeneration View project

Olefin metathesis of industrially relevant feedstocks View project

All content following this page was uploaded by Ionel Mugurel Ciobîcă on 29 September 2016.

The user has requested enhancement of the downloaded file.


This article was originally published in the Comprehensive Inorganic Chemistry II, published by
Elsevier, and the attached copy is provided by Elsevier for the author's benefit and for the benefit of
the author's institution, for non-commercial research and educational use including without
limitation use in instruction at your institution, sending it to specific colleagues who you know, and
providing a copy to your institution’s administrator.

All other uses, reproduction and distribution, including without limitation commercial reprints,
selling or licensing copies or access, or posting on open internet sites, your personal or
institution’s website or repository, are prohibited. For exceptions, permission may be sought for
such use through Elsevier's permissions site at:

http://www.elsevier.com/locate/permissionusematerial

van de Loosdrecht J., Botes F.G., Ciobica I.M., Ferreira A., Gibson P., Moodley D.J., Saib A.M.,
Visagie J.L., Weststrate C.J. and Niemantsverdriet J.W. Fischer–Tropsch Synthesis: Catalysts and
Chemistry. In: Jan Reedijk and Kenneth Poeppelmeier, editors. Comprehensive Inorganic Chemistry
II, Vol 7. Oxford: Elsevier; 2013. p. 525-557.
Author's personal copy

7.20 Fischer–Tropsch Synthesis: Catalysts and Chemistry


J van de Loosdrecht, Sasol Technology Pty (Ltd), Sasolburg, South Africa; Eindhoven University of Technology, Eindhoven,
The Netherlands
FG Botes, Sasol Technology Pty (Ltd), Sasolburg, South Africa
IM Ciobica, Eindhoven University of Technology, Eindhoven, The Netherlands
A Ferreira, P Gibson, DJ Moodley, AM Saib, and JL Visagie, Sasol Technology Pty (Ltd), Sasolburg, South Africa
CJ Weststrate and JW (Hans) Niemantsverdriet, Eindhoven University of Technology, Eindhoven, The Netherlands
ã 2013 Elsevier Ltd. All rights reserved.

7.20.1 Introduction: Processes, Catalysts, and Recent History 526


7.20.1.1 ‘Anything’-to-liquids Technology: Syngas Production, FTS, and Product Workup 526
7.20.1.2 FTS, the Product Distribution 526
7.20.1.3 Fischer–Tropsch Catalysts and Modes of Operation 527
7.20.1.4 Historical Development of the FTS 529
7.20.2 Iron-Based FTS Catalysts 531
7.20.2.1 Introduction 531
7.20.2.2 Commercial Applications 531
7.20.2.3 Iron Fischer–Tropsch Catalyst Preparation 531
7.20.2.3.1 Fusion 532
7.20.2.3.2 Precipitation 533
7.20.2.3.3 Improving iron Fischer–Tropsch catalyst precursors by promoters 534
7.20.2.3.4 Activation and reduction procedures 534
7.20.2.4 Selectivity Manipulation of Iron Catalysts 535
7.20.2.5 Catalyst Stability During FTS 535
7.20.2.6 Spent Catalyst Management 536
7.20.3 Cobalt-Based FTS Catalysts 537
7.20.3.1 Introduction 537
7.20.3.2 Composition of Cobalt Catalysts 537
7.20.3.3 Preparation of Cobalt Fischer–Tropsch Catalysts 539
7.20.3.3.1 Precipitation 539
7.20.3.3.2 Preparation methods involving pre-shaped supports 540
7.20.3.3.3 Calcination 541
7.20.3.3.4 Reduction 541
7.20.3.4 Cobalt Catalyst Fischer–Tropsch Performance 541
7.20.3.5 Deactivation and Regeneration of Cobalt Fischer–Tropsch Catalysts 544
7.20.4 Mechanisms and Kinetics of FTS Over Iron and Cobalt Catalysts 546
7.20.4.1 Introduction 546
7.20.4.2 Surface Science Studies and Model Reactions 547
7.20.4.2.1 Adsorption of CO and hydrogen on model surfaces 547
7.20.4.2.2 C–O bond scission 548
7.20.4.2.3 Hydrogenation and the stability of C1Hx species 549
7.20.4.3 DFT Modeling 549
7.20.4.3.1 CO Dissociation 549
7.20.4.3.2 C þ H reactions 550
7.20.4.3.3 Chain growth 551
7.20.4.4 Macrokinetic Observations and Models 551
7.20.4.4.1 General observations regarding kinetics 551
7.20.4.4.2 Simple macrokinetic models 552
7.20.4.4.3 Selectivity modeling 552
7.20.4.5 Mechanistic and Kinetic Implications 553
7.20.5 Conclusion 554
References 554

Comprehensive Inorganic Chemistry II http://dx.doi.org/10.1016/B978-0-08-097774-4.00729-4 525


Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
526 Fischer–Tropsch Synthesis: Catalysts and Chemistry

7.20.1 Introduction: Processes, Catalysts, gasification. Alternatively, heat can be supplied externally, in
and Recent History which case the gas is reformed only with steam and/or CO2,
but no oxygen is added. Examples of this approach include
The Fischer–Tropsch synthesis (FTS) represents technology steam reforming (where mainly steam is added), dry reforming
from the 1920s1,2 that has continuously been revived to pro- (where mainly CO2 is added), and heat exchange reforming
vide synthetic hydrocarbon fuels and chemicals from initially (where process heat is supplied to the reformer tubes). Typical
coal, later natural gas, and nowadays also biomass. Virtually reforming catalysts are based on nickel as the active metal.10
any source of (hydro)carbon feedstock can be converted to a The second step in the XTL process is to catalytically convert
mixture of synthesis gas, or syngas (CO and H2), which is in the syngas to a range of hydrocarbons via the FT synthesis,
fact a key intermediate on which theoretically the entire chem- which mainly yields linear alkanes and 1-alkenes, and which
ical industry could be based. FTS stands for the reaction(s) of will be the main subject of this chapter hereafter.
synthesis gas to predominantly straight-chain hydrocarbons, The third and last step is usually the workup of the
which can be paraffins from CH4 to waxes (CnH2nþ2 with n hydrocarbons to final products, which are typically fuels, but
from 1 to over 100), olefins from ethylene to much longer optionally also chemicals. A popular application at present is
molecules (CnH2n, with n  2), and to a lesser extent oxygen- to target the production of long chain waxes in the FT synthe-
ated products such as alcohols. It produces as main by- sis, followed by hydrocracking to middle distillate range
products water and/or carbon dioxide, that is, due to the components, such as diesel (C9–C22) and jet fuel (C9–C15).
water-gas shift (WGS) reaction. Being a highly exothermic Hydrocracking catalysts are bifunctional in nature, with either
reaction, it generates large amounts of heat. The process is a noble metal (e.g., Pt) or sulfided base metals (e.g., Ni/W or
represented by the simplified reaction equations Co/Mo) as the hydrogenation function on a catalytically active
acidic support, such as a silica–alumina. We refer to the liter-
FTS : CO þ 2H2 ! CH2  þH2 O  165 kJ mol1 [1] ature for further information on this subject.11,12
WGS : CO þ H2 O⇄H2 þ CO2  42 kJ mol1 [2]

Reaction [1] represents in essence a polymerization, imply-


7.20.1.2 FTS, the Product Distribution
ing that the product will be a mixture of hydrocarbons with
a distribution in molecular weights. Selectivity and control At the chemistry level, the FT synthesis is both a CO hydroge-
thereof are therefore of key importance in FTS technology. nation reaction and a polymerization reaction. The former is
Fischer–Tropsch technology represents a subject of inten- reflected by the fact that the C–O bond must be broken and
sive research both in industry and in academia. Many excellent new C–H bonds formed. Additionally, C–C bonds must be
reviews are available.3–9 formed in order to effect hydrocarbon chain growth. Since the
In this chapter, we first describe the general aspects of the product carbon number distribution approximately follows a
technology in which the FTS features, then the more chemical statistical function called the Anderson–Schulz–Flory relation-
aspects of the process in relation to the iron and cobalt catalysts ship, it is widely accepted that chain growth occurs one carbon
that are used in practical applications, and finally mechanistic atom at a time via a polymerization mechanism. Proposals for
insight, on the basis of kinetics, surface science, and computa- the monomer of chain growth, which is produced in situ, have
tional modeling. included adsorbed CO, an enol species and a CHx species,7,13–15
and will be discussed further in the section on mechanism and
kinetics.
The competition between chain growth (yielding a surface
7.20.1.1 ‘Anything’-to-liquids Technology: Syngas
intermediate with one higher carbon number) and chain ter-
Production, FTS, and Product Workup
mination (yielding a desorbed final product) is determined by
The overall process from original carbon source for the syngas the probability for growth, called the a-value. A higher a-value
to the FTS product is named after the feedstock employed, will result in longer hydrocarbons and thus a heavier product
hence the terminology ‘coal-to-liquids’ (CTL), ‘gas-to-liquids’ spectrum (Figure 1). If a is independent of carbon number, the
(GTL) and ‘biomass-to-liquids’ (BTL), collectively known as scheme presented in Figure 2 applies and the total amount of
XTL (‘anything’-to-liquids). carbon contained in products with n carbon atoms (namely
In all instances, the carbon source is first converted to Cn) can be formulated on a relative basis:
synthesis gas (or ‘syngas’ for short), which is a mixture of CO
C1 ¼ 1ð1  aÞ
and H2. Solid feedstocks such as coal or biomass are gasified,
usually noncatalytically, by partial oxidation with oxygen (sup- C2 ¼ 2ð1  aÞa
plying the heat for the endothermic gasification reactions) and
reaction with steam (which acts as a gasification agent, hydro- C3 ¼ 3ð1  aÞa2
gen source, and coolant). Cn ¼ nð1  aÞan1
When the starting material is natural gas, it can also be
adiabatically reformed in the presence of oxygen and steam. The total amount of carbon in the product spectrum then
There are different embodiments of this approach, such as forms a convergent infinite sum with an analytical solution:
autothermal reforming (ATR), noncatalytic partial oxidation X
1 X
1
1
(POX), and catalytic partial oxidation (CPOX), but in essence Cn ¼ nð1  aÞan1 ¼
the chemistry of all is the same and very similar to that of coal 1 1
1a

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 527

HTFT LTFT
100
90
CH4 C5+

Carbon atom selectivity (%)


80
70 C20+
waxes
60 C2–C4
C5–C11
50 gasoline

40 C9–C22
30 diesel-
distillates
20
10
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Chain-growth probability, a

Figure 1 Hydrocarbon product spectrum that is produced during Fischer–Tropsch synthesis for varying chain growth probability a. High-temperature
Fischer–Tropsch technology (HTFT) corresponds approximately to 0.70 < a < 0.75, and low-temperature Fischer–Tropsch (LTFT) to about
0.85 < a < 0.95.

Products are primary FT products and which are formed subsequently by


C1 C2 C3 secondary reactions. For example, olefins and alcohols can
1-a 1-a 1-a undergo a range of secondary reactions, such as hydrogenation,
a a a double bond isomerization, skeletal isomerization, and con-
C* → C1* → C 2* → C3* → …..
version to heavier compounds.4,16
Intermediates

Figure 2 Carbon chain growth and termination scheme for the


derivation of the Anderson–Schulz–Flory equation, with a the chain
growth probability factor, Cn (n ¼ 1, 2, 3, . . .) the final products with n 7.20.1.3 Fischer–Tropsch Catalysts and Modes of Operation
carbon atoms, and Cn* the intermediates with n carbon atoms. Metals known to catalyze the FT reaction mainly include iron,
cobalt, ruthenium, and nickel.6 Ruthenium is a scarce and
expensive metal, whereas nickel only forms methane at reac-
This means that the selectivity toward products with n tion temperatures sufficiently high to suppress nickel carbonyl
carbon atoms on a carbon atom basis, namely Sn, can be formation (note that methanation is the reverse reaction of
expressed as follows: methane reforming, for which nickel-based catalysts are com-
monly used10). As a result, only iron- and cobalt-based FT
Cn catalysts have found commercial application.18–20 Iron gener-
Sn ¼ X1 ¼ nð1  aÞ2 an1
1
Cn ally produces more olefins and oxygenates than cobalt (i.e., a
less hydrogenated product spectrum), which may be related to
After converting this equation to the logarithmic domain
the lower hydrogenating ability of iron. While cobalt is active
and rearranging, it is found that
in the metallic state,19 iron catalysts change under Fischer–
  Tropsch conditions to a complex mixture of iron carbides
Sn ð1  aÞ2
ln ¼ n ln a þ ln [3] and oxides.20,21
n a
Byproducts of the FTS originate from the way oxygen from
As a result, a plot of ln(Sn/n) versus carbon number (n) gives CO is removed. With cobalt catalysts, essentially all oxygen
rise to a straight line with a slope equal to ln(a). However, from CO dissociation (typically around 99%) is discarded as
deviations in the actual FT product spectrum from the ideal water. Iron catalysts differ in this respect, as a significant por-
Anderson–Schulz–Flory distribution are usually observed.4,14,16,17 tion of the oxygen is also discarded as CO2. The latter is often
These include a higher methane and a lower C2 selectivity than visualized as a separate, consecutive reaction, namely the WGS.
predicted by the equation. There is also an increase in the chain Stoichiometrically, the overall process can be represented by
growth probability factor and concomitant decrease in the reactions [1] and [2], which we repeat here:
olefin/paraffin ratio with hydrocarbon chain length. In addi-
tion to linear alkanes and 1-alkenes, a variety of other products FTS : CO þ 2H2 ! CH2  þH2 O  165 kJ mol1 [1]
are also formed, including branched aliphatic compounds,
WGS : CO þ H2 O⇄H2 þ CO2  42 kJ mol1 [2]
alcohols, aldehydes, ketones, acids, and (at sufficiently high
operating temperatures) even aromatics. This alludes to the The net rate of hydrogen conversion divided by CO conver-
complexity of the reaction and many unresolved issues remain sion (sometimes referred to as the ‘usage ratio’) is extremely
regarding the reaction mechanism. A further complicating important for the gas loop design around an FT reactor. In the
factor is that it is not always clear which of these compounds one extreme, where virtually no WGS takes place, the usage

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
528 Fischer–Tropsch Synthesis: Catalysts and Chemistry

ratio is only determined by the FT reaction (with a high selec- catalysts, yielding a much lower overall usage ratio that is better
tivity to long hydrocarbons and a low selectivity to methane) suited to syngas feeds with a low H2/CO ratio, such as that
and assumes a value of around 2. In the other extreme, where generally obtained from coal gasifiers.
almost all water is shifted to CO2, the usage ratio can approach Table 1 presents the current commercial application of the
a value of 0.5. The low propensity of cobalt catalysts for the FTS. There are two important aspects to note from this table.
WGS makes them the preferred catalysts for GTL application, First, the worldwide FTS capacity is expected to reach a total of
since the H2/CO ratio of syngas derived from natural gas is just over 400 000 barrels per day by 2013 (1 barrel ¼ 159 l),
already close to or above the usage ratio. Any additional WGS which is very small compared to the total crude oil production
will result in an excess of hydrogen that would not be fully of around 80–85 million barrels per day. Second, the FTS has
consumed by the FT reaction, even if CO is converted to been applied in a variety of forms, which determines the type
extinction. Conversely, the WGS is more facile over iron of reactor employed. As indicated in Figure 3, these reactors

Table 1 Fischer–Tropsch synthesis, current commercial plants and plants under construction

Company Location Carbon feedstock Catalyst type Reactor type Start-up date Approximate plant
capacity (barrels
per day)

Sasol Sasolburg, South Initially coal, currently Fused Fe/K HTFT circulating 1955 to
Africa natural gas fluidized bed 1985
Precipitated Fe/K LTFT multitubular 1955
fixed bed 5000
Precipitated Fe/K LTFT slurry phase 1993
(spray dried)
Sasol Secunda, South Mostly coal, now Fused Fe/K HTFT circulating 1980–1999 160 000
Africa supplemented by fluidized bed
natural gas HTFT SAS reactora 1995
Shell Bintulu, Malaysia Natural gas Co/SiO2 LTFT multitubular 1992 14 500
Co/TiO2 fixed bed
PetroSA Mosselbay, South Natural gas Fused Fe/K HTFT circulating 1993 22 000
Africa fluidized bed
(Sasol technology)
Sasol-QP (Oryx) Ras Laffan, Qatar Natural gas Co/Al2O3 LTFT slurry phase 2007 34 000
Shell (Pearl) Ras Laffan, Qatar Natural gas Co/TiO2 LTFT multitubular 2011 140 000
fixed bed
Chevron-Sasol Escravos, Nigeria Natural gas Co/Al2O3 LTFT slurry phase 2013 34 000
a
SAS: Sasol Advanced Synthol, fixed fluidized bed.

Moving bed Stationary bed reactors


1–200 μm particles 200–5000 μm particles

Multitubular
LTFT, 200–250 ∞C
fixed bed
• 3-phase system:
6000 barrels
gas–liquid–solid
per day
• a = 0.85–0.95
• Products: Slurry bubble
wax, diesel, naphta column, Microchannel reactor
• Catalysts: filled with wax ~ 200–1000 b/d (for assembly)
supported cobalt or 24 000 barrels Microchannel process
technology module
precipitated iron per day
Boiling heat transfer

HTFT, 320–350 ∞C High heat flux


10 times higher heat flux
FT

than conventional reactors


• 2-phase:gas–solid
• a = 0.70–0.75
• Products:
petrol and chemicals
• Catalysts:
fused iron, K-promoted
Circulating
fluid bed Fixed fluid bed
7000 barrels per day 20 000 barrels per day

Figure 3 Overview of Fischer–Tropsch technology with reactors (figure microchannel reactor: courtesy of the Oxford Catalysts Group).

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 529

can broadly be classified in two classes, namely two-phase or promise, especially with respect to the small-scale application
three-phase reactors, and moving or stationary catalyst bed of a few hundred or a few thousand barrels per day production
reactors.22–24 capacity, and some relatively new commercial companies are
The high-temperature Fischer–Tropsch (HTFT) synthesis actively pursuing this technology.28
process is characterized by operating temperatures of about Structured reactors (monolith type reactors) for FT applica-
320–350  C and the products are essentially only in the gas tion have also attracted the attention of mainly academia,
phase under reaction conditions, giving rise to a gas–solid although there has been some limited interest from commer-
system without any bulk liquid phase. Originally, this process cial companies as well.29 In this approach, the active FT metal
was operated in circulating fluidized bed reactors and more (e.g., cobalt) is coated onto a large structure with a specific
recently in fixed fluidized bed reactors. Cobalt catalysts would geometry, which is then inserted into a reactor tube.
essentially only produce methane at these temperatures, mak- The LTFT synthesis is ideally suited for the production of
ing alkali-promoted iron catalysts the only option for this high-quality middle distillates (diesel and jet fuel) after hydro-
application. Due to the mechanical demand that these moving cracking of the long chain waxes. In addition, the heavy prod-
bed reactors place on the catalyst, particle strength is an impor- uct spectrum provides chemical opportunities in the form of
tant consideration; consequently, only fused bulk iron catalysts speciality waxes and base oils. The naphtha from the process is
have been employed commercially. The light product spectrum also a high-quality feedstock for naphtha steam crackers that
is best suited to the production of gasoline, but the high produce mainly ethylene, but also some propylene.
selectivity toward linear 1-olefins and (to a lesser extent) oxy-
genates allows for the extraction of chemicals from the product
7.20.1.4 Historical Development of the FTS
slate. These include monomers such as ethylene and propyl-
ene, co-monomers such as 1-hexene and 1-octene, and sol- Historically, the first syngas conversion results were pub-
vents (e.g., propanol, butanol, methyl ethyl ketone (MEK), lished by Sabatier30 in 1902 where it was shown that a mix-
and acetaldehyde).25 ture of carbon monoxide and hydrogen could be converted
The low-temperature Fischer–Tropsch (LTFT) synthesis is into methane over nickel and cobalt catalysts. In the 1920s,
operated between 200 and 250  C.26,27 Both cobalt and iron Franz Fischer and Hans Tropsch took this process a step
catalysts are suitable for this application, although cobalt cat- further and showed that syngas could be converted into a
alysts would typically be used toward the lower half of the mixture of higher hydrocarbons that could be used as petrol
quoted temperature range. The heavy product spectrum or diesel (i.e., FTS).1,2 In their first patent,31 they described
extends well into the domain of waxes, which are liquid the production of higher hydrocarbons using iron- and
under reaction conditions. The presence of a bulk liquid phase cobalt-based catalysts operated at atmospheric pressure and
gives rise to a three-phase gas–liquid–solid system. Originally, at temperatures below 300  C. Further research in Germany
only fixed-bed reactors operating in a trickle bed mode were led to improved versions of this process. The first commercial
employed for this synthesis. In order to limit the pressure plant started in 1936. Several others followed and provided
drop over the stationary catalyst bed, catalyst particle sizes Germany and Japan with synthetic fuel during the Second
must be in the millimeter range, which brings about signifi- World War. These plants used mainly cobalt catalysts
cant intra-particle diffusion limitations. This not only limits supported on kieselguhr (i.e., silica-based supports) and pro-
catalyst utilization, but also adversely affects product selectiv- moted by magnesia and thoria, in fixed-bed reactors. Further,
ities due to the differences in diffusion rates between hydro- China had FTS plants in the 1940 through 1960s, all based on
gen and CO that causes higher H2/CO ratios toward the cobalt catalysts.32
center of the particles. The highly exothermic nature of After the war, the German FT technology came in the hands
the FT reaction causes axial and radial temperature profiles of the Allied Forces. Many scientists and engineers who con-
in the catalyst bed. tributed to the German developments were interrogated and
More recently, slurry bubble-column reactors have been the entire Fischer–Tropsch technology was extensively investi-
developed to overcome some of these drawbacks. Syngas is gated at the US Bureau of Mines, which resulted in new two-
bubbled through a suspension of fine catalyst particles in the phase HTFT technology. The classical textbook by Storch,
liquid product phase. The catalyst particle sizes are usually less Golumbic, and Anderson originates from this period.3 Small
than about 100 mm, which is sufficiently small to prevent intra- plants were built in the US and operated in the 1950s.
particle diffusion limitations, while the well-mixed liquid phase Large-scale FTS developments mainly occurred in South
ensures virtual isothermal operation of the reactor. There are, Africa.33 Sasol started an FTS plant in 1955 based on HTFT to
however, certain technical challenges associated with FTS slurry make petrol and on LTFT to produce wax. Both HTFT and LTFT
reactors. A prerequisite of a slurry process is the development of used iron-based catalysts. The HTFT technology formed the
an efficient solid–liquid separation step to remove product wax basis for the large expansion of Sasol in the late 1970s/early
from the reactor. It is extremely important to ensure the 1980s when Sasol 2 and 3 were built in Secunda (see Table 1).
mechanical integrity of the catalyst to limit the extent of break- The main reasons for this expansion were the oil crises in the
up and attrition in the moving bed environment. 1970s, which led to a significant increase in the crude oil price
Of late there have been some new reactor developments for (see Figure 4).
FTS application, but none of these have been commercially These oil crises also initiated renewed interest in FTS from
applied yet. Microchannel reactors can support very high heat other companies like BP, ExxonMobil, Gulf, Shell, and Statoil,
and mass transfer rates and thereby address the problems of which was mainly based on cobalt FTS catalysts.34,35 In the last
traditional fixed-bed reactors, while the stationary bed circum- 20 years, this has led to new commercial GTL plants by PetroSA
vents the challenges of slurry reactors. This approach shows (South Africa; 1993), Shell (Malaysia; 1992), Sasol-Qatar

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
530 Fischer–Tropsch Synthesis: Catalysts and Chemistry

450 100

400 Patents 90

350 80

Number of articles/patents
Articles

US crude oil price ($)


70
300 US crude oil price
60
250
50
200
40
150 Articles Oil price 30
100 20
50 Patents 10

0 0
1970 1975 1980 1985 1990 1995 2000 2005 2010
Publication year
Figure 4 Patents and articles per year compared with the crude oil price (figure inspired by de Smit and Weckhuysen21).

FTS

ATR
ASU

Figure 5 Photo of the Sasol-QP Oryx GTL plant in Qatar, showing the air separation units (ASUs), the auto-thermal reformers (ATRs), and the
Fischer–Tropsch synthesis (FTS) slurry reactors. The product work-up section located behind the FTS reactors is not visible (photo courtesy of Sasol).

Petroleum (Qatar; 2007; see Figure 5), Shell (Qatar; 2011), oil-derived fuels. However, products from CTL facilities have
and Sasol Chevron (Nigeria; under construction – start-up a much larger CO2 impact, which is immediately clear from the
2013). An overview of the current commercial operations overall stoichiometric equations [1] and [2]. In the hypothet-
using FTS technology is shown in Table 1. ical limit of using a carbon feedstock which does not contain
The investment decision to build the Sasol-Qatar Petroleum any hydrogen, one CO2 molecule is formed for every carbon
Oryx-GTL plant was taken in 2003 when the oil price was $25/ atom that ends up in a hydrocarbon. A large portion of the
barrel. The facility was built at a cost of $1 billion. Currently, a CO2 produced in CTL plants is removed and concentrated, and
yearly profit is generated of about $500 million.36 Sasol’s is therefore ideally suited for capturing, that is, ‘capture ready’.
much larger Secunda CTL facility is generating currently At the same time new focus on products from biomass can
about $2 billion profit annually.36 Shell’s Pearl plant (both stimulate interest in FTS further, as biomass can be used as a
the FTS production – 140 000 barrels per day – and the carbon source for syngas generation. Recently, Oxford Cata-
upstream natural gas condensates – 120 000 barrels per day – lysts has demonstrated their FTS technology using syngas made
together) was built at a cost of $20 billion,37 and Shell from wood chips.28
announced to make annually $4 billion cash when Pearl is at Other opportunities for GTL applications in the future
full production with the oil price at $70/barrel. It is clear that might be the use of associated natural gas in small-scale plants
new GTL/CTL facilities require large capital investments, and (<1000 barrels per day, as Oxford Catalysts and CompactGTL
are heavily dependent on the prevailing crude oil price. How- are pursuing), as well as the use of shale gas in large-scale
ever, over the long term these large-scale GTL/CTL facilities do facilities (as pursued by Sasol).
make economic sense. From an academic point of view, renewed interest in FTS
A new challenge to the GTL/CTL technology is global warm- was clearly observed in two main waves (see Figure 4). The first
ing and the emission of CO2. Fuel products from GTL facilities wave occurred in the late 1970s, while the second one started
have a similar environmental footprint compared to crude around 1995 and is still gaining momentum. The latter

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 531

observation is also matched by an increase in patenting activ- 7.20.2.2 Commercial Applications


ity. Of course, this revived activity is related to the increase in
Sasol is a leader in the field when it comes to commercializing
the crude oil price, although the present academic interest is
iron-catalyzed FTS processes. In the early 1950s, Sasol com-
certainly also inspired by the notion that crude oil resources are
mercialized Fe-catalyzed FTS based on the Ruhrchemie process
limited. Typical topics in Fischer–Tropsch research with a high
to produce a variety of synthetic petroleum products using the
academic interest are catalyst preparation methods, deactiva-
Arge Tubular Fixed bed reactors (see Table 1 and Figure 3).
tion studies, and mechanistic and kinetic studies, in which
Later on, they developed the slurry-bed reactor and this reactor
sophisticated tools such as in situ catalyst characterization,
together with the Arge reactors are used to produce high-
surface science, molecular modeling, and transient kinetic
molecular-weight hydrocarbons for the wax industry.38 In the
studies are the common ingredients. The increased interest
late 1950s, Sasol also commercialized a circulating fluidized
from both the academic as well as the commercial world has
bed reactor at their Sasolburg facilities in which fused iron
created excellent scientific interactions and discussions, which
catalyst is fluidized at high temperatures to produce lighter
enabled further progress on this exciting topic of FTS. Many
hydrocarbons ideally suited for producing fuel and chemical
questions are still outstanding, such as the state of the catalyt-
feedstock. In the late 1970s/early 1980s Sasol’s Secunda plant
ically active surface under reaction conditions, and the reaction
was built using this circulating fluidized bed technology, which
mechanism in terms of elementary steps.
was replaced in the late 1990s by the improved fixed fluidized
bed technology. Subsequent to Sasol’s successful commercial-
ization of iron-catalyzed FTS, South Africa’s national oil
7.20.2 Iron-Based FTS Catalysts
company (PetroSA) commercialized a GTL facility using
7.20.2.1 Introduction previous-generation high-temperature FTS technology (Sasol
licensed technology). This technology is based on a fused
The iron-catalyzed FTS process is, along with ammonia synthe-
iron catalyst operated in a fluidized bed reactor at high tem-
sis, one of the most studied systems in the field of heteroge-
perature (330–350  C). In 2010, it was still recognized as one
neous catalysis. The reason for this is possibly the fact that the
of the world’s largest GTL refineries, producing about 22 000
application of the process is so versatile. Not only can iron FTS
barrels per day of high-quality FTS-derived fuels.
produce a light hydrocarbon product stream ideal for the fuel
Rentech, based in Colorado, USA has long been investing in
and chemical industry, it can also produce heavier hydrocar-
iron-based FTS research. Rentech demonstrated their iron-
bons (C35þ) suited for the waxes market. Iron is also a cheap
based FTS technology in their Product Demonstration Unit
raw material when compared to its cobalt counterpart (cobalt
(PDU) in the middle of 2008. The PDU produces approxi-
is on average 250 times more expensive than iron raw mate-
mately ten barrels per day of ultra-clean diesel, aviation fuels,
rials) and it has been commercially applied since the late 1950s
and naphtha.39
by Sasol38 (Table 1). Iron is believed to be more tolerant of
Synfuels China has recently emerged as an important player
poisons, for example, sulfur in synthesis gas than cobalt. It is
in the Fischer–Tropsch industry.32 Their development of a so-
also known to be responsive to selectivity manipulation by the
called high-temperature slurry-phase technology (HTSFTP™)
addition of promoters and a variation of typical process param-
and associated iron-based catalyst is novel to the industry. The
eters, for example, temperature, pressure, and H2/CO ratio. The
integrated technology promises improvements in process ther-
disadvantage, however, is the fact that iron FTS catalysts deac-
mal efficiency and a highly active catalyst. This technology has
tivate rather quickly (activity or selectivity loss) and this will be
been demonstrated in a 4000 barrels per day semi-commercial
discussed in more detail later in this section. As already men-
CTL facility owned by the YiTai Coal Liquefaction Company in
tioned the iron FTS process can be manipulated to produce a
Xue JiaWan, Erdos, Inner Mongolia.
range of carbon number distributions with the final product
stream depending mainly on the temperature applied during
FTS. At lower temperatures, for example, 220–250  C the chain
7.20.2.3 Iron Fischer–Tropsch Catalyst Preparation
growth probability (a) of the catalyst is approximately 0.94
indicating that the bulk of the products will consist of hydro- There are several preparation methods available in the litera-
carbons longer than C21. In the case of higher temperatures, for ture for the synthesis of Fe FTS catalysts, like precipitation and
example, 320–350  C, the chain growth probability decreases fusion. Iron catalysts prepared commercially are actually iron
to 0.7 and even lower with the main products being light oxides, hydroxides, or oxy-hydroxides, which undergo an acti-
hydrocarbons utilized for the production of transportation vation step such as reduction or pre-treatment in syngas prior
fuel and chemical feedstocks. Figure 1 shows the carbon to FTS. General requirements for the catalysts are, among
chain length as a function of chain growth probability. The others, selectivity (low for methane; high for the targeted
influence of promoters on selectivity will be discussed later in hydrocarbon fraction), activity and stability, and mechanical
this section. robustness. The operation conditions, for example, high or low
Although there are many advantages with regard to iron- temperature (HTFT and LTFT), and the type of reactor
catalyzed FTS, the transformations of the iron catalyst during employed put specific demands on the catalyst synthesis
activation and FTS are rather complex and still not fully under- procedure. For a tubular fixed-bed-type reactor, minimization
stood. During catalyst preparation, iron oxides (e.g., hematite of mass transfer limitations is an important consideration, and
(Fe2O3) and magnetite (Fe3O4)) are produced and these are here catalyst strength is less important than catalyst shape
transformed to either a-Fe or iron carbides during activation and form. For a fluidized bed reactor, however, catalyst
depending on the conditions. strength as well as particle size and density are very important.40

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
532 Fischer–Tropsch Synthesis: Catalysts and Chemistry

Table 2 summarizes the typical preparation methods used for 7.20.2.3.1 Fusion
the various applications. Fusion produces oxidic iron particles of low surface area, high
Key factors when choosing an iron source are cost and density, and high strength, which are ideally suited for appli-
availability. Although most of the iron oxides, hydroxides, cation in circulating fluidized bed reactors (Figure 3). During
and oxy-hydroxides are readily available in nature, precursors the fusion process, the iron oxide raw material together with
for iron catalysts are rather chemical grade raw materials.41 the promoters are fed into an arc furnace where it is subjected
This is done to ensure that impurities that can influence to temperatures above 1000  C. After fusion, the molten mate-
the catalyst are either removed or carefully controlled. It rial is cast into flat bars (ingots) and cooled. These ingots are
is typical for commercial manufacturers of iron catalysts to milled to a specified particle-size range to ensure optimum
produce chemical grade iron(III)nitrate from sufficiently fluidization (Figure 6).
pure scrap iron on site as part of the preparation. Large-scale The disadvantage of fusion is the fact that the inorganic
fusion preparation methods use iron ores or mill scale from impurities, for example, silica and alumina oxides present in
steel mills. Complex preparation methodologies that involve the raw mill scale starting material, form inclusions during
novel chemicals and/or intricate transformations are rarely cooling of the ingot. The alkali promoters migrate and bind
commercially viable when compared to the tried and trusted during cooling to these inclusions, which negates the promo-
methods of precipitation and fusion. The gains from such tion effect. Figure 7(a) is a scanning electron microscopy
novel preparations must be truly unique to justify the addi- (SEM) image of a fused ingot showing clearly the inclusions
tional expense. and with scanning electron microscopy energy-dispersive X-ray

Table 2 Catalyst preparation methods used for high and low temperature iron-based Fischer–Tropsch processes

Reactor Important catalyst properties Raw material Synthesis method

HTFT
Circulating or fixed fluidized bed reactors, Low surface area (<10 g m2), high density, Mill scale Fusion followed by crushing and
320–350  C high strength milling
LTFT
Tubular fixed bed reactor, 220–250  C High surface area, sufficient strength Fe(NO3)3 and silica Precipitation followed by
source extrusion/shaping
Slurry bed reactors, 220–250  C High surface area, small particles Fe(NO3)3 and silica Precipitation followed by spray
(50–250 mm) source drying and calcination

Promoters

Cooling Size
Mill scale Oxidation Arc furnace reduction Reduction
step
(milling)

Figure 6 Catalyst preparation diagram for the HTFT fused iron catalyst. Mill scale means iron metal pieces from the steel industry.

Inclusion

Fe
Si
Alkali 1
Alkali 2

(b)
80 mm
(a)

Figure 7 (a) SEM image of a HTFT Fe catalyst cast ingot, showing inclusions of silica and alkali and (b) SEM EDX mapping on one of these inclusions.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 533

spectroscopy (SEM EDX)(Figure 7(b)) one can see the pres- This high-temperature treatment removes volatile impurities
ence of the alkali in these inclusions. such as water and NOx and increases the strength of the catalyst
The development of fixed fluidized bed reactors diminished particles. Figure 10 shows spherical particles obtained from
the need for particles of high mechanical strength, and permit- spray drying. For fixed-bed reactor applications, the impreg-
ted the use of catalysts with lower density and higher surface nated filter cake is extruded and dried at about 150  C.
area. One of the prospective routes to alternative HT FTS The methodology described above is commercially applied
catalyst precursors is the precipitation of high-density, low- by Sasol for the synthesis of their slurry-bed reactor (SBR) and
surface-area iron oxides, hydroxides, and (oxy)hydroxides tubular fixed-bed reactor catalyst precursors. Rentech uses a
from solutions of iron(III) salts followed by calcination. similar method to synthesize their Fe FTS catalyst precursor.39
As mentioned above, Sasol developed ‘in-house’ precipitation
7.20.2.3.2 Precipitation methodology to synthesize a suitable Fe-HTFTS catalyst that
Precipitation of iron(III)oxides from iron(III)nitrate solutions offers many advantages over the fused Fe HTFT catalyst.43 Some
was one of the first methods reported in the literature for the of these include improved promoter distribution, increased
preparation of iron FTS catalysts.42 In the late 1930s, Ruhrch- strength, and spherical particles which improve fluidization
emie developed a large-scale preparation based on pre- of the catalyst. By replacing fusion with precipitation, the
cipitation. In this procedure, the iron(III) salt is reacted with negating effect of the alkali promoters could be eliminated
a base to form an iron(III) oxide–(oxy)hydroxide precipitate. owing to the purity of the starting iron(III)salt.
By variation in process conditions, for example, pH, precipita-
tion rate, and temperature, catalyst properties such as surface
Hematite: 12 – 27 m2 g-1
area and crystallite size can be controlled. Figure 8 illustrates
the decrease of crystallinity versus surface area for the various
iron oxides known as precursors for Fe FTS catalysts. Magnetite: 4 – 100 m2 g-1
Degree
After precipitation, the slurry is filtered and washed to
of
remove all the salts (e.g., NH4NO3) from the filter cake (see crystallinity
Figure 9). The latter is then reslurried and impregnated with Goethite: 8 – 200 m2 g-1
structural promoters like Si, Al, etc. The application of chemical
promoters in the iron catalyst is discussed in more detail later in
Ferrihydrite: 100 – 700 m2 g-1
this section. Next, the slurry is spray-dried to yield spherical
particles that are suited to slurry-bed and fixed fluidized-bed Figure 8 Degree of crystallinity is decreasing with an increased surface
reactors. The final step in the catalyst preparation is calcination. area for various iron oxides.

Filtration/
Dissolution Precipitation washing
HNO , Base
Water
(aq) (aq)

Iron

Re-slurry and
impregnation

Salt
(aq)

OR

Spray drying Extrusion

Calcination Drying

Slurry bed Fixed bed


catalyst catalyst

Figure 9 Catalyst preparation diagram for the slurry-bed and fixed-bed precipitated iron LTFT catalyst.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
534 Fischer–Tropsch Synthesis: Catalysts and Chemistry

The stability and composition of the final activated iron phase


determine the performance of the catalyst under Fischer–
Tropsch conditions. Because of this, the activation procedure
has a strong effect on selectivity and activity of the catalyst.
For activation in hydrogen, the extent of reduction of iron is
governed by the role of water removed during the activation.
The degree of reduction is governed by the equation47:
pH2 O
DG ¼ DG þ RT ln [4]
pH2
where ΔG is the free energy change for the reduction under the
conditions employed, ΔG is the standard free energy change
for the reduction reaction, R is the gas constant, T is the
temperature, and p is the partial pressure of the gases indicated.
The equation implies that the rate at which water is
removed from the reactor plays a critical role: the faster the
water is removed, the faster the reduction process proceeds,
Figure 10 Spray-dried iron LTFT slurry-bed catalyst particles.
and the higher the degree of reduction is. Rewriting exp-
ression [4] to include the equilibrium constant in the form of
7.20.2.3.3 Improving iron Fischer–Tropsch catalyst the ratio (pH2O/pH2)eq at equilibrium, enables one to estimate
precursors by promoters the degree of reduction that can be obtained:
FTS processes catalyzed by unmodified and unpromoted " 
  #
iron catalysts suffer from poor selectivity, low activity, and pH2 O pH2 O
DG ¼ nRT ln [5]
sintering, but the addition of structural and chemical pro- pH2 pH2 eq
moters addresses most of these issues.
As the equilibrium ratio for reduction from Fe2O3 to FeO is
Structural promoters, for example, Si, Al, and Mg, may sup-
0.7, and for FeO to a-Fe is 0.1, the theoretical degree of reduc-
press sintering, stabilize the active phase, and improve mechan-
tion would be 50% at 10% water in the gas phase. Hence, for
ical strength. The addition of alumina and silica typically
high degrees of reduction the water content should be well
increases the stability of hematite under FTS conditions.44 In
below 1%.47,48 Structural promoters such as silica and alumina
general, it is observed that in the presence of a structural pro-
increase the resistance against reduction.
moter such as silica the surface area of the iron oxide remains
To fully understand activation, it is necessary to understand
high even after calcination at relatively high temperatures.
the possible phase transformations that can occur. x-Ray dif-
A potential disadvantage, however, of adding structural
fraction (XRD) and Mössbauer spectroscopy are ideal tech-
promoters is that the activation, for example, reduction of the
niques to distinguish between the different iron phases that
iron oxide, becomes more difficult due to the formation of iron
can arise, while small particle effects, which so often limit the
silicates or aluminates. For this reason, chemical promoters
information content of XRD, are generally absent in the unsup-
such as Cu or Ag are added during catalyst synthesis to increase
ported iron FT catalysts.49–51 Activation with CO present in the
the rate of reduction, most likely due to hydrogen spillover
gas typically leads to a mixture of metallic iron (a-Fe), iron
from the Cu surface to the iron oxide surface. Apart from
carbides (general formula FexCy), and magnetite (Fe3O4). The
increasing the rate of reduction, chemical promoters are
relative quantities are a function of the reducing gas, the gas
known to (i) enhance nucleation of iron intermediates which
hourly space velocity, and the temperature.
leads to higher surface areas, (ii) increase the number/type of
It is generally believed that carbides such as Hägg-carbide
CO adsorption sites, (iii) stabilize selected phases, and (iv)
(w-Fe5C2) are the active phase for FTS.52 The exact nature of the
influence the rate of secondary reactions.45
surface carbidic species is still a subject of debate. The stabilities of
Addition of alkali metals (e.g., potassium) to LTFT iron oxide
different bulk iron carbides have been reported to be in decreas-
catalyst precursors is known to enhance the chain growth prob-
ing order: e0 -Fe2.2C> e-Fe2C > w-Fe5C2 > y-Fe3C.53 Depending on
ability (increased C5þ selectivity), to diminish methane forma-
the type of catalyst (e.g., fused or precipitated), different iron
tion, and inhibit secondary hydrogenation reactions, leading to
carbides were found to be characteristic for each type after acti-
higher olefin to paraffin ratios.46 In a similar way, but perhaps
vation. However, the possibility that a-Fe plays a role during FTS
not as effective, alkali earth metals have been shown to increase
cannot be ignored.21 Typically during activation, precipitated
alpha values and to suppress methane formation.
catalyst precursors, for example, hematite (Fe2O3), are converted
to magnetite (Fe3O4) irrespective of the activation gas used.
7.20.2.3.4 Activation and reduction procedures However, after this transformation the final iron phase will
Iron oxides and (oxy)hydroxides are inactive for FTS and must depend on the activation gas used, for example, a-Fe in the case
be activated to render an active catalyst. Depending on the of hydrogen or Hägg-carbide (w-Fe5C2) in the case of CO or
application – HTFT or LTFT – and iron oxide precursor, activa- synthesis gas. In a study by Herranz et al., it was found that the
tion is performed in hydrogen, carbon monoxide, or synthesis activation of hematite using CO resulted in mainly cementite
gas. The optimum activation conditions are influenced by the (y-Fe3C) while activation in synthesis gas yielded Hägg carbide
type and quantity of chemical and structural promoters. (w-Fe5C2) (Table 3).54

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 535

The fused magnetite HTFT catalyst is reduced with hydrogen from lighter to heavier hydrocarbons. Although the Anderson–
at temperatures between 350 and 450  C and high linear flow Schultz–Flory (ASF) distribution curve (see Section 7.20.1 and
rates to avoid re-oxidation by water, as explained above. During Figure 2) gives a good indication of expected selectivities, alkali
reduction, oxygen atoms are removed from the lattice leading to promotion of iron catalysts leads to selectivities that tend to
an increase in surface area from <1 g m2 to  5–8 g m2 55. deviate from ASF and are characterized by two alpha values
The extent of reduction under these conditions was measured at (see Figure 11).
about 80% (a-Fe). Under CO or synthesis gas, virtually no In attempts to increase the selectivity toward valuable base
reduction of the nonporous magnetite was observed. chemicals, mixed-metal oxides and/or multicomponent metals
In the case of precipitated iron oxide catalyst precursors for are typically incorporated in precipitated iron catalysts. Addi-
LTFT catalysts, the activation is usually done under much milder tion of manganese to a typical Ruhrchemie catalyst increases
conditions than for fused catalysts. These catalysts are more the selectivity toward alpha-olefins at low-temperature FTS
amorphous with a high pore volume and surface area and the conditions (230  C and 20 bar total pressure).57 Another
oxide crystallites can sinter under too harsh activation condi- example is the Fe/Zn/Mn/Cu/K/SiO2 catalyst for direct conver-
tions. It is important to note that the success of activation of sion of synthesis gas to chemicals.58 The development of such a
precipitated iron catalyst precursors is coupled to FTS activity, technology is known as ChemFT, and has as primary focus to
stability, and selectivity. This is dependent not only on the type shift the selectivity toward alcohols. Table 4 compares the
of reduction gas but also on process conditions, for example, selectivities of the various iron FTS technologies (HTFT, LTFT,
temperature and pressure. From the literature, it seems as and ChemFT) and illustrates that the alcohol selectivity in
though activation under CO yields the optimally activated pre- ChemFT is much higher than that of a typical Ruhrchemie
cipitated iron catalyst for FTS synthesis, as these gave the best catalyst under similar LTFT conditions.58
syngas conversion and lowest methane selectivity when com-
pared to catalysts activated with H2 or synthesis gas.56 However,
the final catalyst also had a relatively high WGS activity. 7.20.2.5 Catalyst Stability During FTS
Stability is a key characteristic of a successful catalyst. The ideal
FTS catalyst should maintain constant activity and a correspond-
7.20.2.4 Selectivity Manipulation of Iron Catalysts
ing stable selectivity during time on stream. Commercial reac-
A key advantage of iron-catalyzed FTS is the fact that the tors and product workup sections are designed for a very narrow
selectivity of the process can be manipulated, either by process set of optimum process conditions. The catalyst must perform
conditions (less responsive) or by catalyst composition (more within these design constraints for as long as possible. This
responsive). Lowering the temperature shifts the selectivity determines the useful catalyst life. Unfortunately, iron catalysts
show considerable loss of performance over time. During recent
Table 3 Names of the various iron phases years, the focus of research has shifted from improving catalyst
activity to increasing the lifetime of the catalyst.
a-Fe2O3 Hematite Deactivation of iron FTS catalysts is usually attributed to the
a-FeOOH Goethite following factors:
Fe3O4 Magnetite
FeO Wustite (i) ‘free’ carbon formation, leading to catalyst fouling,
w-Fe5C2 Hägg carbide (ii) activity loss due to transformation of the phase, for exam-
y-Fe3C Cementite ple, oxidation,

LT FT
HT FT
Ln (X )

Alpha 1
(LT FT) = 0.80 Alpha 2 (LT FT) = 0.94

Alpha (HT FT) = 0.75

0 5 10 15 20 25 30 35 40 45 50
Carbon number
Figure 11 Anderson–Schulz–Flory distribution of hydrocarbons formed over an LTFT catalyst (typical Ruhrchemie catalyst) and a fused HTFT catalyst.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
536 Fischer–Tropsch Synthesis: Catalysts and Chemistry

(iii) mechanical break-up of the catalyst, so-called competition model.59 After the adsorption and dis-
(iv) deposition of poisons in the synthesis gas on the catalyst’s sociation of CO and H2, three reactions are possible:
surface, and
(i) C* þ iron ! carbides
(v) sintering.
(ii) C* þ xH* ! CHx*
Buildup of ‘free’ carbon is one of the major causes of (iii) C* þ yC* ! inactive carbon
deactivation in HTFT. It leads to a decrease in density and
The first reaction describes the formation of iron carbides
strength of catalyst particles and results in catalyst bed expan-
from the reduced a-Fe under FTS conditions. The dissociated
sion, particle break-up, and carry-over of fine catalyst material
carbon (C*) can either react with dissociated hydrogen atoms
into downstream processes.25 Figure 12 shows an SEM image,
(H*) to yield hydrocarbons or react with another carbon atom
along with the distribution of elements of a ‘spent’ catalyst
(C*) to yield inactive/ so-called ‘free’ carbon.59 This type of
retrieved from a commercial fixed fluidized bed reactor. Cata-
deactivation can be suppressed by chemical promoters. In
lyst break-up and fines formation are easily recognized. As
recent years, Sasol developed another propriety catalyst involv-
carbon is dispersed through the bulk of the particle, break-up
ing the addition of chromium to reduce the amount of ‘free’
may lead to exposure of new active surfaces, thus helping to
carbon formed during FTS.60
maintain activity. At the same time, the carbon that is lost in
The formation of ‘free’ carbon is less pronounced in the case
the form of fines is rich in alkali and thus removes some of the
of the precipitated LTFT iron catalysts. The main deactivation
chemical promoter, which degrades the selectivity.
mechanisms in this case are sintering and oxidation of the
Much has been discussed regarding the origin of the free
active phase. Interconversion of different carbides may lead
carbon in the catalyst. A plausible explanation is given by the
to a stoichiometric excess of carbon which in turn leads to
weakening of catalyst particles. Figure 13 shows a deactivation
curve for a typical Rührchemie catalyst under low-temperature
Table 4 Selectivity comparison between LTFT, HTFT, and FTS conditions. Samples of this catalyst taken from the reactor
ChemFT25,58 usually contain mixtures of highly dispersed magnetite and
iron carbide (both containing around 2 nm particles).61 The
Product Fe HTFT Fe LTFT Fe ChemFT highly dispersed magnetite particles can either react in synthe-
sis gas to the required iron carbide, or they can agglomerate or
CH4 (%) 8.0 3.0 18.0
sinter into larger inactive particles (about 40 nm). The larger
C2–C4 olefins (%) 24.0 4.0 21.0
C2–C4 paraffin (%) 6.0 4.5 17.7 magnetite particles can agglomerate further to yield large glob-
C5–C6 (%) 16.0 7.0 13.3 ules (around 400 nm). Surprisingly, agglomeration or sinter-
C7–350  C (middle distillate 36.0 26.5 20.5 ing of highly dispersed iron carbide into less active or inactive
product) iron carbide particles of about 20 nm has also been observed
350  C (wax products) 5.0 51.0 0.0 (Figure 14).
Oxygenates as alcohols (%) 2.8 3.8 8.3
Oxygenates as acids þ ketones 2.2 0.2 0.9
% breakdown (C5–C12 cut) 7.20.2.6 Spent Catalyst Management
% total paraffins 13.0 29.0 49.6
% total olefins 70.0 64.0 37.8 Regeneration of ‘spent’ iron FTS catalysts is difficult, due to the
% aromatics 5.0 0.0 0.0 sintering of the particles during FTS. Successful regeneration
% oxygenates 12.0 7.0 12.5 requires redispersion of the sintered phase, and this cannot
easily be achieved. Reactivation by re-reduction is possible, but
the activity of the reactivated catalyst is lower because the

2.0
Relative activity ratio

1.5

1.0

0.5

0.0
-Fe -C -Si 0 50 100 150 200 250
Time on line (h)
Figure 12 Scanning electron microscopy (SEM) image of a spent,
fused Fe HTFT catalyst; color coding: red, iron; yellow, carbon; and green, Figure 13 Relative activity versus time on stream for a precipitated
silicon. Ruhrchemie-type iron LTFT catalyst.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 537

H2 + CO
CnHm
FTS
H2 + CO CO CO
(Fe3O4)HD (FeO)HD (Fe)HD (FexCy)HD
-H2O -CO2 -CO2

Sintering H2O Oxidation


Sintering

Sintering
(Fe3O4)LP (Fe3O4)G (Fe2C)LP

50 nm
50 nm

200 nm

HD – highly dispersed phase » 2 nm


LP – larger particles » 20 nm
G – globules » 400 nm
Figure 14 Deactivation mechanism for a typical Ruhrchemie iron catalyst under low-temperature Fischer–Tropsch synthesis conditions.

original surface area cannot be recovered. Multiple reactivation Cobalt FTS catalysts are currently commercially applied by
steps are therefore not viable. Spent HTFT catalysts may in Sasol/QP in the Oryx GTL plant, Qatar (Co/Al2O3), in a slurry-
principle be recycled to make new catalysts. However, using phase reactor, and by Shell in the SMDS plant in Bintulu,
the material in the fusion process has a high energy cost associ- Malaysia, as well as in the Pearl plant, Qatar (both Co/Mn/
ated with it, due in part to its high carbon content. Iron is a TiO2), in a fixed-bed reactor. Figure 15 shows the catalyst that
cheap material, and there is little economic incentive for recov- is used in slurry-phase application.
ering it. Therefore, spent catalysts have usually been landfilled. Exciting academic and industrial research in the last 20
Currently, awareness of the environmental impact of such pro- years has increased the fundamental knowledge of cobalt FTS
cedures is growing, and reclamation of metal – even iron – from catalysts substantially on topics like the nature of the active
spent catalysts, for example, by acid dissolution is more and site, impact of crystallite size on activity and selectivity, and
more seen as a social responsibility of the industry to reduce the deactivation mechanisms, owing to the application of surface
impact of commercial processes on the environment. science techniques, model catalysts, in situ analyses at relevant
industrial conditions, and molecular modeling.8,19,27,62–67 The
literature of the last 20 years shows that quite a wide variety of
cobalt catalyst compositions prepared by numerous methods
7.20.3 Cobalt-Based FTS Catalysts can result in academically and industrially relevant cobalt-
based FTS catalytic systems.
7.20.3.1 Introduction
Cobalt as an FTS catalyst was already claimed by Fischer and
Tropsch in their original patent of 1925.31 The commercializa-
7.20.3.2 Composition of Cobalt Catalysts
tion of the FTS by Germany and Japan in the period 1938–45
relied fully on cobalt catalysts. Only after World War II did the Modern cobalt catalysts are similar to the ones prepared by
focus shift to the use of iron catalysts for FTS applications. Fischer and Tropsch in the sense that they consist of promoted
Since the oil crises of the 1970s the interest in cobalt-based cobalt on a metal oxide support. An inspection of the literature
FTS catalysts reappeared, which has resulted in numerous sci- and patents on this topic reveals the following general charac-
entific papers and patents (see Figure 4). Many companies teristics, with almost all companies with FTS catalysts having a
showed interest in cobalt FTS, for example, BP, ConocoPhilips, similar formulation for them18,35,68 (Table 5):
Gulf, ExxonMobil, IFP, Johnson Matthey, Sasol, Shell, Statoil,
(a) Cobalt as the FTS active metal (typically 10–30 wt%)
and Syntroleum. Almost all focused on wax production, fol-
(b) A second metal (usually noble) as a reduction promoter
lowed by hydrotreating to produce diesel. This is also the
(0.05–1 wt%)
application that will receive most attention in this section.
(c) A structural oxidic promoter (e.g., Zr, Si, and La) (1–10 wt%)
Cobalt FTS catalysts are exclusively utilized in low-
(d) A refractory oxidic support (most likely modified)
temperature synthesis or LTFT, and are applied in fixed-bed,
slurry-phase, and micro-channel FTS reactors. Catalyst design Cobalt is expensive and to maximize its use, it needs be well
needs to be adjusted to the targeted reactor as well as the applied dispersed on the support. Since cobalt metal is considered the
FTS conditions. Important for catalyst design are the composi- active phase, it is imperative that there is a high density of
tion, method of preparation, activity and selectivity behavior, cobalt metal sites available. The number of cobalt surface
deactivation and regeneration, and mechanical integrity. sites is a function of particle size and morphology, extent of

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
538 Fischer–Tropsch Synthesis: Catalysts and Chemistry

Cobalt catalyst

~ 60 m
Wax

0.1 m

-4
5 ⫻ 10 m
Support particles
10 m

Cobalt

Slurry phase reactor Support

-8
1 ⫻ 10 m 1 ⫻ 10-7 m
A cobalt nanoparticle Cobalt and support

Structural promoter

Reduction promoter

Cobalt

Cobalt
Cobalt

Modified support

Figure 15 Cobalt catalysts for application in a slurry-phase reactor, and schematical composition of a typical cobalt-based Fischer–Tropsch catalyst.

Table 5 Examples of catalyst formulations, as patented by several industrial Fischer–Tropsch synthesis companies

Company Composition Reference Preparation route

Co (wt%) 2nd metal Structural promoter Support

Shell 20 – MnO (Co/Mn ¼ 12) TiO2 WO 199700231 Coprecipitation


ExxonMobil 12 Re (1 wt%) Al2O3 (6 wt%) TiO2 US 5268344 Impregnation
Syntroleum 20 Ru (0.1 wt%) La (1 wt%), SiO2 (0.1–10.6 Si/nm2) Al2O3 WO 2005058493 Impregnation
BP 10 – Al2O3 (0.5 wt% Al) ZnO WO 19913400 Impregnation
Sasol 20 Pt (0.05 wt%) SiO2 (0.8 Si/nm2) Al2O3 US 7365040B2 Impregnation

reduction, and particle stability.68 It is preferred to have a fairly reported that for cobalt particles less than 40 nm, the predom-
high extent of reduction (>60%), but it should also be noted inant phase should be fcc.72 The mode of activation, addition
that the cobalt is further reduced during the FTS reaction. An of promoters, and support may influence the relative amounts
optimum cobalt particle size of just above 8–10 nm is pre- of the phases.19 Some authors have reported that the hcp phase
ferred as particles below those have shown to have a lower is more active for FTS.73
turnover frequency (TOF).69 Additionally, very small particles Nanometer-sized cobalt particles when supported on tradi-
(4–6 nm) could be more prone to sintering and also may prove tional oxidic carriers like silica, alumina, and titania are diffi-
very difficult to reduce due to an increased metal-support cult to reduce due to strong interactions with the support.
interaction. It is important that there is a minimum amount Therefore, catalysts are often promoted with a second metal
of cobalt-support compounds as these are reducible at very (e.g., Ru, Pt, or Re) which leads to improved reducibility of the
high temperatures and are inactive for the FTS reaction.70 The cobalt oxide particles; the increase in amount of active sites
two most common phases of metallic cobalt in supported results in higher activity compared to un-promoted catalysts.
cobalt FTS catalysts are face-centered cubic (fcc) and hexago- The more facile cobalt reduction is attributed to faster hydro-
nally close-packed (hcp), which often coexist.19,71 It has been gen activation in the presence of promoter metals and

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 539

subsequent spillover of hydrogen to cobalt oxides and reduc- There has also been work conducted on less conventional
tion of cobalt species.19,68 In many cases, the promotion with supports such as MCM-41, SBA-16, and carbon nanofibers,
noble metals leads to a smaller average size of either cobalt nanotubes, and spheres.19,69,87 These studies are mainly aca-
oxide or cobalt metal particles. Promotion with noble metal demic in nature but further fundamental understanding of
may also play a role during the decomposition of the cobalt cobalt FTS catalysts considerably. Carbon supports interact
precursors and can lead to crystallization of smaller cobalt weakly with cobalt and allow for a high degree of cobalt reduc-
oxide particles. This increased dispersion is most likely due to tion, thus enabling the study of cobalt particle-size effects.69,87
a higher rate of nucleation, enabled by the promoter.68
Promoter metals such as Ru have also been claimed to lead
to the formation of bimetallic particles and alloys. This influ- 7.20.3.3 Preparation of Cobalt Fischer–Tropsch Catalysts
ences catalyst activity and selectivity, may inhibit deactivation The preparation of cobalt FTS catalysts aims to achieve the
by keeping the surface clean, and allows easier regeneration of optimal crystallite size distribution in a particle that is optimal
the cobalt surface.74 The metal promoter is usually present at for its application in a specific fixed-bed, bubble-column, or
levels of 0.1–0.5 wt%. At these low concentrations reduction is microchannel reactor. As the optimum size range of catalyst
efficiently promoted, and the hydrocarbon selectivity is hardly particles for the different reactor types varies (see Figure 3),
negatively affected. preparation methods and equipment depend on the targeted
Structural promoters affect the formation and stability of the reactor application. Important considerations for choosing a
active phase of a catalyst material. For Co/silica catalysts, it has particular method of preparation and starting components are
been shown that promotion with Zr results in a decreased cobalt– to minimize poisons (e.g., Na, S, Cl) in the catalyst and the
silica interaction, which in turn leads to a higher degree of cobalt type of waste streams resulting from the chosen method.
reduction and increase in the metallic atoms on the surface.19,75 A number of procedures for preparing cobalt FT-catalyst
Zr promotion of cobalt/alumina catalysts has been claimed to precursor exist:
prevent formation of cobalt aluminate.76 Incorporation of ele-
ments such as B77 and Ni78 increases the stability of cobalt • coprecipitation of cobalt, promoters, and support, followed
catalysts by suppressing carbon formation. Irreducible oxides by catalyst particle shaping. In a variation of this method,
such as MnO and CeO2 may also slow down cobalt sintering.63 the support is added just before particle shaping,
A wide range of promoters has been studied; the reader is referred • precipitation or impregnation of cobalt and promoters
to a detailed review by Morales and Weckhuysen.63 onto pre-shaped support particles, and
The support provides mechanical strength and thermal sta- • impregnation of cobalt (oxide or metal) particles onto pre-
bility to the cobalt crystallites, while facilitating high cobalt shaped supports.
dispersion. The properties of the support are an important
factor. For alumina, high purity, low acidity, and relatively 7.20.3.3.1 Precipitation
high surface area (150–250 m2 g1) are required, according to Most of the initial FTS-catalysts (e.g., Co/ThO2/kieselguhr)
patents from the 1980s.79–81 More recently, however, alumina- were made by coprecipitation.88 This method has been applied
based supports of relatively low surface area (50 m2 g1), such for some of the modern cobalt catalysts as well, for example,
as Ni-promoted a-Al2O3, have been reported to have a positive for Co/Mn catalysts,89 Co/Mg/SiO2 and Co/ZnO2.90
effect on both mechanical strength and C5þ selectivity.82 The Catalyst preparation based on coprecipitation usually con-
pore size of the support can also influence the size of the cobalt sists of three steps: precipitation, washing and drying, and
crystallites, as shown by Saib et al. for SiO2-supported shaping. Selection of chemicals is of course an important con-
catalysts.83 Van Steen and Claeys reported that the desired sideration in view of the associated waste streams.
pore size of the support for the optimum cobalt crystallite Chemical precipitation of the cobalt, promoter, and sup-
size should be around 12–16 nm.61 port by a precipitation agent can be done batchwise or contin-
The support needs to be robust under FTS conditions, imply- uously at constant pH. The cobalt precipitates as a hydroxide,
ing that it should be able to cope with the presence of several which can exist as green a-Co(OH)2 or pink b-Co(OH)2 poly-
bars of steam that occur at high conversion levels. Van Berge morphs. The former is metastable and readily transforms into
et al. found that an unprotected alumina-supported cobalt FTS the stable b-phase. Crystallite size and composition of the
catalyst was susceptible to hydrothermal attack during realistic precipitate are controlled by temperature, precipitation agent,
FTS conditions, which resulted in contamination of product wax precursor salts, structure directing or organic hydrolysis
with ultra-fine, cobalt-rich particulates.23,84,85 This problem was reagents, aging time, and reaction atmosphere (air or N2).
solved by pre-coating the support with silica as structural pro- Using Na2CO3 or KOH as precipitation agents in the prepara-
moter. TiO2 seems to be the support of choice for both Exxon tion of Co/SiO2 catalysts would lead to cobalt silicate
and Shell based on the most recent patents (Table 5). An advan- formation.91 To prevent formation of the inactive cobalt sili-
tage of TiO2 is that it has a high hydrothermal stability and can cate, the silica is added after the precipitation.
withstand high water partial pressures. The rutile/anatase ratio Filtration and washing of the precipitate is required to
can be tailored, which influences the surface area and mechan- remove excess chemicals. Even low levels of alkali metals and
ical properties. halogens left in the washed precipitate can severely degrade the
Supported cobalt catalysts should also be resistant to attrition catalyst’s performance.
especially if applied in a slurry bubble-column environment. Shaping of the catalyst precursor depends on the reactor
Wei et al.86 noted that the attrition resistance of supported cobalt application. For bubble beds, the precipitate is usually reslurried
catalysts follows the sequence: Co/Al2O3 > Co/SiO2 > Co/TiO2. and spray-dried to obtain the required particle-size distribution.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
540 Fischer–Tropsch Synthesis: Catalysts and Chemistry

For fixed-bed reactors, the precipitate is extruded or pelletized. on particle geometry (diffusion path length and pore size),
Addition of acids to the washed and dried precipitate is done to viscosity of the suspension, interaction between solution and
improve the final catalyst’s particle strength.92 support surface (contact angle and surface tensions), and
diffusion coefficients.93–95 Figure 16 shows the effect of dif-
fusion on the different cobalt distributions observed on small
alumina particles for incipient wetness impregnation com-
7.20.3.3.2 Preparation methods involving pre-shaped pared to slurry-phase impregnation, as obtained from SEM/
supports EDX line scans.96,97
Support morphology and characteristics play an important role
in optimizing the preparation of cobalt catalysts on pre-shaped
• For fixed-bed catalysts, eggshell-type cobalt distributions
are sometimes preferred to overcome pore diffusion limita-
support particles. As the aim is to get a desired amount of
tions on performance and selectivity. Concentrating the
cobalt crystallites onto the support and maintain a crystallite
cobalt in the outer layers of the support is, among others,
size of around 8–10 nm, the following support characteristics
achieved by adding viscosity enhancers or using cobalt salt
need consideration:
melts for impregnation.98
• The support pore volume dictates how much cobalt precursor • For deposition precipitation onto pre-shaped supports, the
can be added per impregnation. As shown in Table 6, 30 g of same parameters that determine the time required during
metallic cobalt per 100 g of support occupies only 0.03 ml g1 impregnation to get a homogenous distribution (e.g., par-
of support material, but when using Co(NO3)26H2O as the ticle geometry, diffusion path length, pore size, viscosity of
precursor, a pore volume of 0.79 ml g1 is required. the suspension, interaction between solution and support
• The time required to get a homogeneous cobalt distribution surface, contact angle and surface tensions, and diffusion
throughout a support particle during impregnation depends coefficients) are important.99

Table 6 Pore volume requirements for different cobalt components

Cobalt compound Molar mass (g mol1) Cobalt mass fraction (%) Density (g cm3) Pore volume required for a loading of 30 g
of Co per 100 g of support (ml)

Co 59 100 8.9 3.4


CoO 75 0.79 6.4 5.9
Co3O4 241 0.73 6.1 6.7
Co2O3 166 0.71 5.2 8.2
CoOOH 92 0.64 5.0 9.4
Co(OH)2 93 0.64 3.6 13.2
Co(NO3)2 183 0.32 2.5 37.7
Co(NO3)2 6H2O 291 0.20 1.9 79.0
CoCl2 6H2O 237 0.25 1.9 62.5

60
60

50 50

40 40
Mass% Co3O4
Mass% Co3O4

30 30

20 20

10 10

0 0
-10 0 10 20 30 40 50 60 70 80 -10 0 10 20 30 40 50 60 70
Distance from edge of catalyst particle (mm) Distance from edge of the catalyst particle (mm)
Figure 16 Macroscopic cobalt crystallite distribution, as measured by scanning electron microscopy (SEM) line scans for: (a) incipient wetness
impregnation followed by immediate fast drying and (b) slurry-phase impregnation allowing 3 h for the cobalt to slowly disperse throughout the alumina
particle, followed by fast drying.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 541

• In addition, diffusion differences between the precipitation For optimal reduction, care must be taken to optimize heat
agent and cobalt will impact the final cobalt distribution transfer, minimize hydrogen diffusion and mass transfer limi-
during precipitation. When depositing bulky cobalt tations, and to remove water effectively. The latter benefits
hydroxide crystallites or cobalt metal particles100 onto pre- from high hydrogen space velocities and application of low
shaped support particles, the important parameters are pore heating rates.104 Figure 18 shows the impact of the water
size, particle diameter, and bulkiness of the precipitate. partial pressure during reduction of a 30 g Co/0.075 g
• During the drying of the catalyst precursor, the same Pt/100 g alumina catalyst on the starting FTS performance,
parameters as highlighted above for impregnation and indicating that low water content should be targeted for max-
deposition should be taken into account to prevent the imum activity, in agreement with the thermodynamics of
cobalt precursor (if not chemically fixed to the support) reduction as expressed in eqn [5]. Hence, the hydrogen stream
from migrating out of the particle again. In addition, heat used for reduction must be as dry as possible. For small catalyst
transfer coefficients, evaporation enthalpies, and particle particles as used in slurry-bed reactors, fluidized bed reduction
outer surface area need consideration for optimizing the reactors are preferred and the above requirements are easily
drying phase of the cobalt catalyst preparation. met. For reductions in fixed-bed reactors, more care must be
taken to overcome the limitations especially toward the reduc-
Each preparation method needs to be optimized carefully, tion reactor outlet.
and one cannot assume that the optimum procedures for one The maximum temperature required for reduction of cobalt
type of support and support particle shape will be the same for catalysts depends on the level of reduction promoter present
all other supports and support particle shapes. (Pt, Ru, Pd, etc.), the presence of other promoters (e.g., alkali
metals make reduction more difficult), the support, the sup-
port modifiers, and the catalyst precursor used in the prepara-
7.20.3.3.3 Calcination tion. Reduction temperatures that are too high can cause
Usually, drying is not fully accomplished and therefore the first sintering and loss of cobalt metal surface area. In the case of
stages of calcination actually complete the drying phase. To Co/SiO2 catalysts, cobalt silicate formation has been reported
maintain the cobalt distribution achieved by impregnation or for temperatures higher than 350  C.
precipitation during drying and calcination, the cobalt compo- Catalyst performance depends critically on the reduction
nent mobility must be hindered. One way of achieving this is procedure.105 Application of reduction–oxidation–reduction
to ensure that the cobalt component stays in a viscous or solid (ROR) cycles has been reported to improve the FTS perfor-
form. For catalysts obtained by impregnation from cobalt mance of cobalt catalysts by up to 30%.106 Some of the reasons
nitrate solutions, this implies that during calcination the com- given in the literature for this improved performance from
bination of heating rate and air flow must be such that water ROR treatment are: (1) rougher (more steps on the surface)
and NOx are immediately removed.101 As the mobility of the cobalt crystallites, (2) higher degree of reduction, and (3) re-
cobalt phase can be minimized by fast calcination, heat flow dispersion of the cobalt on the support surface.
into the system is also important as both the drying and nitrate
decomposition are endothermic.
Performing calcination under different atmospheres pro- 7.20.3.4 Cobalt Catalyst Fischer–Tropsch Performance
vides a way to affect the dispersion of the cobalt phase. NO Both activity and selectivity are of course important parameters
addition during calcination leads to the formation of a less for cobalt-based FTS catalysts. High activity is important for
mobile cobalt hydroxyl nitrate.102 Using H2 or CO as decom- slurry-phase catalysts, while for fixed-bed catalyst the heat
position medium at temperatures below those where reduction removal capacity needs to be balanced with the activity of the
takes place also gives catalysts with good cobalt dispersions.103 catalyst. From a selectivity point of view, a low methane selec-
Adding organic additives during impregnation is another tivity is normally desired, combined with a high C5þ selectivity
method to influence cobalt nitrate decomposition. Oxidation or a high chain growth probability (a). Determining the intrin-
of the additive is an exothermic process, which provides heat sic catalytic performance of cobalt FTS catalysts is not a
for the endothermic nitrate decomposition, and thus acceler- straightforward exercise, as it is influenced by the choice of
ates its decomposition. reactor and conditions. Khodakov et al.19 summarize a num-
The transmission electron microscopy (TEM) images in ber of issues and choices related to the testing of FTS catalysts,
Figure 17 illustrate how cobalt distributions change when for example: (i) reactor choice: fixed-bed, slurry-bed (or con-
different calcination conditions are applied. tinuous stirred-tank reactor, CSTR), or high-throughput reac-
tors, (ii) hydrodynamics, (iii) heat transfer and hot spots, (iv)
intra particle and external mass-transfer limitations, and (v)
7.20.3.3.4 Reduction atmospheric or elevated pressure. As testing catalysts under
Cobalt catalysts are usually reduced in hydrogen or a diluted different FTS conditions (H2/CO ratio, T, and P) will result in
hydrogen atmosphere. Examples of CO reductions are also different catalyst performances35,107 and therefore possibly
found, but carbon formation on the cobalt crystallites should selection of different catalysts, it is important in the early stages
be avoided. Reduction of cobalt oxide to cobalt metal occurs in of research to understand the long-term scaling-up view, with
two exothermic steps: respect to reactor choice and FTS conditions. Taking the same
Co3 O4 þ H2 ! 3CoO þ H2 O [6] cobalt catalyst and testing it in different manners can result in
very different catalytic activity behavior. Figure 19(a) clearly
CoO þ H2 ! Co þ H2 O [7] shows that the FTS activity is influenced by the water partial

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
542 Fischer–Tropsch Synthesis: Catalysts and Chemistry

0.5 μm 0.5 μm

(a) (b)

0.5 μm 0.5 μm

(c) (d)

Figure 17 Cobalt crystallite distributions as measured with STEM for cobalt alumina catalysts calcined in different manners. (a) Cobalt oxide
microglobical formation of a 30 g Co/100 g alumina catalysts using a heating rate of 1  C min1 and an air space velocity of 1 m3n per kg Co(NO3)2.6H2O
per hour. (b) Cobalt oxide distribution of a 30 g Co/100 g alumina catalysts using optimized heating rate and air space velocity to ensure optimum
calcination. (c) Cobalt oxide distribution on a 30 g Co/100 g alumina catalysts using carbon coated alumina , using the same heating rate and air flow rate
as in (a). (d) Cobalt oxide distribution on 30 g Co/100 g alumina catalysts using the same heating rate and flow rate as (a) but with 1% NO in He as
calcination atmosphere.

pressure applied during the test, which possibly impacts factors was confirmed by Iglesia108 who showed that the activity of
such as sintering and carbon deposition, as well as surface and cobalt catalysts is directly proportional to the amount of cobalt
active site reconstruction.65 The selectivity behavior of cobalt metal surface in the catalyst. The TOF or the reaction rate per
catalysts is also strongly influenced by parameters such as unit of cobalt surface area was stable over the range of cobalt
temperature, hydrogen and carbon monoxide partial pressures, particles that was investigated (9–200 nm). FTS over cobalt
and conversion. Comparing catalysts tested at different condi- catalysts was therefore regarded to be structure insensitive.
tions should thus be done with care. Thereafter, a number of authors have investigated the effect
Figure 19(a) clearly shows that cobalt catalysts are more of cobalt metal particle size on the intrinsic activity of sup-
active in fixed-bed than in slurry-bed reactors. However, cobalt ported cobalt catalysts for smaller cobalt particles, that is, well
catalysts in slurry-phase reactors are normally applied at tem- below 10 nm.69,87,109–112 Bian et al.110 using Co/SiO2 con-
peratures around 230  C, while in fixed-bed reactors they are firmed Iglesia’s results for samples with cobalt particles
normally used at temperatures around 210  C. The productiv- between 11 and 29 nm, but Barbier et al.,109 Bezemer et al.,69
ity per gram of catalyst is therefore higher in slurry-phase Martinez and Prieto,111 Coville and coworkers87 all showed
reactors (Figure 19(b)). that the TOF was stable for catalysts with cobalt particles above
For heterogeneous catalysts the activity often increases with 8–10 nm, while it decreased sharply for catalyst with smaller
smaller particle size, as the metal surface area increases. This particles. Only Borg et al.113 reported no sensitivity for the

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 543

activity of cobalt particles with sizes down to 3 nm. The As mentioned above, the C5þ selectivity depends strongly
relationship between TOF (or reaction rate per unit cobalt on the FTS process conditions. For a constant set of conditions,
surface area) and cobalt particle size for above-mentioned Iglesia114 found no particle-size effect on selectivity for parti-
publications is summarized in Figure 20. As the TOF numbers cles between 9 and 100 nm. A few years later, however, Barbier
were obtained under different FTS conditions (i.e., tempera- et al.109 reported a strong dependency of the chain growth
ture and partial pressures), they were normalized to enable probability, a, on particle size. Increasing the cobalt diameter
comparison of trends in the different papers and therefore from 4.5 to 9.5 nm caused the a-value to increase from 0.74 to
expressed in arbitrary units. It is clear that for catalysts with 0.87 (at 170  C, 1 bar). Bian et al.110 showed a similar, though
cobalt crystals above 10 nm the TOF is structure insensitive, less pronounced, trend with the a-value increasing from 0.85
while there is a sharp decrease in activity for particles smaller to 0.89 when the particle size increased from 11 to 29 nm (at
than 8–10 nm. 200  C, 10 bar). Bezemer et al.69 reported a very clear particle-
size effect on selectivity at atmospheric pressure, with a meth-
ane selectivity that was stable for particles larger than 6 nm, but
increased sharply for smaller particles (220  C, 1 bar).
However, the reported data at high pressure (35 bar and
210  C) clearly show that the C5þ selectivity still increases with
1.0 increasing particle size up to 15 nm. Xiong et al.,87 Prieto
et al.,112 and Borg et al.113 all confirmed the general trend on
Relative FT activity

an increasing C5þ selectivity with increasing particle size


0.9 extending beyond 10 nm, and up to 20 nm.
Little fundamental understanding has been offered to
explain this particle-size effect on both activity and selectivity.
0.8 Interestingly, the effect of size on activity is very pronounced
for particles smaller than 10 nm, while the impact on selectiv-
ity seems to be more gradual and does not level off above
0.7
10 nm. The particle-size effects cannot, as previously suggested,
10 30 50 70 90 110 130 150 170
be explained by the oxidation of the smallest particles. Bezemer
PH 0 (mbar)
2
et al.69 showed with x-ray absorption near-edge structure
Figure 18 Impact of water partial pressure during reduction of a 30 g (XANES) measurements that oxidation did not occur. This is
Co/0.075 g Pt/100 g alumina catalyst on its initial Fischer–Tropsch in line with extensively reported research that cobalt oxidation
synthesis performance. during FTS does not occur for cobalt particles larger than

1.6
Productivity (mole CO converted/g cat/s)

1.4

1.2
Activity (au)

0.8

0.6

0.4

0.2

0
0 200 400 600 800 1000 0 200 400 600 800 1000
Time (h) Time (h)

Figure 19 (a) Three Fischer–Tropsch synthesis runs with same Co/Al2O3 catalyst, as tested at 20 bar, 230  C, and H2/CO ¼ 2. Red solid: fixed bed,
PH2 O inlet ¼ 0; PH2 O outlet ¼ 3.0 bar. Red open: fixed bed with water co-feeding; PH2 O inlet ¼ 2.6, PH2 O outlet ¼ 4.0 bar. Blue solid: slurry bed;
PH2 O inlet ¼ 0; PH2 O outlet ¼ 4.5 bar; all the catalyst is exposed to outlet water partial pressure in a CSTR laboratory slurry reactor. (b) Two Fischer–
Tropsch synthesis runs with same Co/Al2O3 catalyst, as tested at 20 bar, 60% conversion, with the fixed bed run at 210  C (red circles) and the
slurry phase run at 230  C (blue triangles).

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
544 Fischer–Tropsch Synthesis: Catalysts and Chemistry

2 nm.65 The particle-size effect can also not be explained by of Codþ–SiO2 sites in small particles, caused by flattening of
sintering as this was not observed by Bezemer et al.69 and by the crystals during FTS, and possibly due to carbon-induced
the Coville group.87 It is clear that the Co particle-size effect in surface reconstruction. Another study showed that carbon and
FTS extends beyond the classical impact of size, which derives oxygen atoms, originating from dissociated CO, were very
from the fraction and type of surface atoms as a function of strongly bonded on small particles, possibly blocking active
crystallite size and normally does not extend beyond 4 nm sites for further CO dissociation.117
particles. It was suggested69 that the optimum combination As the impact of the particle size is different for activity from
of active sites for the different elemental reactions of FTS (i.e., for selectivity, the fundamental explanation of these effects
CO dissociation, hydrogenation, and insertion) requires rela- seems to involve more than one kinetically relevant mechanis-
tively large cobalt particles, possibly combined with a CO- tic step in the FTS process. Understanding this requires more
induced surface reconstruction. This might be related to the fundamental research.
presence or absence of the so-called B5 site,65 which has been
speculated to be the most active site for CO dissociation, and
7.20.3.5 Deactivation and Regeneration of Cobalt
needs a certain particle size to be present in high abundance.
Fischer–Tropsch Catalysts
Another speculation is that the particle-size effect might be
related to specific bonding modes of CO, such as the bridge- Catalyst stability is crucial for the economics of cobalt FTS,
bonded CO coordination; this mode is believed to be favored in addition to other important factors such as high activity,
on large particles and held responsible for an increased CO selectivity, and mechanical strength. Understanding catalyst
dissociation rate, which would lead to an increased reaction deactivation is essential for improving catalyst stability and for
rate.87,109 Based on in situ Fourier transform infrared developing effective regeneration procedures. Figure 21 shows a
spectrometry results, Prieto et al.112 proposed an enhancement typical deactivation profile for Co FTS catalysts under
TOF (au)

Coville (2011)
Fischer (2010)
Pietro (2009)
Borg (2008)
Martinez (2007)
De Jong 1 bar (2006)
De Jong 35 bar (2006)
Bian (2003)
Barbier (2001)
Iglesia (1997)

0 5 10 15 20 25 30
Co particle size (nm)

Figure 20 The TOF or Fischer–Tropsch synthesis rate per unit surface area, as a function of the cobalt metal particle size.69,109,110,112–116 The TOF had
to be scaled due to variations in process conditions, and is therefore reported in arbitrary units.

1.0

0.8
RIAF

0.6

0.4

0.2

0.0
0 10 20 30 40 50 60
Time on line (days)

Figure 21 Normalized activity stability for a Co/Pt/Al2O3 catalyst during realistic Fischer–Tropsch synthesis in a laboratory scale micro-slurry
reactor at fixed CO conversion (230  C, 20 bar, H2 þ CO conversion of 50–70%, feed gas composition of 50–60 vol.% H2 and 30–40 vol.% CO). Adapted
from van de Loosdrecht, J.; Bazhinimaev, B.; Dalmon, J. A.; Niemantsverdriet, J. W.; Tsybulya, S. V.; Saib, A. M.; van Berge, P. J.; Visagie, J. L.
Catal. Today 2007, 123, 293–302.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 545

commercially relevant FTS conditions.118 Deactivation is initially 1.8


stronger, after which it starts leveling off. Similar deactivation

Hydrogen resistant carbon (wt%)


1.6
profiles have been reported by other laboratories.18,62,119,120
Fundamental studies on catalyst deactivation essentially 1.4
involve understanding differences in the fresh and spent cata- 1.2
lysts with respect to the active site. The B5 sites on metallic
1.0
cobalt are currently considered as active sites for the FTS.65
Hence, changes to the number or nature of these sites will 0.8
contribute to deactivation. However, due to the complexity of
0.6
the FTS process and the lack of suitable techniques to charac-
terize any particular site on a surface, most studies on deacti- 0.4
vation are necessarily limited to ‘observables’ such as changes 0.2
in metallic cobalt surface area. This in itself is no trivial matter,
0.0
as cobalt catalysts are sensitive to their environment, and spent 0 20 40 60 80 100
FT catalysts are embedded in wax (which actually protects Time on stream (days)
them from being exposed to the air). As is so often the case
in fundamental studies of catalysts, a combined approach Figure 22 Build up of polymeric carbon on a cobalt slurry-phase
Fischer–Tropsch synthesis catalyst as a function of time on stream.121
using real catalyst and model systems, advanced in situ and
ex situ characterization, combined with molecular modeling
has given detailed insight into FTS catalyst deactivation.65 FT activity. Although detailed studies are unknown to us, pore
The main deactivation mechanisms of cobalt FTS catalysts blocking is generally accepted as a deactivation mechanism.
and its active sites, as proposed in the literature, are: (1) oxi- Recently, strong evidence has been found for the accumula-
dation, (2) mixed metal-support interaction, (3) carbon depo- tion of a rather inactive polymeric carbon on the metallic surface
sition and carburization, (4) sintering, (5) poisoning, and (6) of cobalt resulting in catalyst deactivation (Figure 22).121 Tech-
surface reconstruction.19,62,64,65,70,71,108,114,118,121–133 niques involved in this work were temperature-programmed
Over the last 15 years, oxidation of cobalt by the product hydrogenation, low-energy ion scattering, and energy-filtered
water was seen as the major deactivation mechanism in the TEM. Molecular modeling suggested that the polymeric carbon
open literature.65 However, many recent publications have dis- might be a form of graphene. In general, there is good agree-
proved this.62,64,65,71,118,134Following an in-depth study on oxi- ment that bulk cobalt carbide is metastable and will not be
dation using model systems, molecular modeling, surface present in substantial quantities during FTS for cobalt, although
thermodynamic calculations, and an industrial catalyst tested Karaca et al.71 observed small quantities of Co2C in their in situ
under commercially relevant conditions, the key finding is that XRD experiments (which, interestingly, was entirely absent for
oxidation is crystallite size and condition dependent, that is, the first 9 h on stream, but appeared in the measurements after
under realistic FTS conditions (H2O/H2 ¼ 0.5–3). Co crystallites 10 h; exposure to pure CO made the signal grow further).
with diameters larger than about 2 nm will not undergo oxida- Although characterization of spent cobalt FT catalysts run for
tion. In fact, from XANES analyses of spent Co catalysts from an several months in a slurry bubble column did not show bulk
extended FTS run, it was found that a further reduction took cobalt carbide formation,118,121 a possible role of subsurface
place during FTS118 and has been confirmed by others.135–137 carbon cannot be ruled out. To date, our knowledge on subsur-
Further, the formation of metal-support compounds such face carbon comes from molecular modeling. Calculations on
as cobalt aluminate have been considered as a deactivation carbon clusters by Zonneville et al.140 indicate that carbon in
mechanism.129,138 Although thermodynamically favorable, subsurface positions affects the CO dissociation rate and may
this reaction needs CoO formation as an intermediate, which therefore affect the FTS activity as well. More work is needed to
does not take place under realistic FTS conditions. Indeed, a ascertain the impact of subsurface carbon.
recent study showed that the minor cobalt aluminate formed Carbon deposition can be decreased by adding promoters.
during FTS originates from unreduced CoO present in the fresh Examples from the literature include ruthenium74 and
catalyst and not from Co metal.65,122 This leaves sintering and boron.77,141,142
carbon deposition as the major contributors to Co FTS catalyst Sintering is a thermodynamically driven process whereby
deactivation. smaller, more unstable particles grow to form larger, more
Carbon deposition on an FTS catalyst that is covered by stable particles that are lower in surface energy. Sintering as a
growing hydrocarbons and is entirely embedded in product deactivation mechanism is easier to investigate than carbon
wax represents a real challenge. Nevertheless, deleterious car- deposition, with TEM and XRD being the most common tech-
bon arising from CO or FT products can have a wide range of niques used. Due to this in general there is good agreement in
negative effects on Co FTS catalysts. We mention pore block- the literature on the importance of sintering as a deactivation
age, resulting in mass transfer limitations, formation of bulk or mechanism of cobalt FTS catalysts.62,65,70,71,121,134,137,143–145
surface carbides, and blockage or alteration of active sites.121 Key factors that affect the rate of sintering are the reaction
Pore blockage by long hydrocarbon products resulting in dif- temperature and the partial pressure of water: an increase in
fusion limitations of the reactants CO and H2 has been men- either of these results in enhanced sintering. The choice of
tioned as a deactivation mechanism from the onset of the support also plays a key role. Alumina is considered to provide
discovery of the FTS.139 A hydrogen treatment or solvent more stability against sintering than silica does, due to the
wash of the spent catalyst resulted in a partial recuperation of improved metal support interaction in the former. Both

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
546 Fischer–Tropsch Synthesis: Catalysts and Chemistry

particle migration and Ostwald ripening are expected to be mechanism of redispersion of cobalt has been proposed to be a
important, but further studies are desirable to gain full under- two-step process, that is, (1) oxidation to form hollow spheres
standing on the contribution of each. by the Kirkendall effect and (2) multinucleation of Co3O4
Poisoning of cobalt catalysts by S, NH3, HCN, Hg, and Cl is during reduction to produce smaller crystallites.65,153
well known and is an issue for cobalt-based FTS62,146–149 espe- Poisons such as sulfur are removed by oxidation with steam
cially for coal-to-liquid applications.65 Sulfur is a strong, irre- and air to sulfates, followed by washing them out. However,
versible poison with a large adsorption energy. Due to its size phases originating from strong metal-support interaction are
and electronic effect, sulfur will also poison adjacent cobalt very difficult to reverse and their formation should be
sites. Once adsorbed, sulfur is difficult to remove and will prevented.154
accumulate with time.146 Sulfur poisoning can however rela- Substantial progress has been made toward understanding
tively easily be prevented by cleaning the synthesis gas feed deactivation and regeneration of Co FTS catalysts, but more spe-
properly, for example, by using zinc-oxide or lead-oxide guard cific knowledge, for example, on deactivation by carbonaceous
beds. Poisoning of cobalt-based FTS catalysts by means of species and sintering mechanisms, is definitely necessary. With
nitrogen-containing compounds such as NH3 and HCN is a currently available in situ characterization, synchrotron-based
known effect and postulated to arise from competitive techniques, and molecular modeling, we expect major advance-
adsorption.150 The impact of N-compounds is less severe ments in the coming years.
than that of S-compounds and can be reversed by mild hydro-
gen treatment. Nevertheless, reducing their level to parts per
billion is recommended.150 7.20.4 Mechanisms and Kinetics of FTS Over Iron
Surface reconstruction is a thermodynamically driven pro- and Cobalt Catalysts
cess which results in a lowering of the surface energy and
7.20.4.1 Introduction
therefore can contribute to catalyst deactivation. Using molec-
ular modeling, Ciobica et al.124 showed that atomic carbon The complicated nature of the FTS is among others reflected by
from dissociated CO can cause a reconstruction of the Co fcc its complex product spectrum, consisting of methane, C2þ
(111) surface to a Co (100)-like structure, followed by a clock olefins and paraffins (linear and branched), oxygenates
reconstruction. This surface is less active and could therefore (mainly alcohols, but also aldehydes and ketones), and even
contribute to deactivation. As the reconstruction is accompa- aromatics (at sufficiently high operating temperatures). Three
nied by a change in surface density, it could, ironically, also classes of mechanisms have been proposed, each assuming a
assist in the formation of more reactive sites, as proposed by different monomer for chain propagation.
Wilson and de Groot.151 This is a complex phenomenon that The carbide mechanism (Table 7) was first formulated by
needs further investigation. Fischer and Tropsch in 1926, which proposes that CO dissoci-
Methods to reverse deactivation and regenerate deactivated ates before the carbon atom is partially hydrogenated to a CHx
Co FTS catalysts have been around since the early days of species.2 These CHx species combine by the addition of one
Fischer and Trospch.3 The most common methods reported in monomer at a time to effect hydrocarbon chain growth. The
the open literature are treatment of the deactivated catalyst growing intermediate can then terminate in different ways
in hydrogen or in steam, applying oxidation–reduction cycles, before leaving the catalyst surface, giving rise to an alkane, an
and combinations of these.65,152 With carbon and sinte- alkene, or an oxygenate. A number of variations have been
ring being the major deactivation mechanisms during FTS, considered within the basic carbide mechanism, arising from,
‘oxidation–reduction’ is considered to be the most robust and for instance, whether CO dissociation occurs unassisted or via
preferred method to regenerate spent Co FT catalysts.65 By interaction with hydrogen, and the number of hydrogen atoms
careful control of the oxidation step, deleterious carbon is in the monomer (CH or CH2). The enol mechanism (Table 7),
removed at temperatures above 250  C. The oxidation step proposed by Storch in the 1950s, assumes that CO is partially
is also key for the redispersion of cobalt (see Figure 23). The hydrogenated to a CHOH species (oxymethylene) which then
20 nm

20 nm 20 nm

Figure 23 Co particles supported on a flat SiO2 support during different stages of an oxidation–reduction cycle. Cobalt particles in the metallic
state before oxidation (left). Hollow oxide particles, formed upon oxidation (middle). Reduction of the hollow particles, which break up into several
smaller metallic particles (right).65,153

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 547

Table 7 Fischer–Tropsch Reaction Mechanisms

Carbide mechanism Enol mechanism CO insertion mechanism

Initiation CO ! C þ O CO þ 2H ! CHOH CO ! C þ O
O þ 2H ! H2O O þ 2H ! H2OH
C þ xH ! CHx
Propagation R þ CHx ! R–CHx R–C–OH þ CHOH ! R–C–COH þ H2O R–CH þ CO ! R–CH–CO
R–CHx þ (2  x)H ! R–CH2 R–C–COH þ H ! R–CH2–COH R–CH–CO þ H ! R–CH2–CH þ H2O
Termination R–CH2 þ H ! R–CH3 R–CH2–COH þ 4H ! R–CH2–CH3 þ H2O R–CH þ 2H ! R–CH3
R–CH2–CH2–H ! R–CH ¼ CH2
R þ CO ! oxygenates R–CH2–COH þ nH ! oxygenates R–CH–CO þ nH ! oxygenates

NB : R ¼ H, alkyl(CH3, CH3CH2, CH3CH2CH2, . . .).

acts as the monomer.3 Chain growth occurs via a condensation temperature of 180 K.156 Such a low-temperature experiment
reaction with water elimination. Again, different termination in ultrahigh vacuum (UHV) conditions is equivalent to incre-
routes determine what final product molecule is formed. In the asing the CO pressure in a room temperature experiment.157
1970s, Pichler and Schultz introduced the CO insertion mech- Initially CO adsorbs, with a sticking coefficient of 0.7, on top
anism (Table 7), which assumes a similar initiation step to the sites up to a coverage of 0.33 ML (monolayer), with an adsorp-
carbide mechanism.155 The difference resides in the way that tion energy of  115 kJ mol1.156 This translates into a desorp-
chain growth is proposed to occur, namely by direct CO inser- tion temperature around 400 K. Upon further dosing, the CO
tion into the growing intermediate followed by hydrogenation coverage increases, until a saturation coverage of  0.65 ML is
to remove the oxygen atom. reached. Increasing the CO coverage beyond 0.33 ML leads to
Even though the FTS has been known since the 1920s, it is complex overlayers where CO occupies bridge and threefold
evident from the foregoing that its mechanism is still a matter sites as well as top sites. The downward shift of the CO desorp-
of debate. In order to progress the understanding of the mech- tion temperature for coverages beyond 0.33 ML is mainly
anism, a multidisciplinary approach is required. Subsequently, caused by repulsive interactions between CO molecules rather
three such disciplines (model surface science experiments, than by the difference in adsorption site. This has important
density functional theory (DFT) calculations, and macrokinetic implications for the interpretation of vibrational spectra on
studies) will be briefly discussed with particular emphasis on supported catalyst particles, where occupation of both top
their implications for mechanistic and kinetic understanding. and bridge/threefold sites is typically detected. Occupation of
bridge and threefold sites can simply be caused by a high CO
coverage on the facets of the particle rather than by the pres-
7.20.4.2 Surface Science Studies and Model Reactions ence of special sites.
Hydrogen/deuterium adsorption on a close-packed Co
A surface science approach is very powerful to study elemen-
surface was studied in a similar fashion: hydrogen adsorbs at
tary reaction steps in isolation. Conceptually, it is very close to
180 K with a low sticking coefficient, up to a coverage of 0.5
the approach taken by DFT calculations: take a well-defined
ML hydrogen atoms, with an adsorption energy of 33 kJ mol1
surface, that is, a single-crystal surface of the material you want
(per H atom). Recombinative desorption occurs between 300
to study and use an ultrahigh vacuum so that the adsorbate of
and 400 K.158 Hydrogen desorbs at lower temperature from
interest can be introduced with high purity and with a high
more open surfaces, around 300 K, indicating a weaker adsorp-
accuracy, down to a sub-monolayer coverage. The model sys-
tion onto those surfaces. The sticking coefficient on the other
tem can then be studied with many different sophisticated
hand is much higher than on close-packed surfaces: 0.76
analysis techniques that give information on the atomic level.
for an open surface compared to 0.05 on a close-packed
The number of studies with cobalt and iron single-crystal
surface.159,160 This enhanced hydrogen sticking is commonly
surfaces related to the FTS is relatively small. Iron carbide
observed on the more open crystal planes of different metal
single-crystal work is not available, as far as we know. Studies
surfaces.
on nickel and rhodium crystals are more numerous, as they are
When CO is dosed at 180 K on hydrogen-covered surfaces,
easier to use and because of the fact that both metals are used
CO partially replaces the hydrogen, and only 50% of the initial
as a catalyst for a number of reactions. In this section, we
coverage remains on the surface. The remaining hydrogen is
discuss the most relevant surface science findings on cobalt,
less strongly bound due to the CO, and as a result the desorp-
with a focus on studies where a single elementary step was
tion peak shifts downward by  100 K. CO, on the other hand,
studied in isolation.
is only mildly influenced by the presence of hydrogen, and any
influence of hydrogen is only seen at low temperatures.158
7.20.4.2.1 Adsorption of CO and hydrogen on model These experiments show that repulsive interactions exist
surfaces between hydrogen and CO, which adds to the barrier to form
One of the simplest experiments one can do is to study the hydrogenated HxCO species. Other adsorbates such as sulfur,
interaction of a surface with CO and H2, the reactants in the FT oxygen, and carbon give rise to a similar downward shift of the
reaction. In a typical experiment, a clean close-packed Co hydrogen desorption temperature, and those species also
surface is exposed to increasing amounts of CO at a sample (partly) block the surface for hydrogen adsorption.158

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
548 Fischer–Tropsch Synthesis: Catalysts and Chemistry

7.20.4.2.2 C–O bond scission rapidly and are not observed in significant concentrations on
A key step in any FT mechanism is the cleavage of the C–O the surface. This is in line with the finding that formaldehyde
bond. The close-packed surface of cobalt, Co (0001), is unable (H2CO) decomposes with 100% selectivity to CO and H2
to cleave the CO bond: CO desorbs as a molecule with the CO between 100 and 200 K.160 The experiments show that partially
156
bond hydrogenated species such as HCO and H2CO are very unstable,
 intact.
161,162
Some open surfaces of cobalt, such as Co
1012 and Co 1120 ,163 and cobalt foils are capable and the barrier to decompose via dehydrogenation is obviously
of cleaving the CO molecule directly: after a CO thermal much smaller than that of (Hx)C–O bond cleavage. Similar
desorption experiment161,162 or a prolonged exposure to CO experiments have been reported on 2-nm cobalt particles sup-
at elevated temperature,163 carbon and oxygen are found to be ported on alumina. In those experiments, 60% of the methanol
left on the surface. CO dissociation stops when enough carbon that was present dehydrogenated, while 40% underwent C–O
and oxygen has built up to block all the active sites for disso- bond cleavage.164 In this experiment, it was not clear whether
ciation. UHV studies on 2-nm Co particles on alumina show the C–O bond scission was assisted by the presence of hydrogen,
that CO dissociates during a CO thermal desorption experi- as CO (produced by methanol dehydrogenation in the metha-
ment, which can be explained by the high defect density on nol experiment) also dissociates on those particles in the
such small nanoparticles.164 These studies did not report exact absence of hydrogen.
temperatures or activation barriers for CO dissociation, but in In the CO insertion mechanism (Table 7), a CO molecule is
all cases a typical reaction temperature in the order of 400 K inserted into a CxHy, after which the C–O bond has to be
can be deduced. In short, surface science shows that direct CO cleaved to generate a Cxþ1Hz intermediate that can insert
dissociation is possible, but not on the (most abundant) close- another CO molecule. In other words, the CO molecule is
packed surface. chemically modified by insertion of an alkyl group on the
As direct CO dissociation is not possible on the close-packed C-end of the molecule. Experiments using ethanol on a close-
surfaces of cobalt, one might consider if the CO molecule is packed Co surface gave information about the effect of alkyl
(partly) hydrogenated before the C–O bond breaks. When CO modification on the C–O bond cleavage.166 Figure 24 shows
and hydrogen are co-adsorbed onto a close-packed surface at the result of such an experiment: ethanol adsorbs as an ethoxy
low temperature, the molecules just desorb upon heating, with- species at 160 K. This ethoxy species decomposes around
out reaction.158,160 An alternative experimental approach is to 350 K, via an acetaldehyde intermediate, with an activation
study the decomposition of (partly) hydrogenated CO mole- barrier of  70 kJ mol1. The products of this dissociation
cules such as methanol and formaldehyde. Methanol adsorbs as step are atomic O and a C2Hx species, demonstrating C–O
methoxy (H3CO) when dosed at 165 K. During heating this bond cleavage. This means that alkyl insertion in the CO
methoxy species is stable up to  300 K, after which it decom- molecule facilitates C–O bond scission. For the CO insertion
poses to CO and H2.165,166 Experimentally, this is seen as a mechanism, it implies that after the CO is inserted the C–O
single step, indicating that the first dehydrogenation is rate- bond can be readily broken and a Cxþ1Hy species is formed that
limiting. The intermediate species, H2CO and HCO, decompose can undergo further chain growth.

O1s O1s, TP-XPS C1s, TP-XPS C1s (high res.)


250 K

hv = 650 eV hv = 650 eV hv = 380 eV hv = 321 eV


O–CH2–CH3 O–CH2–CH3
Photo-emission intensity (a.u.)

Photo-emission intensity (a.u.)


300 K

250 K
250 K
350 K

C2H2

Oad

370 K 370 K

534 532 530 528 532 530 528 286 285 284 283 287 286 285 284 283 282
Binding energy (eV) Binding energy Binding energy (eV)
Figure 24 Ethanol decomposition on a close-packed cobalt surface, followed with temperature programmed (TP) synchrotron x-ray photoelectron
spectroscopy (XPS). At low temperature, the ethoxy intermediate is found. During heating ethoxy decomposes around 350 K, yielding atomic oxygen
and C2Hx (acetylene) on the surface, showing clear evidence for C–O bond cleavage. Adapted from Weststrate, C. J.; Gericke, H. J.; Verhoeven, M.;
Ciobica, I. M.; Saib, A. M.; Niemantsverdriet, J. W. J. Phys. Chem. Lett. 2010, 1, 1767–1770.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 549

7.20.4.2.3 Hydrogenation and the stability of C1Hx species to estimate pre-exponential factors, although these are generally
C1Hx species are important ingredients in a carbide mechanism less often calculated than enthalpies.
as they are the monomeric species responsible for chain initi- Validation of calculations has to be sought by comparing
ation and growth. Surface science experiments can give infor- calculated adsorption energies and activation barriers or vibra-
mation about the stability of the different C1Hx species. tional frequencies from stable adsorption states with experi-
Regarding cobalt there is only one article that addresses this mental values, obtained from surface science experiments with
question directly, using CHxCly intermediates on a cobalt single crystals. However, relatively little experimental data are
foil.167 The authors mention that the behavior on nickel foil available, partly also because surface science studies are by
was essentially the same. In an experiment on a close-packed necessity often performed in vacuum, whereas FTS reaction
nickel surface, a molecular beam was used to dissociate meth- steps occur at higher pressures.
ane at low surface temperature, which generates CH3 species As this is an area of research that is emerging rapidly, we
(þH) on the surface.168 Heating of this CH3 layer showed intend to present some examples of how DFT modeling is used
transformation of CH3 to CH around 200 K, and no sign of in mechanistic studies. Although DFT results refer to tempera-
CH2 was found. On close-packed Pt reported in the same tures of zero K and pressure, the results give valuable insight
study, a very similar trend was seen: CH3 decomposes around into the energetic of the underlying surface chemistry. It is not
250 K, yielding solely CH, which decomposes around our intention to give a full review here, as it is still too early for
500 K.168 These surface science results demonstrate that the conclusive statements on the FTS reaction mechanism.
CH2 species, which is typically seen as the monomer for
chain growth, is particularly unstable, which means that its
7.20.4.3.1 CO Dissociation
concentration under equilibrium concentrations will be
Conversion of CO and H2 into CxHy þ H2O necessarily implies
much lower than that of CH3 and CH.
that the C–O bond has to broken. The question is now if this
Generally speaking, surface science studies on cobalt and
happens before or after reaction with H-atoms. As an example, we
iron are scarce in comparison to those on nickel and the more
show a computational study of direct CO dissociation on the
noble metals. Studies on iron foils and single crystals,169–176
square (100) surface of bcc-iron in Figure 25.179,180 On this
although very interesting from the point of view of surface
surface, the CO molecule is known to adsorb in a tilted
chemistry, are even further removed from the reality of
geometry.171 In the transition state for dissociation, the C–O
Fischer–Tropsch reactions than cobalt, as iron FTS catalysts
bond elongates, and the energy rises by 1.14 eV ( 109 kJ mol1),
are essentially carbides. Surface science studies on iron carbides
which is the activation energy for dissociation. This value com-
are, to the best of our knowledge, not available.
pares well with the experimentally measured activation energy of
110 kJ mol1 reported by Bernasek and coworkers.171 When the
bond breaks, the C and O atoms each end up in a fourfold hollow
7.20.4.3 DFT Modeling
site of the Fe(100) surface. However, the two atoms significantly
Molecular modeling by DFT offers a relatively new way to repel each other, implying that the total energy decreases substan-
understand reaction mechanisms at the molecular level.177,178 tially when the two atoms move apart, as shown in the last
Adsorption configurations along with their energetics as well as structure of Figure 25. This series of calculations demonstrates
transition states can be modeled, and thus adsorption energies, that the direct dissociation of a CO molecule on this (100) surface
activation energies, and heats of reactions can be obtained. In is very well possible under reaction conditions (a barrier of
addition, entropy changes over elementary reactions enable one 110 kJ mol1 corresponds roughly to a reaction temperature of

Transition state

Adsorbed CO

Dissociated CO
1.14 eV
(exp 110 kJ mol-1)
Dissociated CO
repulsion relieved
2.30 eV
-0.34 eV

0.82 eV

-1.16 eV

Extent of reaction
Figure 25 Energy diagram for the dissociation of carbon monoxide on the (100) surface of iron, showing the exothermicity of the dissociation, and the
effect of repulsion between carbon and oxygen atoms that are adsorbed onto adjacent sites. Energies are given in electron volt (eV);
1 eV  96.5 kJ mol1). Adapted from Bromfield, T. C.; Ferre, D. C.; Niemantsverdriet, J. W. ChemPhysChem 2005, 6, 254–260.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
550 Fischer–Tropsch Synthesis: Catalysts and Chemistry

400–450 K, whereas the FTS reaction temperature is at least dissociation is favored, and often with a lower barrier than for
475 K), provided the space is available for the dissociation the reaction H þ CO.186 Nevertheless, it is good to realize that
products to move apart. Hence, this elementary step needs a so- on certain surfaces the H-assisted pathway may be an alternative
called ‘ensemble’ of iron atoms, which is determined by the that gives less reactive facets the chance to play a role in FTS.
condition that the C and the O atom can end up on next nearest
neighbor sites, that is, at distances at least equal to √2 times the 7.20.4.3.2 C þ H reactions
lattice constant. Many studies have addressed the formation of CHx fragments
Table 8 compares the activation energy for dissociation on a all the way to methane on several surfaces. Particularly inter-
number of iron surfaces with increasing reactivity.181,182 The esting is the comparison made by the Nørskov group.187 Their
trend is that dissociation becomes easier when the surface be- calculations are based on a full set of DFT calculations on the
comes more reactive. Note that the (110) surface is unlikely to play fcc (211) step of ruthenium surfaces, from which they esti-
a role in CO activation, as the barrier is forbiddingly high.181–183 mated the adsorbate energies on several other transition and
Several authors have proposed that CO dissociation becomes group IB surfaces. According to the Sabatier Principle,188 the
easier when the CO first reacts with hydrogen.183,185,186 The optimum pathway for a reaction is that in which the interme-
reasoning is evident, as the C–O bond in an HC–O or C–OH diates adsorb at the catalyst surface in a moderate way, that is,
fragment is expected to be weaker than in the CO molecule. not too strongly and not too weakly. The set of profiles in
However, forming the HCO or COH fragment also costs Figure 26 illustrates that metals such as Ni, Rh, and Co are
energy.186 It appears that on surfaces of low reactivity, such as close to ideal methanation catalysts, but that (the stepped
Co(0001), Fe(110), or the (100) surface of the Hägg carbide, H- surfaces of) metals such as Ru, Fe, and W bind the intermediate
assisted dissociation indeed leads to a lower activation barrier species too strongly. On the other extreme are Au and Ag,
than direct CO dissociation does.183,185 On more reactive sur- where formation of intermediates is strongly endothermic
faces, and notably on surfaces containing steps, the direct and therefore unfeasible. Metals such as Cu, Pt, and Pd are

Table 8 Activation energies for direct CO dissociation

Iron Characteristics CO dissociation activation Reference


surface energy

(110) Flat, close packed, least reactive surface 149 kJ mol1 Sorescu181
(100) Flat, somewhat more open, and thus more 103–110 kJ mol1 Bromfield et al.,179 Scheijen et al.,184 and Sorescu181
reactive than (110)
(310) Stepped surface with narrow terraces 70–87 kJ mol1 Sorescu181
(710) Stepped surface with broader terraces 64–86 kJ mol1 Sorescu181

5 Ag

Au

Cu
Energy (eV)

0 CO + 3H2
Pd, Pt

CH4 + H2O
Ni, Rh
Co
-5
Ru
Fe

W
)

O*

O*

C*

2*

3*

4 (g)

4 (g)

4 (g)
O(g

CH

H
+C

*+
*+

+C

+C

CH

CH
+C

*+

+C
6H
+C
2 (g)

5H

H*

H*

g) +
2 (g)

H*

H*
2 (g)

+4

+3
3H

+
O*

2 O(
+2
3H

*+
O*
3H

O*

O*

H
OH
O*

Figure 26 Density functional calculation for the energy profile of the methanation reaction on the group VIII and I-B metals, showing that nickel,
rhodium, and cobalt fall close to the optimum profile for the reaction, while iron and tungsten form too strong bonds with the carbon, oxygen, and the
CHx intermediates. Adapted from Jones, G.; Bligaard, T.; Abild-Pedersen, F.; Norskov, J. K. J. Phys. Condens. Matter 2008, 20.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 551

incapable of breaking the CO bond, but can carry out CO formation.200 It is obvious that mechanistic understanding
hydrogenation to methanol. would greatly benefit from more surface science work, but
Not shown in the figure are the barriers for the individual unfortunately the possibilities for this branch of physical
steps. As shown by several authors, activation energies for chemistry research are limited for reactions that by necessity
hydrogenation of adsorbed CHx (x ¼ 0–3) species on uncorru- require high pressure conditions.
gated (atomically flat) surfaces are in the range of 50–
100 kJ mol1 and therefore not expected to be problematic
7.20.4.4 Macrokinetic Observations and Models
for the reaction mechanism.189–194
In order to derive a macrokinetic model, a scheme of elemen-
7.20.4.3.3 Chain growth tary reaction steps is usually assumed to represent the
Next comes the question how the hydrocarbon chains grow. mechanistic pathway of the reaction. Normally, a number of
This has been and still is a matter of scientific debate. The simplifying assumptions are made regarding the catalyst sur-
conventional view is that chains grow by a reaction between face and the adsorption of species on it, for example, that only
an alkyl and a CH2 species, for example, CH3 þ CH2 to C2H5, one type of adsorption site is considered, and that these sites
and C2H5 þ CH2 to C3H7. Termination to an alkane would are homogeneously distributed over the catalyst surface, that
then occur by hydrogenation of the alkyl, and olefins would species only adsorb in a monolayer, and that adsorbed species
form by b-H abstraction from the alkyl fragment. This view do not interact apart from the involved chemical reactions. It is
originates from the work of Biloen et al., who, however, pro- usually further assumed that most reaction steps are suffi-
posed this mechanism with some reservation, and with a judi- ciently fast to reach equilibrium, but that one or more rate-
cious discussion of the assumptions involved.195,196 Ciobica determining steps exist that are relatively slow and control the
et al. found that reactions between CxHy fragments that contain overall rate of reaction.188 These assumptions allow for the
less hydrogen are energetically more favorable, making chain derivation of simple, manageable kinetic expressions, such as
growth via reactions of the type CH2 þ CH and CH¼CH2 þ CH those presented in Table 9.
more likely.197 Nevertheless, the class of mechanisms, based
on CO dissociation, formation of CHx species, and incorpora- 7.20.4.4.1 General observations regarding kinetics
tion of CHx in a growing chain, albeit with a rich variety in the Macrokinetic models are mainly developed for use in process
details, comes close to the original proposal of Fischer and modeling, yet their mechanistic importance stems from the
Tropsch, entitled the carbide mechanism (see Table 7). fact that they capture the overall behavior of the FT synthesis.
Another class of mechanisms considers CO insertion as the Considering eqn [3] from Table 9 as an example, it is seen that
step leading to chain growth (see Table 7). This line of thought the reaction rate is predicted to increase with the square root of
also has a long history, the archetype being the Pichler–Schulz the hydrogen partial pressure (everything else being constant).
mechanism.155 The most detailed mechanism has been pre- Therefore, it is said that the reaction order of hydrogen is 0.5 in
sented by Saeys and coworkers,198 who proposed that CO this kinetic model. The variation in reaction rate with CO
inserts in an adsorbed CH2, which then further hydrogenates partial pressure is more complex, since the overall reaction
to a fragment in which the C–O bond breaks, leading to a C2 order is not constant. At very low CO partial pressures or at
intermediate in which the next CO inserts. As discussed in the high reaction temperature, KCOPCO is much smaller than 1 and
section on surface science, experimental proof exists that C–O the denominator term can be ignored, yielding an overall CO
bond breaking in adsorbed species derived from ethanol, such reaction order of 1. In the other extreme, at very high CO
as ethoxy, is a facile step, even on the least reactive surface of partial pressures or at low temperatures where KCOPCO is much
cobalt, that is, the close-packed (0001) surface.166 larger than 1, the constant term in the denominator can be
At the time of writing, medio 2011, the authors believed ignored and the overall CO reaction order strives to 1.
that DFT is a highly valuable tool for getting insight into
reaction mechanisms at the level of elementary steps. It is,
however, much too early to draw definitive conclusions on Table 9 Kinetic expressions for the rate of the Fischer–Tropsch
which mechanism is prevalent in the FTS. Further, it should synthesis, rFT
be acknowledged that not all catalysts and conditions can be Equation number Rate expression
captured under one dominating mechanism, and it is even not
PH PCO
at all certain that this would be the case on one catalyst. For 1 rFT ¼ A PCO þK2H P
2 O H2 O
example, the initiation by CO dissociation might simulta- PH0:5 PCO
neously occur via direct dissociation on highly reactive parts 2 rFT ¼ A PCO þK2H P
2 O H2 O
PH0:5 PCO
of a cobalt particle, and by H-assistance on facets of moderate 3 rFT ¼ A 2
ð1þKCO PCO Þ2
reactivity. We also point to the differences between cobalt and
PH2 PCO
iron. Whereas the former is believed to operate as a metal, the 4 rFT ¼ A
ð1þKCO PCO Þ2
latter is active as a carbide, in which the intrinsically high PH0:75 PCO
0:5

5 rFT ¼ A  2
2
reactivity of the iron atoms is considerably decreased by carbon 0:5 P0:25 þK P0:5 P0:25
1þKC=OH PCO H O CO H
neighbors. The DFT literature has suggested mechanisms vary- PH0:75 PCO
0:5
2 2

ing from Mars-van Krevelen-type reactions199 (in which carbon 6 rFT ¼ A 2

ð1þKCO PCO Þ
0:5 2

atoms from the lattice become the CHx species for initiation
and chain growth) to H-assisted CO dissociation and K stands for equilibrium constant, P for partial pressure, and A is an effective rate
CO insertion as the major ingredients for hydrocarbon constant.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
552 Fischer–Tropsch Synthesis: Catalysts and Chemistry

Similarly, it can be said for eqn [1] that the reaction order in experimental approach was then followed to eliminate non-
hydrogen is constant at 1, while the overall reaction order in applicable models until ultimately a robust kinetic expression
CO varies from 0 to 1 as the conversion increases from low to remained as the preferred rate equation. A further notable
high values. feature of the FTS kinetic studies was to operate the reaction
There are distinct differences between the chemical reaction at a baseline condition, where the catalyst is known to be quite
kinetics over iron and cobalt FTS catalysts. The reaction rate stable, for most of the run. Changes to other conditions were
increases roughly linearly with pressure over iron catalysts at only made for short intervals sufficient to allow for hydrody-
constant temperature and H2/CO ratio,201 but only to an order namic steady state inside the reactor, but insufficient to effect
of around 0.5 for cobalt catalysts.202 Furthermore, CO has a changes in the intrinsic catalyst behavior. The importance of
strong inhibiting influence over cobalt catalysts to the extent this approach is related to the fact that FT catalysts, especially
that it has a significantly negative reaction order under com- those based on iron, readily respond to changes in operating
mercially relevant reaction conditions. To the contrary, CO has conditions. It is imperative to avoid reversible and irreversible
a positive effect on the rate over iron catalysts in the normal changes in the catalyst when chemical reaction kinetics is
operating regime and it has been estimated that CO will only studied.
start to negatively influence the kinetics above a partial pres- Originally within Sasol, the rate equation [1] in Table 9, by
sure of around 11 bar.203 Anderson201 was used to describe the FTS over iron. Following
The influence of water on the reaction kinetics has proven a systematic in-house study focusing on the reaction order of
to be a highly controversial topic. Historically, it was firmly hydrogen, its exponent was later reduced to a value of 0.5,
believed that water (and possibly CO2) inhibited the reaction yielding eqn [2] (Table 9). The most recent study on iron-FT
rate over iron catalysts via competitive adsorption, but more kinetics has considered the implications of the historic rate
recently it has been shown that there is no convincing evidence equations, specifically the single order denominator which
for this notion.203 The results of water co-feeding studies over implies that hydrogen reacts directly from the gas phase. By
cobalt catalysts have been inconsistent, since some have applying a second order denominator to obtain a more appro-
reported a positive and some a negative influence of water on priate Langmuir–Hinshelwood–Hougen–Watson-type equa-
the reaction rate, while others have found no effect at all. At tion where both CO and H2 first absorb onto the surface
least for alumina-supported cobalt catalysts, it appears as before reaction, and also including a constant term in the
though water in the range of 1–6 bar has no significant influ- denominator to provide for the possibility of vacant sites, it
ence on the overall rate of CO conversion, but that it does could in fact be shown that there is no statistical justification
decrease the methane selectivity.204 for including a water term in the kinetic model.207 Experiments
were designed to conclusively show that eqn [3] (Table 9) is
7.20.4.4.2 Simple macrokinetic models more accurate than the foregoing expressions. This study
For the derivation of macrokinetic models, a key question is highlighted the fact that the historic perception regarding the
whether or not there is a rate-determining step involved in the effect of water on iron-FT kinetics was self-specified by the old
formation of the monomer of chain growth. If monomer models, but never tested. Furthermore, the CO order of unity is
formation is relatively facile, then the rate of CO conversion consistent with a mechanism where CO interacts with a hydro-
and the product distribution obtained are intimately linked gen atom before being dissociated.
and must be modeled together. This typically yields a complex, The most recent cobalt kinetic study involved the derivation
implicit type model that requires an advanced numerical rou- of several rate equations to cover various reaction schemes of
tine to solve. However, in order to keep these models manage- CO hydrogenation.202 Models assuming hydrogen-assisted CO
able, questionable simplifying assumptions are often made, for dissociation, such as eqn [4] that was originally proposed by
example, the assumption by Yang et al.205 that the monomer is Yates and Satterfield,208 generally described the measured data
in thermodynamic equilibrium with the gas phase concentra- poorly and could be eliminated early on in the study. Figure 27
tions of CO, H2, and water. Furthermore, such models require illustrates that eqn [4] underestimates the reaction rate at low
a large number of parameters that are inevitably highly cross- CO partial pressure, but overestimates it at higher CO partial
correlated, implying that it is virtually impossible to accurately pressures. Ultimately, after further work to distinguish between
estimate their values. If, however, there is a rate-determining those models where CO first dissociates before it is hydroge-
step in the formation of the monomer, the overall reaction rate nated, eqn [5] (Table 9) was the only rate expression that could
can be decoupled from the product distribution. In such a case, not be eliminated. Therefore, it was selected as the most appro-
the overall rate of CO conversion is determined by the CO priate kinetic model. It should be noted that this expression can
hydrogenation reaction (monomer formation), while the be very closely approximated by eqn [6], which contains one
product distribution is determined by the polymerization model parameter less and is thus preferably used from a practi-
part of the FT reaction (i.e., the competition between chain cal perspective. Figure 27 shows that the preferred equation [6]
growth and desorption). The steady state isotopic transient does not suffer from the same systematic errors as eqn [4], since
kinetic analysis study of van Dijk206 has indeed provided it is reasonably accurate across a range of CO partial pressures.
microkinetic support for the notion that there is a rate-
determining step in the formation of the monomer. 7.20.4.4.3 Selectivity modeling
The approach followed during FTS kinetic studies by Botes The vast number of components in the FT product slate does
et al.202,207 has been to consider different reaction schemes and not allow for the prediction of individual product selectivities;
rate-determining steps in the formation of the monomer to consequently, the product spectrum is rather represented by a
obtain a variety of explicit rate expressions. A systematic product characterization model with a limited number of

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 553

1.5

Equation 4

PH2PCO
rFT = A
(1 + KCOPCO)2

rFT (measured) / rFT (predicted) 1


Equation 6
P 0H.75P 0.5
CO
2
rFT = A
(1 + KCOP 0.5
CO
)2

0.5
0 1 2 3 4 5 6 7 8
CO partial pressure (bar)

Figure 27 Performance of rival kinetic expressions as a function of CO partial pressure for a data series where the CO flow rate into a slurry reactor was
varied while that of H2 was maintained constant.

parameters. After correlating these parameters with process with certainty. For example, and as described before, DFT calcu-
conditions, an explicit selectivity model can be obtained. The lations have shown that the coupling reactions of CHx fractions
simplest model for describing selectivity is the Anderson– are reasonably facile, which lends support to the steps of chain
Schulz–Flory distribution, with the chain growth probability growth as proposed by the carbide mechanism. However, DFT
(a-value) as the only parameter. Approaches have been pro- calculations have also shown that each of the steps required for
posed to account for deviations from the ideal distribution. the CO insertion mechanism is energetically feasible, while the
The double-a model by Donnelly et al.209 assumes that two surface science approach has demonstrated experimentally that
types of catalytic sites or two types of mechanisms simulta- the scission of the C–O bond of a CO molecule into which an
neously form the observed product spectrum. However, the alkyl group has been inserted is a facile reaction, even on a low
three model parameters have a high degree of covariance reactivity cobalt surface.166
when estimated from experimental data. Furthermore, neither The current inability to clearly discriminate between rival
the C1 and C2 selectivities, nor the olefin content of the prod- mechanisms partly stems from the overlap between the pro-
uct spectrum can be predicted. Some of these limitations have posed reaction pathways. For instance, the two most popular
been addressed by considering a chain length-dependent mechanisms are not mutually exclusive with respect to each
desorption model, which assumes that termination by desorp- other. The carbide mechanism has to assume direct CO inser-
tion becomes increasingly more difficult as the chain length tion as a termination step at least in order to explain the
increases.17 It has also been reported that the chain length observed formation of oxygenates in the FTS. On the other
effects in the FT product spectrum, in particular the positive hand, the initiation step in the CO insertion mechanism is
deviation of methane and the negative deviation of ethylene, similar as for the carbide mechanism. The most plausible con-
can be explained by symmetry effects as accounted for by the clusion currently is therefore that a variety of reaction pathways
single event kinetic theory.210 Some have ascribed the chain simultaneously contribute to the overall synthesis. The question
length-dependent deviations to secondary olefin reactions, but still remains though whether one pathway is dominant over the
many concerns remain over this approach.203 These include rest and individually determines the bulk of the observed behav-
the observation that secondary olefin reactions are much less ior of the system, or whether two or more parallel pathways
facile (almost negligible) over iron catalysts compared to have similar contributions to the overall kinetics. The answer to
cobalt catalysts, yet the bend in the Anderson–Schulz–Flory this question may not even be absolute, as it may depend on
graph is much more pronounced with iron catalysts. what aspect of the reaction is of interest. Hypothetically speak-
ing, if the carbide mechanism predominates in the formation of
olefins and paraffins (the main products of the synthesis), it
7.20.4.5 Mechanistic and Kinetic Implications
may well be accurate to describe the overall rate of syngas
Despite the large number of kinetic and mechanistic studies on conversion by only considering this mechanism. However,
FTS, there is still substantial uncertainty regarding the most even in such an event, CO insertion cannot be ignored if the
relevant steps in the reaction pathway(s). A variety of elementary object is to model oxygenate formation.
reaction steps have been proven to be realistically possible under Despite all the foregoing uncertainty, some consistencies
typical FTS conditions, while very few steps could be eliminated have also emerged from the studies performed in the various

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
554 Fischer–Tropsch Synthesis: Catalysts and Chemistry

disciplines. For example, the results of DFT calculations suggest barrels of synfuels is undoubtedly significant, particularly
that unassisted CO dissociation readily occurs on the more locally where the production takes place.
open (high reactivity) metal surfaces, while hydrogen-assisted The technology has much potential for wider use, for exam-
CO dissociation would be required on the close packed (low ple, in emerging economies, or at a smaller scale in the utili-
reactivity) surfaces.179,181,183,185,186 As described in the section zation of biomass. Interesting applications of FTS have been
on surface science above, unassisted CO dissociation is not proposed for conversion of remote natural gas at off-shore oil
easy on close packed surfaces, while it is facile over open production locations.
surfaces. It has further been found that HxCO species are Both GTL and BTL can be important tools in strategies
more inclined to dehydrogenate than to undergo C–O bond aimed at reduction of CO2 emissions. CTL technology is clearly
cleavage on cobalt surfaces. During a macrokinetic study on an disadvantaged here, and will in the future have to be combined
actual cobalt-FT catalyst, models assuming hydrogen-assisted with CO2 sequestration technology. The rapid increase in dis-
CO dissociation failed comprehensively, while the preferred coveries of shale gas (in, for example, USA and Canada) can
model based on unassisted CO dissociation could describe the also provide the GTL industry with a significant boost.
experimental data over a range of commercially relevant con- Although proven technology, FTS continues to pose chal-
ditions. Together all these findings suggest that, in the case of lenges from an industrial perspective. Stability improvement of
the cobalt-based FT synthesis, the main pathway in the conver- the catalysts is an important aspect, but also selectivity
sion of CO to a CHx species proceeds via unassisted cleavage of improvement would be very advantageous. Economically one
the C–O bond. would like to have the highest possible Cþ 5 and the lowest
To the contrary, it is known that the carbiding of iron cata- possible CH4 selectivity, because recycling of CH4 means that
lysts substantially decreases the reactivity of iron surfaces. There- the carbon atoms involved have to go through the expensive
fore, one may well expect hydrogen-assisted CO dissociation to syngas generation more than once, with the associated effi-
predominate over iron-FT catalysts,199 which are in the carbided ciency losses. Syngas production is the most expensive part of
state under actual synthesis conditions. The most preferred a GTL plant; it accounts for 40–60% of the capital investments.
macrokinetic model for iron is indeed consistent with a CO Increased research efforts on reducing the costs of syngas pro-
dissociation step that occurs via interaction with hydrogen. duction will make XTL projects even more viable. Although
A further consistency that is steadily emerging relates to the new XTL facilities require large capital investments and are
most likely nature of the species responsible for chain propaga- dependent on the price ratio of crude oil to natural gas, in
tion in terms of the carbide mechanism. Originally, it was the long term they are expected to be economically successful.
believed (not necessarily based on strong evidence) that these From a more academic perspective, understanding the
species are quite saturated with hydrogen, that is, that the grow- mechanism of the FTS has been and will be a challenge. It is
ing intermediate is a CH3–CH2  CH2 species, while the mono- more and more realized that mechanisms may differ with
mer being added is a CH2 species.195 DFT calculations have conditions and catalysts. It is highly unlikely that one unique
shown that reactions between intermediates that are leaner in mechanism can account for all different forms of FTS. Molec-
hydrogen (e.g., CH¼CH2 and CH) are energetically more favor- ular modeling represents a very important tool for getting
able than reactions between more hydrogen-saturated mechanistic insight, but the problem is that experimental val-
species.197 In line with this, it has been found during surface idation of its predictions at the level of elementary steps is very
science experiments the coupling of two CH species to from difficult to achieve, as the opportunities for relevant surface
acetylene is facile over nickel catalysts.168,211Further support is science experiments are limited. Mechanistic studies aimed at
provided by the steady state isotopic transient kinetic analysis describing FTS selectivity from first principles are in their
(SSITKA) study of Govender,212 who concluded from H–D infancy and have a long way to go before accurate predictions
switching experiments that the C2H species is the only abundant can be expected.
C2 intermediate on the fully carbided, working iron-FT catalyst Describing the physical/chemical state of the catalysts
surface. Therefore, even though the carbide mechanism as a under reaction conditions is another field where significant
whole cannot be discarded as a prominent reaction pathway progress has been booked, but major advances would still be
for the FTS, it seems unlikely that it proceeds in the form that very welcome. The advent of in situ imaging tools in combina-
was originally proposed. This has particular significance for the tion with realistic catalysts,213 as well as the use of planar
termination toward olefins, since it implies that a hydrogen model catalysts in simulated environments,153 has proven
abstraction is not required (possibly even a hydrogen addition). promising and will almost certainly lead to improved insight
in the relation between catalyst properties on the nanoscale
and performance in the reaction.
7.20.5 Conclusion The FTS is therefore expected to remain an inspiring source
of industrial and academic research for many years to come.
The FTS represents proven technology, which has secured its For a related chapter in this Comprehensive, we refer to
position in modern energy technology. Originally used to Chapter 7.01
convert coal into liquid fuels, nowadays the emphasis is on
monetizing natural gas, by converting it to diesel fuel, waxes,
and naphtha. It is expected that some 500 000 barrels of fuel
References
per day will be produced using Fischer–Tropsch technology by
2013. Although small in comparison to the 85 million barrels 1. Fischer, F.; Tropsch, H. Brennst. Chem. 1923, 4, 276–285.
of crude oil that are produced daily, the 0.5 million daily 2. Fischer, F.; Tropsch, H. Brennst. Chem. 1926, 7, 97–104.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 555

3. Storch, H.; Golumbic, N.; Anderson, R. B. The Fischer–Tropsch and Related 48. Kock, A. J. H. M.; Geus, J. W. Prog. Surf. Sci. 1985, 20, 165–272.
Syntheses. Wiley: New York, 1951. 49. de Smit, E.; Cinquini, F.; Beale, A. M.; Safonova, O. V.; van Beek, W.; Sautet, P.;
4. van der Laan, G. P.; Beenackers, A. A. C. M. Catal. Rev. 1999, 41, 255–318. Weckhuysen, B. M. J. Am. Chem. Soc. 2010, 132, 14928–14941.
5. Kolbel, H.; Ralek, M. Catal. Rev. 1980, 21, 225–274. 50. Raupp, G. B.; Delgass, W. N. J. Catal. 1979, 58, 348–360.
6. Dry, M. E.; Hoogendoorn, J. C. Catal. Rev. 1981, 23, 265–278. 51. Niemantsverdriet, J. W.; Van der Kraan, A. M.; Van Dijk, W. L.; Van der Baan, H. S.
7. Roferdepoorter, C. K. Chem. Rev. 1981, 81, 447–474. J. Phys. Chem. 1980, 84, 3363–3370.
8. Iglesia, E., Reyes, S. C., Madon, R. J., Soled, S. L., Eds.; Advances in Catalysis; 52. Dry, M. E. In Studies in Surface Science and Catalysis; Steynberg, A. P.,
1993; Vol. 39, pp 221–302. Dry, M. E., Eds.; 152, Elsevier: Amsterdam, 2004; pp 533–600.
9. Steynberg, A. P.; Dry, M. E. Fischer–Tropsch Technology; Elsevier: Amsterdam, 53. Luo, M. S.; Hamdeh, H.; Davis, B. H. Catal. Today 2009, 140, 127–134.
2004; Vol. 152. 54. Herranz, T.; Rojas, S.; Perez-Alonso, F. J.; Ojeda, M.; Terreros, P.; Fierro, J. L. G.
10. Rostrup-Nielsen, J. R., Sehested, J., Norskov, J. K., Eds.; Advances in Catalysis; J. Catal. 2006, 243, 199–211.
2002; Vol. 47, pp 65–139. 55. Dry, M. E. In Catalysis, Science and Technology; Anderson, J. R., Boudart, M.,
11. Dancuart, L. P.; de Haan, R.; de Klerk, A. In Studies in Surface Science and Eds.; Springer-Verlag: New York, 1981; Vol. 1, pp 159–255.
Catalysis; Steynberg, A. P., Dry, M. E., Eds.; 152, Elsevier, 2004; pp 482–532. 56. Luo, M. S.; Davis, B. H. Fuel Process. Technol. 2003, 83, 49–65.
12. De Klerk, A. Fischer–Tropsch Refining. Wiley-VCH: Weinheim, 2011. 57. Fiato, R. A.; Soled, S. L. Fischer–Tropsch Hydrocarbon Synthesis with High
13. Claeys, M.; van Steen, E. Fischer–Tropsch Technology; Elsevier: Amsterdam, Surface Area Copper and Potassium Promoted Reduced-Carbided Iron/
2004; Vol. 152. Manganese Spinels. US Patent 4621102A, 1986.
14. Schulz, H.; Vansteen, E.; Claeys, M. In Natural Gas Conversion II 1994; Vol. 81, 58. Du Toit, E. Ph.D. Thesis, University of the North West, 2002.
pp 455–460. 59. Niemantsverdriet, J. W.; van der Kraan, A. M. J. Catal. 1981, 72, 385–388.
15. Claeys, M.; Van Steen, E. In Fischer–Tropsch Technology; Studies in Surface 60. Bromfield, T. C.; Visagie, R. Chromium Oxide Incorporation into Precipitated Iron-
Science: An Catalysis; Steynberg, A. P., Dry, M. E., Eds.; Elsevier: Amsterdam, Based Fischer–Tropsch Catalysts for Increased Production of Oxygenates and
2004; Vol. 152, Chapter 8. Branched Hydrocarbons. Patent WO 2005049765A1, 2005.
16. Botes, F. G.; Govender, N. S. Energy Fuel 2007, 21, 3095–3101. 61. van Steen, E.; Claeys, M. Chem. Eng. Technol. 2008, 31, 655–666.
17. Botes, F. G. Energy Fuel 2007, 21, 1379–1389. 62. Tsakoumis, N. E.; Ronning, M.; Borg, O.; Rytter, E.; Holmen, A. Catal. Today
18. Davis, B. H. Ind. Eng. Chem. Res. 2007, 46, 8938–8945. 2010, 154, 162–182.
19. Khodakov, A. Y.; Chu, W.; Fongarland, P. Chem. Rev. 2007, 107, 1692–1744. 63. Morales, F.; Weckhuysen, B. M. Catalysis 2006, 19, 1–40.
20. Dry, M. E. Catal. Lett. 1991, 7, 241–251. 64. van de Loosdrecht, J.; Bazhinimaev, B.; Dalmon, J. A.; Niemantsverdriet, J. W.;
21. de Smit, E.; Weckhuysen, B. M. Chem. Soc. Rev. 2008, 37, 2758–2781. Tsybulya, S. V.; Saib, A. M.; van Berge, P. J.; Visagie, J. L. Catal. Today 2007,
22. Steynberg, A. P.; Dry, M. E.; Davis, M. E.; Davis, B. H.; Breman, B. B. Stud. Surf. 123, 293–302.
Sci. Catal. 2004, 152, 64–195. 65. Saib, A. M.; Moodley, D. J.; Ciobica, I. M.; Hauman, M. M.; Sigwebela, B. H.;
23. Van Berge, P. J.; Van De Loosdrecht, J.; Caricato, E. A.; Barradas, S. Process for Weststrate, C. J.; Niemantsverdriet, J. W.; van de Loosdrecht, J. Catal. Today
Producing Hydrocarbons from a Synthesis Gas, and Catalysts Therefore. Patent 2010, 154, 271–282.
WO 9942214A1, 1999. 66. Claeys, M.; van Steen, E. In Studies in Surface Science and Catalysis;
24. Davis, B. H. Catal. Today 2002, 71, 249–300. Steynberg, A. P., Dry, M. E., Eds.; 152, Elsevier: Amsterdam, 2004;
25. Steynberg, A. P.; Espinoza, R. L.; Jager, B.; Vosloo, A. C. Appl. Catal. A: Gen. pp 601–680.
1999, 186, 41–54. 67. Iglesia, E.; Reyes, S. C.; Madon, R. J.; Soled, S. L. Adv. Catal. 1993, 39,
26. Espinoza, R. L.; Steynberg, A. P.; Jager, B.; Vosloo, A. C. Appl. Catal. A: Gen. 221–302.
1999, 186, 13–26. 68. Diehl, F.; Khodakov, A. Y. Oil Gas Sci. Technol. 2009, 64, 11–24.
27. Geerlings, J. J. C.; Wilson, J. H.; Kramer, G. J.; Kuipers, H.; Hoek, A.; 69. Bezemer, G. L.; Bitter, J. H.; Kuipers, H.; Oosterbeek, H.; Holewijn, J. E.; Xu, X. D.;
Huisman, H. M. Appl. Catal. A: Gen. 1999, 186, 27–40. Kapteijn, F.; van Dillen, A. J.; de Jong, K. P. J. Am. Chem. Soc. 2006,
28. Deshmukh, S. R.; Tonkovich, A. L. Y.; McDaniel, J. S.; Schrader, L. D.; 128, 3956–3964.
Burton, C. D.; Jarosch, K. T.; Simpson, A. M.; Kilanowski, D. R.; LeViness, S. 70. Jacobs, G.; Patterson, P. M.; Zhang, Y. Q.; Das, T.; Li, J. L.; Davis, B. H. Appl.
Biofuels 2011, 2, 315–324. Catal. A: Gen. 2002, 233, 215–226.
29. Visconti, C. G.; Tronconi, E.; Lietti, L.; Groppi, G.; Forzatti, P.; Cristiani, C.; 71. Karaca, H.; Hong, J. P.; Fongarland, P.; Roussel, P.; Griboval-Constant, A.;
Zennaro, R.; Rossini, S. Appl. Catal. A: Gen. 2009, 370, 93–101. Lacroix, M.; Hortmann, K.; Safonova, O. V.; Khodakov, A. Y. Chem. Commun.
30. Sabatier, P.; Senderens, J. B. Hebd. Seances Acad. Sci. 1902, 134, 514. 2010, 46, 788–790.
31. Fischer, F.; Tropsch, H. Patent DE 484 337, 1925. 72. Kitakami, O.; Sato, H.; Shimada, Y.; Sato, F.; Tanaka, M. Phys. Rev. B 1997, 56,
32. Liu, Z.; Shi, S.; Li, Y. Chem. Eng. Sci. 2010, 65, 12–17. 13849–13854.
33. Dry, M. E. Catal. Today 2002, 71, 227–241. 73. Enache, D. I.; Rebours, B.; Roy-Auberger, M.; Revel, R. J. Catal. 2002, 205,
34. Sie, S. T. Rev. Chem. Eng. 1998, 14, 109–157. 346–353.
35. Oukaci, R.; Singleton, A. H.; Goodwin, J. G. Appl. Catal. A: Gen. 1999, 186, 74. Iglesia, E.; Soled, S. L.; Fiato, R. A.; Via, G. H. J. Catal. 1993, 143, 345–368.
129–144. 75. Feller, A.; Claeys, M.; van Steen, E. J. Catal. 1999, 185, 120–130.
36. Sasol Financial Report July–December 2010. http://www.sasol.com. 76. Moradi, G. R.; Basir, M. M.; Taeb, A.; Kiennemann, A. Catal. Commun. 2003,
37. Brown, A. Pearl GTL Presentation, XTL Summit, June 2011. 4, 27–32.
38. Steynberg, A. P. In Studies in Surface Science and Catalysis; Steynberg, A. P., 77. Tan, K. F.; Chang, J.; Borgna, A.; Saeys, M. J. Catal. 2011, 280, 50–59.
Dry, M. E., Eds.; 152, Elsevier: Amsterdam, 2004; pp 1–63. 78. Rytter, E.; Skagseth, T. H.; Eri, S.; Sjastad, A. O. Ind. Eng. Chem. Res. 2010,
39. Penning, R. In New Developments in Synthetic Fuels, CTL/GTL Conference 2010, 49, 4140–4148.
Brisbane, Australia. 79. Beuther, H.; Kobylinski, T. P.; Kibby, C. L.; Pannell, R. B. Synthesis Gas
40. Davis, B. H. Catal. Today 2003, 84, 83–98. Conversion Using Ruthenium-Promoted Cobalt Catalyst Prepared by
41. Schwertmann, U.; Cornell, R. M. Iron Oxides in the Laboratory, Preparation and Nonaqueous Impregnation. US Patent 4585798A, 1986.
Characterization. Wiley-VCH: Weinheim, 2000. 80. Beuther, H.; Kibby, C. L.; Kobylinski, T. P.; Pannell, R. B. Fluid Bed Catalyst for
42. Smith, D. F.; Hawk, C. O.; Golden, P. L. J. Am. Chem. Soc. 1930, 52, Synthesis Gas Conversion and Its Utilization for Preparation of Diesel Fuel. US
3221–3232. Patent 4413064A, 1983.
43. Bromfield, T. C.; Botes, F. G.; Visagie, R.; Espinoza, R.; Gibson, P.; Van Lawson, 81. Beuther, H.; Kibby, C. L.; Kobylinski, T. P.; Pannell, R. B. Conversion of Synthesis
K. H. Hydrocarbon Synthesis Catalyst and Process. US Patent 6844370, 2002. Gas to Diesel Fuel and Gasoline. US Patent 4605680A, 1986.
44. Hayakawa, H.; Tanaka, H.; Fujimoto, K. Appl. Catal. A: Gen. 2006, 310, 82. Eri, S.; Kinnari, K. J.; Schanke, D.; Hilmen, A.-M. Preparation and Use of
24–30. Promoted Cobalt Catalyst with Low Surface Area Alumina for Fischer–Tropsch
45. O’Brien, R. J.; Xu, L. G.; Spicer, R. L.; Bao, S. Q.; Milburn, D. R.; Davis, B. H. Reaction with High Olefin Selectivity. Patent WO 2002047816A1, 2002.
Catal. Today 1997, 36, 325–334. 83. Saib, A. M.; Claeys, M.; van Steen, E. Catal. Today 2002,
46. Luo, M. S.; O’Brien, R. J.; Bao, S. Q.; Davis, B. H. Appl. Catal. A: Gen. 2003, 239, 71, 395–402.
111–120. 84. Van Berge, P. J.; Van De Loosdrecht, J.; Barradas, S. Method of Treating an
47. Niemantsverdriet, J. W. Spectroscopy in Catalysis; An Introduction, 3rd ed.; Untreated Catalyst Support, and Forming a Catalyst Precursor and Catalyst from
Wiley-VCH: Weinheim, 2007. the Treated Support. Patent EP 1 303 350 B1, 2000.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
556 Fischer–Tropsch Synthesis: Catalysts and Chemistry

85. Van Berge, P. J.; Van De Loosdrecht, J.; Barradas, S. Production of Fischer– 122. Moodley, D. J.; Saib, A. M.; van de Loosdrecht, J.; Welker-Nieuwoudt, C. A.;
Tropsch Synthesis Produced Wax. Patent EP 1 432 778 B1, 2001. Sigwebela, B. H.; Niemantsverdriet, J. W. Catal. Today 2011, 171, 192–200.
86. Wei, D. G.; Goodwin, J. G.; Oukaci, R.; Singleton, A. H. Appl. Catal. A: Gen. 2001, 123. Moodley, D. J.; van de Loosdrecht, J.; Saib, A. M.; Niemantsverdriet, J. W. Chem.
210, 137–150. Ind. 2010, 128, 49–81.
87. Xiong, H. F.; Motchelaho, M. A. M.; Moyo, M.; Jewell, L. L.; Coville, N. J. J. Catal. 124. Ciobica, I. M.; van Santen, R. A.; van Berge, P. J.; van de Loosdrecht, J. Surf. Sci.
2011, 278, 26–40. 2008, 602, 17–27.
88. Stranges, A. N. Germany’s Synthetic Fuel Industry 1927–1945. Presented at the 125. Soled, S. L.; Iglesia, E.; Fiato, R. A.; Baumgartner, J. E.; Vroman, H.; Miseo, S.
AIChE 2003, New Orleans, 2003. http://www.fischertropsch.org. Top. Catal. 2003, 26, 101–109.
89. Keyser, M. J.; Everson, R. C.; Espinoza, R. L. Appl. Catal. A: Gen. 1998, 171, 126. Hilmen, A. M.; Schanke, D.; Holmen, A. In Natural Gas Conversion IV;
99–107. dePontes, M., Espinoza, R. L., Nicolaides, C. P., Scholtz, J. H., Scurrell, M. S.,
90. Co-Precipitated Cobalt–Zinc Catalysts for Fischer–Tropsch Reaction or Eds.; Elsevier: Amsterdam, 1997; Vol. 107, pp 237–242.
Functional Group Hydrogenation. Patent EP1358934A1, 2003. 127. Rothaemel, M.; Hanssen, K. F.; Blekkan, E. A.; Schanke, D.; Holmen, A. Catal.
91. Puskas, I.; Fleisch, T. H.; Full, P. R.; Kaduk, J. A.; Marshall, C. L.; Meyers, B. L. Today 1997, 38, 79–84.
Appl. Catal. A: Gen. 2006, 311, 146–154. 128. Craje, M. W. J.; van der Kraan, A. M.; van de Loosdrecht, J.; van Berge, P. J. Catal.
92. Joustra, A. H.; Scheffer, B. Process for the Preparation of Alumina-Based Today 2002, 71, 369–379.
Extrudates. Patent EP 455307A1, 1991. 129. Kiss, G.; Kliewer, C. E.; DeMartin, G. J.; Culross, C. C.; Baumgartner, J. E.
93. Neimark, A. V.; Kheifets, L. I.; Fenelonov, V. B. Ind. Eng. Chem. Prod. Res. Dev. J. Catal. 2003, 217, 127–140.
1981, 20, 439–450. 130. Das, T. K.; Jacobs, G.; Patterson, P. M.; Conner, W. A.; Li, J. L.; Davis, B. H. Fuel
94. Kheifets, L. I.; Neimark, A. V.; Fenelonov, V. B. Kinet. Catal. 1979, 20, 2003, 82, 805–815.
626–632. 131. Li, J. L.; Jacobs, G.; Zhang, Y. Q.; Das, T.; Davis, B. H. Appl. Catal. A: Gen. 2002,
95. Neimark, A. V.; Fenelonov, V. B.; Heifets, L. I. React. Kinet. Catal. Lett. 1976, 5, 223, 195–203.
67–72. 132. Storsaeter, S.; Borg, O.; Blekkan, E. A.; Holmen, A. J. Catal. 2005, 231, 405–419.
96. Van Berge, P. J. In Scaling Up of an Alumina Supported Cobalt Slurry Phase 133. Huffman, G. P.; Shah, N.; Zhao, J. M.; Huggins, F. E.; Hoost, T. E.; Halvorsen, S.;
Fischer–Tropsch Catalyst Preparation, CatCon – World Wide Catalyst Industry Goodwin, J. G. J. Catal. 1995, 151, 17–25.
Conference, Houston, TX, USA, June 12–13, 2000. 134. Bezemer, G. L.; Remans, T. J.; van Bavel, A. P.; Dugulan, A. I. J. Am. Chem. Soc.
97. Van Berge, P. J.; Van De Loosdrecht, J.; Caricato, E. A.; Barradas, S.; Sigwebela, 2010, 132, 8540–8541.
B. H. Impregnation Process for Catalysts. Patent WO 2000020116A1, 2000. 135. Yan, Z.; Wang, Z. J.; Bukur, D. B.; Goodman, D. W. J. Catal. 2009, 268, 196–200.
98. Soled, S. L.; Baumgartner, J. E.; Reyes, S. C.; Iglesia, E. In Preparation of 136. den Breejen, J. P.; Sietsma, J. R. A.; Friedrich, H.; Bitter, J. H.; de Jong, K. P.
Catalysts VI: Scientific Bases for the Preparation of Heterogeneous Catalysts, J. Catal. 2010, 270, 146–152.
Studies in Surface Science and Catalysis, 1995; Vol. 91, pp 989–997. 137. Ronning, M.; Tsakoumis, N. E.; Voronov, A.; Johnsen, R. E.; Norby, P.; van
99. de Jong, K. P. Deposition Precipitation onto Pre-shaped Carrier Bodies, Beek, W.; Borg, O.; Rytter, E.; Holmen, A. Catal. Today 2010, 155, 289–295.
Possibilities and Limitations; Elsevier: Amsterdam, 1991; Vol. 63. 138. Li, J. L.; Zhan, X. D.; Zhang, Y. Q.; Jacobs, G.; Das, T.; Davis, B. H. Appl. Catal. A:
100. Boutonnet, M.; Jaras, S.; Logdberg, S., Method for Depositing Metal Particles on Gen. 2002, 228, 203–212.
a Support. Patent EP 1 985 361, 2008. 139. British Intelligence Objectives Sub-Committee, Interrogation of Dr Otto Roelen of
101. van de Loosdrecht, J.; Barradas, S.; Caricato, E. A.; Ngwenya, N. G.; Ruhrchemie A.G., B.I.O.S. Final Report No. 447; Item no 30 (1945). http://www.
Nkwanyana, P. S.; Rawat, M. A. S.; Sigwebela, B. H.; van Berge, P. J.; fischer-tropsch.org.
Visagie, J. L. Top. Catal. 2003, 26, 121–127. 140. Zonnevylle, M. C.; Geerlings, J. J. C.; van Santen, R. A. Surf. Sci. 1990, 240,
102. Wolters, M.; Munnik, P.; Bitter, J. H.; De Jongh, P. E.; De Jong, K. P. Method for 253–262.
Producing a Supported Metal Nitrate. Patent WO 2010109216A1, 2010. 141. Tan, K. F.; Xu, J.; Chang, J.; Borgna, A.; Saeys, M. J. Catal. 2010, 274, 121–129.
103. Soled, S.L.; Baumgartner, J.E.; Reyes, S.C.; Iglesia, E.; Poncelet, G., J. M. B. D. P. 142. Saeys, M.; Tan, K. F.; Chang, J.; Borgna, A. Ind. Eng. Chem. Res. 2010, 49,
A. J. a. P. G. In Studies in Surface Science and Catalysis; Elsevier, 1995; Vol. 91; 11098–11100.
pp 989–997. 143. Jacobs, G.; Sarkar, A.; Ji, Y.; Luo, M.; Dozier, A.; Davis, B. H. Ind. Eng. Chem.
104. Hoek, A.; Moors, J. H. Catalyst Activation and Rejuvenation Process. Patent WO Res. 2008, 47, 672–680.
9717137A1, 1997. 144. Tavasoli, A.; Abbaslou, R. M. M.; Dalai, A. K. Appl. Catal. A: Gen. 2008, 346,
105. Behrmann, W. C.; Davis, S. M.; Mauldin, C. H. Method for Preparing Cobalt- 58–64.
Containing Hydrocarbon Synthesis Catalyst. Patent WO 9206784A1, 1992. 145. Zhou, W.; Chen, J. G.; Fang, K. G.; Sun, Y. H. Fuel Process. Technol. 2006, 87,
106. Oosterbeek, H. Phys. Chem. Chem. Phys. 2007, 9, 3570–3576. 609–616.
107. van Berge, P. J.; Barradas, S.; van de Loosdrecht, J.; Visagie, J. L. Erdol Erdgas 146. Bartholomew, C. H. Appl. Catal. A: Gen. 2001, 212, 17–60.
Kohle 2001, 117, 138–142. 147. Madon, R. J.; Shaw, H. Catal. Rev. 1977, 15, 69–106.
108. Iglesia, E. Appl. Catal. A 1997, 161, 59–78. 148. Liu, Z. T.; Zhou, J. L.; Zhang, B. J. J. Mol. Catal. 1994, 94, 255–261.
109. Barbier, A.; Tuel, A.; Arcon, I.; Kodre, A.; Martin, G. A. J. Catal. 2001, 200, 149. Bartholomew, C. H.; Bowman, R. M. Appl. Catal. 1985, 15, 59–67.
106–116. 150. Leviness, S. C.; Mart, C. J.; Behrmann, W. C.; Hsia, S. J.; Neskora, D. R. Slurry
110. Bian, G. Z.; Fujishita, N.; Mochizuki, T.; Ning, W. S.; Yamada, M. Appl. Catal. A: Hydrocarbon Synthesis Process with Increased Catalyst Life. Patent WO
Gen. 2003, 252, 251–260. 9850487A1, 1998.
111. Martinez, A.; Prieto, G. J. Catal. 2007, 245, 470–476. 151. Wilson, J.; De Groot, C. J. Phys. Chem. 1995, 99, 7860–7866.
112. Prieto, G.; Martinez, A.; Concepcion, P.; Moreno-Tost, R. J. Catal. 2009, 266, 152. Luo, M. S.; Davis, B. H. In Catalyst Deactivation 2001, Proceedings, Studies in
129–144. Surface Science and Catalysis, 2001; Vol. 139, pp 133–140.
113. Borg, O.; Dietzel, P. D. C.; Spjelkavik, A. I.; Tveten, E. Z.; Walmsley, J. C.; 153. Thune, P. C.; Weststrate, C. J.; Moodley, P.; Saib, A. M.; van de Loosdrecht, J.;
Diplas, S.; Eri, S.; Holmen, A.; Ryttera, E. J. Catal. 2008, 259, 161–164. Miller, J. T.; Niemantsverdriet, J. W. Catal. Sci. Technol. 2011, 1, 689–697.
114. Iglesia, E. Appl. Catal. A: Gen. 1997, 161, 59–78. 154. Reynhout, M. J. Process for Regenerating a Cobalt Catalyst. Patent EP
115. Martinez, A.; Rollan, J.; Arribas, M. A.; Cerqueira, H. S.; Costa, A. F.; S- 1920836A1, 2008.
Aguiar, E. F. J. Catal. 2007, 249, 162–173. 155. Pichler, H.; Schulz, H. Chemie Ingenieur Technik 1970, 42, 1162–1174.
116. Fischer, N.; van Steen, E.; Claeys, M. Catal. Today 2011, 171, 174–179. 156. Lahtinen, J.; Vaari, J.; Kauraala, K. Surf. Sci. 1998, 418, 502–510.
117. Yang, J.; Tveten, E. Z.; Chen, D.; Holmen, A. Langmuir 2010, 26, 16558–16567. 157. Beitel, G. A.; Laskov, A.; Oosterbeek, H.; Kuipers, E. W. J. Phys. Chem. 1996,
118. Saib, A. M.; Borgna, A.; de Loosdrecht, J. V.; van Berge, P. J.; 100, 12494–12502.
Niemantsverdriet, J. W. Appl. Catal. A: Gen. 2006, 312, 12–19. 158. Habermehl-Cwirzen, K. M. E.; Kauraala, K.; Lahtinen, J. Phys. Scripta 2004,
119. Reynhout, M. J. Cobalt-Based Hydrocarbon Synthesis Catalysts Prepared from T108, 28–32.
Solid Solution Mixture of Metal Compound Precursors. Patent WO 159. Ernst, K. H.; Schwarz, E.; Christmann, K. J. Chem. Phys. 1994, 101, 5388–5401.
2008061970A2, 2008. 160. Bridge, M. E.; Comrie, C. M.; Lambert, R. M. J. Catal. 1979, 58, 28–33.
120. White Paper – Fischer–Tropsch Catalyst Test on Coal Derived Synthesis Gas, 161. Prior, K. A.; Schwaha, K.; Lambert, R. M. Surf. Sci. 1978, 77, 193–208.
Syntroleum Corporation. http://www.syntroleum.com. 162. Geerlings, J. J. C.; Zonnevylle, M. C.; Degroot, C. P. M. Surf. Sci. 1991, 241,
121. Moodley, D. J.; van de Loosdrecht, J.; Saib, A. M.; Overett, M. J.; Datye, A. K.; 315–324.
Niemantsverdriet, J. W. Appl. Catal. A: Gen. 2009, 354, 102–110. 163. Papp, H. Surf. Sci. 1985, 149, 460–470.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
Author's personal copy
Fischer–Tropsch Synthesis: Catalysts and Chemistry 557

164. Nowitzki, T.; Borchert, H.; Jurgens, B.; Risse, T.; Zielasek, V.; Baumer, M. 188. Chorkendorff, I.; Niemantsverdriet, J. W. Concepts of Modern Catalysis and
ChemPhysChem 2008, 9, 729–739. Kinetics. Wiley-VCH: Weinheim, 2003.
165. Habermehl-Cwirzen, K.; Lahtinen, J.; Hautojarvi, P. Surf. Sci. 2005, 598, 189. Ciobica, I. M.; van Santen, R. A. J. Phys. Chem. B 2002, 106, 6200–6205.
128–135. 190. Sorescu, D. C. Phys. Rev. B 2006, 73.
166. Weststrate, C. J.; Gericke, H. J.; Verhoeven, M.; Ciobica, I. M.; Saib, A. M.; 191. Lo, J. M. H.; Ziegler, T. J. Phys. Chem. C 2007, 111, 11012–11025.
Niemantsverdriet, J. W. J. Phys. Chem. Lett. 2010, 1, 1767–1770. 192. Govender, A. Towards a Mechanism for the Fischer–Tropsch Synthesis on
167. Steinbach, F.; Kiss, J.; Krall, R. Surf. Sci. 1985, 157, 401–412. Fe(100) Using Density Functional Theory. Ph.D. Thesis, Eindhoven University of
168. Denecke, R. Appl. Phys. Mater. Sci. Process. 2005, 80, 977–986. Technology, Eindhoven, The Netherlands, 2010.
169. Krebs, H. J.; Bonzel, H. P.; Gafner, G. Surf. Sci. 1979, 88, 269–283. 193. Cheng, J.; Hu, P.; Ellis, P.; French, S.; Kelly, G.; Lok, C. M. J. Phys. Chem. C
170. Dwyer, D. J.; Gland, J.; Albert, M.; Bernasek, S., Intermediates to the Dissociative 2010, 114, 1085–1093.
Chemisorption of Co and Ch3oh on Fe(100). Abstracts of Papers of the American 194. Cheng, J.; Gong, X. Q.; Hu, P.; Lok, C. M.; Ellis, P.; French, S. J. Catal. 2008,
Chemical Society 1987, 193, 30. 254, 285–295.
171. Moon, D. W.; Cameron, S.; Zaera, F.; Eberhardt, W.; Carr, R.; Bernasek, S. L.; 195. Biloen, P.; Helle, J. N.; Sachtler, W. M. H. J. Catal. 1979, 58, 95–107.
Gland, J. L.; Dwyer, D. J. Surf. Sci. 1987, 180, L123–L128. 196. Biloen, P.; Sachtler, W. M. H. Adv. Catal. 1981, 30, 165–216.
172. Dwyer, D. J.; Somorjai, G. A. J. Catal. 1978, 52, 291–301. 197. Ciobica, I. M.; Kramer, G. J.; Ge, Q.; Neurock, M.; van Santen, R. A. J. Catal.
173. Dwyer, D. J.; Hardenbergh, J. H. J. Catal. 1984, 87, 66–76. 2002, 212, 136–144.
174. Wedler, G.; Colb, K. G.; McElhiney, G.; Heinrich, W. Appl. Surf. Sci. 1978, 2, 30–42. 198. Zhuo, M. K.; Tan, K. F.; Borgna, A.; Saeys, M. J. Phys. Chem. C 2009, 113,
175. Wedler, G.; Colb, K. G.; Heinrich, W.; McElhiney, G. Appl. Surf. Sci. 1978, 2, 8357–8365.
85–101. 199. Gracia, J. M.; Prinsloo, F. F.; Niemantsverdriet, J. W. Catal. Lett. 2009, 133, 257–261.
176. Vink, T. J.; Gijzeman, O. L. J.; Geus, J. W. Surf. Sci. 1985, 150, 14–23. 200. Deng, L. J.; Huo, C. F.; Liu, X. W.; Zhao, X. H.; Li, Y. W.; Wang, J. G.; Jiao, H. J.
177. van Santen, R. A.; Neurock, M.; Shetty, S. G. Chem. Rev. 2010, 110, J. Phys. Chem. C 2010, 114, 21585–21592.
2005–2048. 201. Anderson, R. B. In Catalysis; Emmett, P. H., Ed.; Reinhold Publishing Company:
178. Hammer, B.; Norskov, J. K. In Advances in Catalysis, Impact of Surface Science New York, 1956; Vol. IV.
on Catalysis; Academic Press: San Diego, 2000; Vol. 45, pp 71–129. 202. Botes, F. G.; van Dyk, B.; McGregor, C. Ind. Eng. Chem. Res. 2009, 48,
179. Bromfield, T. C.; Ferre, D. C.; Niemantsverdriet, J. W. ChemPhysChem 2005, 6, 10439–10447.
254–260. 203. Botes, F. G. Catal. Rev.: Sci. Eng. 2008, 50, 471–491.
180. Curulla-Ferre, D.; Govender, A.; Bromfield, T. C.; Niemantsverdriet, J. W. J. Phys. 204. Botes, F. G. Ind. Eng. Chem. Res. 2009, 48, 1859–1865.
Chem. B 2006, 110, 13897–13904. 205. Yang, J.; Liu, Y.; Chang, J.; Wang, Y. N.; Bai, L.; Xu, Y. Y.; Xiang, H. W.; Li, Y. W.;
181. Sorescu, D. C. J. Phys. Chem. C 2008, 112, 10472–10489. Zhong, B. Ind. Eng. Chem. Res. 2003, 42, 5066–5090.
182. Sorescu, D. C.; Thompson, D. L.; Hurley, M. M.; Chabalowski, C. F. Phys. Rev. B 206. van Dijk, H. A. J. Ph.D. Thesis, Eindhoven University of Technology, 2001
2002, 66, 035416. 207. Botes, F. G.; Breman, B. B. Ind. Eng. Chem. Res. 2006, 45, 7415–7426.
183. Ojeda, M.; Nabar, R.; Nilekar, A. U.; Ishikawa, A.; Mavrikakis, M.; Iglesia, E. 208. Yates, I. C.; Satterfield, C. N. Energy Fuel 1991, 5, 168–173.
J. Catal. 2010, 272, 287–297. 209. Donnelly, T. J.; Yates, I. C.; Satterfield, C. N. Energy Fuel 1988, 2, 734–739.
184. Scheijen, F. J. E.; Ferre, D. C.; Niemantsverdriet, J. W. J. Phys. Chem. C 2009, 210. Lozano-Blanco, G.; Thybaut, J. W.; Surla, K.; Galtier, P.; Marin, G. B. Ind. Eng.
113, 11041–11049. Chem. Res. 2008, 47, 5879–5891.
185. Inderwildi, O. R.; Jenkins, S. J.; King, D. A. J. Phys. Chem. C 2008, 112, 211. Yang, Q. Y.; Maynard, K. J.; Johnson, A. D.; Ceyer, S. T. J. Chem. Phys. 1995,
1305–1307. 102, 7734–7749.
186. Shetty, S.; van Santen, R. A. Phys. Chem. Chem. Phys. 2010, 12, 6330–6332. 212. Govender, N. S.; Botes, F. G.; de Croon, M.; Schouten, J. C. J. Catal. 2008, 260,
187. Jones, G.; Bligaard, T.; Abild-Pedersen, F.; Norskov, J. K. J. Phys. Condens. 254–261.
Matter 2008, 20, 064239. 213. Weckhuysen, B. M. Angew. Chem. Int. Ed. 2009, 48, 4910–4943.

Comprehensive Inorganic Chemistry II: From Elements to Applications, (2013), vol. 7, pp. 525-557
View publication stats

Das könnte Ihnen auch gefallen