You are on page 1of 64

Master’s Programme in Advanced Energy Solutions

Power-to-X: Modelling of Fischer-Tropsch


synthesis in Aspen Plus

Hassan Sayed Ahmed

Master’s Thesis
2021
Copyright ©2021 Hassan Sayed Ahmed
Author Hassan Sayed Ahmed
Title of thesis Power-to-X: Modelling of Fischer-Tropsch synthesis in Aspen Plus
Programme Advanced Energy Solutions
Major Sustainable Energy Systems and Markets
Thesis supervisor Associate Prof. Annukka Santasalo-Aarnio
Thesis advisor(s) Judit Nyári, MSc
Date 24.12.2021 Number of pages 64 / 61 Language English

Abstract
Global warming is the most challenging problem we are facing nowadays. Simulta-
neously, the energy sector is responsible for about three-quarters of the global
greenhouse gas emissions. Electrification has been the main solution for defossiliz-
ing the energy sector in the last years, to tackle the emissions problem. In spite of
that, many parts of the energy sector are hard to electrify, such as industry, aviation,
marine, and heavy transport vehicles. Power-to-X (PtX) can contribute to tackling
this problem. When using renewable electricity to produce hydrogen and capture
carbon dioxide, different near carbon-neutral fuels can be produced via different
PtX technologies, such as methanol, kerosene, diesel, gasoline, and natural gas. In
this thesis, more focus will be given to producing diesel via Fischer-Tropsch (FT)
synthesis.
In this thesis, first, a summary of some PtX technologies is given, where their
contribution in defossilizing the energy sector is investigated. Then, a techno-eco-
nomic analysis of carbon-neutral FT synthesis is conducted, where Aspen Plus is
used to model the process with having hydrogen and carbon dioxide as feedstock.
Moreover, Carbon-neutral FT synthesis readiness for large-scale production is in-
vestigated, and the overall efficiency of carbon-neutral FT synthesis and the Lev-
elized cost of FT fuels (LCOF) are calculated.
It was found that PtX technologies can produce fuels that directly replace fossil
ones with using the same infrastructure or with little modifications. Moreover, it
can add more flexibility and reliability to the electricity sector, which enables more
dependence on solar and wind energies. Besides, it was found that carbon-neutral
FT synthesis is not yet ready for large scale production, mainly because of reverse
water gas shift reaction which has TRL 6, and it was found that the overall efficiency
of carbon-neutral FT synthesis is about 39%, so more work needs to be done to in-
crease the efficiency of the process and its readiness for large scale production. On
the other hand, FT fuels were found to be from 2.5 to 5 times more expensive than
their equivalent fossil fuels. Hydrogen price was found to be the main contributor
to this high cost, as it accounted for about 80% of the LCOF. If green hydrogen
prices reached 0.5 €/kg, FT hydrocarbons will start to be economically competitive
with their fossil equivalent hydrocarbons. Therefore, more efforts have to be given
to reduce the cost of green hydrogen production to make FT fuels as economically
competitive as possible with their fossil counterparts.
Keywords Power-to-X, Fischer-Tropsch, CO2 hydrogenation, FT diesel, Techno-
economic, Aspen Plus, Modelling
Preface
I would like to thank Annukka Santasalo-Aarnio and Judit Nyári for their assistance and
guidance throughout the thesis and my master’s studies.

Also, I would like to give special thanks to my family and friends for their support and giving
me motivation during good and bad times.

Otaniemi, 24 December 2021


Hassan Sayed Ahmed

4
Contents
Preface ......................................................................................................... 4
List of figures ............................................................................................... 6
List of tables................................................................................................. 7
Symbols and abbreviations .......................................................................... 8
Abbreviations ........................................................................................... 8
Symbols.................................................................................................... 9
1 Introduction........................................................................................ 10
2 Introduction to Power-to-X technologies.............................................12
2.1 Power-to-heat ............................................................................... 13
2.2 Power-to-gas................................................................................. 15
2.2.1 Power-to-hydrogen ................................................................ 16
2.2.2 Power-to-methane.................................................................. 16
2.2.3 Power-to-syngas......................................................................17
2.3 Power-to-liquid............................................................................. 19
3 Carbon-neutral Fischer-Tropsch process ............................................ 23
3.1 Potential global market of Fischer-Tropsch products ................... 26
3.2 Environmental impact .................................................................. 29
4 Techno-economic analysis of Fischer-Tropsch synthesis .................... 32
4.1 Methodology ................................................................................. 32
4.2 Technical Modelling of Fischer-Tropsch synthesis........................ 33
4.2.1 Syngas production.................................................................. 33
4.2.2 FT synthesis ........................................................................... 34
4.2.3 FT-products upgrading........................................................... 34
4.3 Economic analysis of Fischer-Tropsch synthesis ..........................40
5 Results ................................................................................................ 44
5.1 Technical performance ................................................................. 44
5.2 Economic Feasibility..................................................................... 47
6 Conclusion ...........................................................................................51
References.................................................................................................. 53
Compounds and elements used in the Aspen Plus model ................... 62
Reactions rates parameters for LHHW kinetic models ....................... 63

5
List of figures
Figure 1: Power-to-X concept [7] ....................................................................................10
Figure 2: Power-to-X role in sector coupling [12] ............................................................12
Figure 3: Power-to-heat Scheme [18] ..............................................................................14
Figure 4: Power-to-gas overview [24]..............................................................................15
Figure 5: Products produced from Syngas and their usage [26] .......................................18
Figure 6: Syngas Production methods [26] ......................................................................19
Figure 7: Power-to-liquid overview [41] ..........................................................................20
Figure 8: Power-to-Methanol economically competitiveness [41] ...................................21
Figure 9: Variation of methanol cost with electricity prices [42] ......................................21
Figure 10: Relation between growth probability (α) and FT-product chain length (n) when
ASF distribution is assumed (dotted red line) and when experimental data is measured
(black line) [54] ...............................................................................................................25
Figure 11: Relation between carbon atom selectivity in FT-product spectrum (y-axis) ,
growth probability (x-axis), and type of FT-synthesis (highlighted in red and blue) [47] ...25
Figure 12: Carbon-neutral FT synthesis............................................................................26
Figure 13: Oil products demand in EU-28 from 2009 to 2020 [61] ...................................27
Figure 14: Relation between kerosene production and consumption in EU-28 from 2000
to 2019 [61] ....................................................................................................................28
Figure 15: Relation between diesel/gasoil production and consumption in EU-28 from
2000 to 2019 [61] ...........................................................................................................28
Figure 16: Main Environmental Footprints[63] ................................................................29
Figure 17: Land use comparison between PtL via FT and biofuels [43] .............................30
Figure 18: Water demand comparison between PtL via FT and biofuels [43] ...................30
Figure 19: Aspen Plus model flowsheet ...........................................................................37
Figure 20: Relation between log of molar fraction of FT-products (LogMn) and FT-product
chain length (n) when ASF distribution is assumed (dotted red line) and when
experimental data is measured (black line)[54] ...............................................................39
Figure 21: Relation between carbon selectivity of FT products in Aspen Plus model and in
normalized ideal ASF distribution....................................................................................44
Figure 22: Sensitivity analysis of FT synthesis pressure, T=220 0C, H2/CO=2 .....................46
Figure 23: Sensitivity analysis of FT synthesis temperature, P=25bar, H 2/CO=2................46
Figure 24: Sensitivity analysis of FT synthesis H 2/CO ratio, T=220 0C, P=25bar .................47
Figure 25: Contribution percentages of cost sources .......................................................48
Figure 26: Relation between FT-products cost divided by fossil-based cost and change in
H2 prices .........................................................................................................................48
Figure 27: Relation between FT-products cost divided by fossil-based cost and change in
CO2 prices .......................................................................................................................49
Figure 28: Relation between FT-products cost divided by fossil-based cost and change in
Capex..............................................................................................................................49

6
List of tables
Table 1: Comparison between LTFT and HTFT .................................................................23
Table 2: Equations used for calculating parameters used in hydroisomerization and
hydrocracking LHHW kinetic rates [50]............................................................................36
Table 3: Parameters of the main blocks in the model ......................................................38
Table 4: Parameters used in efficiency calculation...........................................................40
Table 5: OPEX calculations parameters............................................................................42
Table 6: Parameters showing the performance of the Aspen Plus model.........................45
Table 7: LCOF of FT-products...........................................................................................47
Table A 1: Components used in the Aspen Plus model
Table B 1: Parameters used for calculating LHHW kinetic rates in FT synthesis
Table B 2: Values of LHHW kinetic rates used for alkanes in FT synthesis
Table B 3: Values of LHHW kinetic rates used for alkenes in FT synthesis

7
Symbols and abbreviations
Abbreviations

APEA Aspen process economic analyzer


ASF Anderson-Schulz-Flory
Btl Biomass-to-liquid
Capex Capital expenses
CCU Carbon capture and utilization
CEPCI Chemical engineering plant cost index
COP Coefficient of operation
COP21 Conference of the Parties 21
COP26 Conference of the Parties 26
CPI Consumer price index
DE Design and engineering
FT Fischer-Tropsch
GHG Greenhouse gas
HC Hydrocarbons
HTFT High temperature Fischer-Tropsch
IEA International Energy Agency
IPCC Intergovernmental Panel on Climate Change
IRENA International renewable energy agency
ISBL Inside battery limits
LCOF Levelized cost of fuel
LHHW Langmuir–Hinshelwood–Hougen–Watson
LHV Low heating value
LTFT Low temperature Fischer-Tropsch
OPEX Operating expenses
OS Outside battery limits
PEM Proton exchange membrane
PSA Pressure swing adsorption
PtG Power-to-gas
PtH Power-to-heat
PtL Power-to-liquid
PtX Power-to-X
RWGS Reverse water gas shift reaction
SNG Synthetic natural gas
SOEC Solid oxide electrolysis cells
SRK Soave-Redlich-Kwong
TRL Technology readiness level
WGS Water gas shift reaction
WWSHP Wastewater source heat pump

8
Symbols

A Pre-exponential factor
C Cost of component
E Activation energy
FTP-C5+ Fischer-Tropsch alkaline products with number of car-
bon atoms 5 or more
f Correction factor
fug(n) Fugacity
GJ/ha.yr Giga joule of fuel is produced per one hectare of land in
one year
gco2eq/MJfuel Grams of equivalent CO2 emissions per 1 Mega joule of
fuel
H2R Recycled hydrogen to the beginning of the process
h Annual running hours
Keq Equilibrium constant
KL Langmuir constant
k Rate constant
k 0 Pre-exponential factor
Mn Molar selectivity of hydrocarbons
Mtonnes/year Million tonnes per year
ṁ Mass flow rate
m /GJ
3 Cubic meter of water is used for producing 1 giga joule of
fuel
PCO Partial pressure of carbon monoxide
PH2 Partial pressure of hydrogen
R Reaction rate
r Discount rate
S Equipment size
α Growth probability

9
1 Introduction
According to the Intergovernmental Panel on Climate Change (IPCC), the
global surface temperature in the last decade (2011-2020) was already 1.09
0C higher than the pre-industrial level (1850-1900) [1]. At COP 21 in Paris,

the parties involved agreed to limit the global warming this century to less
than 2 0C, with making efforts to keep it at 1.5 0C more than the pre-industrial
level [2]. While at COP 26 in Glasgow, 6 years after COP 21, the parties in-
volved realized that if their climate plans are fully and successfully imple-
mented, this will cause an increase in global warming from 1.8 0C to 2.4 0C
by the end of this century, which is more than the limits agreed on in COP 21
[3]. At the same time, energy sector is responsible for about three-quarters
of the global greenhouse gas (GHG) emissions [4]. So, if we really want to
adhere to the Paris Agreement, more efforts and more radical procedures are
required to limit the GHG emissions from the energy sector.

Solar and wind energies are paving the road for a carbon-neutral energy
system. While they are clean and cheap sources of energy, they are also in-
termittent. So, the quality, reliability, and stability of energy systems that
have high share of intermittent renewable sources are questioned. Energy
storage and demand response plays an important role in solving these issues
by storing the surplus energy produced when solar and wind energies are
available and decreasing energy curtailment [5]. One of the most prominent
technologies for energy storage and which can provide demand response to
the energy sector is Power-to-X (PtX) [6]. Figure 1 shows the main concept
of PtX where renewable electricity can be converted to different energy
forms, such as thermal energy via Power-to-Heat (PtH) and chemical energy
via Power-to-Gas (PtG) and Power-to-Liquid (PtL).

Figure 1: Power-to-X concept [7]


10
Different liquid fuels and chemicals can be produced from PtL, such as
ammonia, methanol, oxymethylene ether, and Fischer-Tropsch (FT) prod-
ucts [7]. FT products include diesel, gasoline, kerosene, naphtha, wax, and
olefin [8], [9]. Liquid fuels are widely used in transport sector, which is re-
sponsible for more than 16% of the global GHG emissions and globally uses
the largest share of oil when compared to other sectors [4], [10]. Carbon-
neutral liquid fuels can contribute in defossilizing the transport sector. In this
thesis, focus will be given to PtL via Fischer-Tropsch (FT) route and the main
research questions that this thesis is answering are:

 What is Power-to-X? And how can it contribute in defossilizing the


energy and transport sector?
 What are the main pros and cons of carbon neutral Fischer-Tropsch
synthesis?
 What is the relation between Fischer-Tropsch synthesis operating
conditions and the produced liquid hydrocarbons?
 What are the main parameters affecting the cost of Fischer-Tropsch
based hydrocarbons? And what is the price ratio between Fischer-
Tropsch based and fossil-based diesel?

11
2 Introduction to Power-to-X technologies
PtX is a broad term that includes different processes to convert renewable
electricity to various energy carriers, such as heat, hydrogen, methane, chem-
icals, and liquid fuels [11]–[13]. These products can directly replace fossil
fuels in different sectors, such as industry, heavy transport vehicles, aviation,
and marine, where electrification is challenging. Besides, many of these prod-
ucts can be cheaply stored, when compared to electrical energy storage, easily
transported to other places, and used for re-producing electricity when there
is a demand [6].

Figure 2: Power-to-X role in sector coupling [12]

Figure 2 shows how PtX promotes coupling between different sectors. In


addition to defossilizing different hard-to-defossilize sectors, PtX provides
more flexibility to the electricity sector [6], [11], [13], which improves the in-
tegration of the intermittent renewable energies, wind and solar energies in
particular, into the electricity grid [11] and increases its reliability and quality
[14].

PtX enables us to take immediate actions against global warming by re-


placing fossil fuels with carbon-neutral synthetic fuels using the currently ex-
isting infrastructure [15]. Because of that, it can act as short and medium-
term solutions for defossilizing the energy and transport sector. On the other
hand, nowadays, the cost of most Power-to-X processes is still higher than its

12
alternatives which depend on fossil fuels [16]. However, Schnuelle et al. sug-
gest that if analysts considered the costs associated with climate change and
the penalties for missing the climate targets, Power-to-X technologies would
be more economically attractive [17].

In this chapter, a summary of some PtX processes will be given, such as


PtH, PtG, and PtL, and how each technology can contribute to the defossili-
zation of the energy and transport sector.

2.1 Power-to-heat

International Renewable Energy Agency (IRENA) indicates in its analysis of


the heating sector that heating and cooling applications in industrial and res-
idential sectors are responsible for about 35% of the global energy consump-
tion. While in Europe, 51% of the annual energy demand is for heating and
cooling ,and 75% of this European heating and cooling demand is met by fos-
sil fuel [18].

For defossilizing this sector, energy can be produced directly from renew-
able sources, such as solar and geothermal energies, or by using renewable
electricity to produce heat energy. The latter is also known as Power-to-heat
(PtH). As shown in Figure 3, PtH uses renewable electricity, such as solar and
wind energies, to produce thermal energy by heat pumps and electric boilers
[18]. This energy can be used directly or stored for later when there is a de-
mand. PtH can be used in both centralized and decentralized systems. So, it
can help in defossilizing district heating and cooling networks and in decreas-
ing renewable energy curtailment from small, medium, and large scale inter-
mittent renewable energy sources.

13
Figure 3: Power-to-heat Scheme [18]

PtH is already an economically attractive option for replacing fossil fuels.


The levelized cost of heat in the European Union when using air-to-air heat
pumps with an average coefficient of operation (COP) of 3.83 is lower than
when using natural gas or oil [18]. On the other hand, the levelized cost of
heat when using electric heaters is higher than its fossil fuel alternatives [18].
This difference in cost between using heat pumps or electric heaters is caused
by the principle of operation of each one of them. Heat pumps use the energy
in the surrounding medium in addition to electricity to produce the thermal
energy needed, so COP is always greater than one [19], while electric heaters
produce heat with efficiency lower than one [20]. In addition to that, the
prices of PtH technologies can be reduced if coupled with thermal energy
storage [18]; since heat can be produced at off-peak times. Besides, increas-
ing the flexibility of the electricity sector and reducing the curtailment of in-
termittent renewable energies is achieved by adding more demand response,
since heating demand is not as instantaneous as electricity demand, and in-
stalling thermal energy storage systems, which is much cheaper than in-
stalling electric energy storage systems for the same purpose [21], [22].

In spite of the maturity of PtH residential and low-temperature industrial


applications and their competitive prices, high-temperature industrial appli-
cations still need more research and development, since high-temperature
heat pumps have low technology readiness levels and electric boilers are not
economically competitive when used at high temperatures [23]. For fully

14
defossilizing the heating sector carbon-neutral synthetic fuels are needed to
replace fossil fuels, which is used as feedstock for heating purposes, in short
and medium terms.

2.2 Power-to-gas

Hydrogen, methane, and syngas are the three main products of PtG technol-
ogies [24]. In carbon-neutral PtG processes, renewable electricity is used to
produce hydrogen and capture carbon dioxide from industry or directly from
air [24], [25]. Renewable electricity is also needed in reducing captured car-
bon dioxide into carbon monoxide which can be used with hydrogen to pro-
duce syngas either through co-electrolysis or reverse water gas shift reaction
(RWGS) [26]. Hydrogen and captured carbon dioxide are required to pro-
duce methane [27]. Power-to-hydrogen, Power-to-methane, and Power-to-
syngas will be discussed in more details in this chapter. Figure 4 shows an
overview of the PtG processes, where produced gases can be stored, con-
verted to other chemicals and fuels, or directly used in different applications,
such as industry, transportation, electricity production, and heat production.

Figure 4: Power-to-gas overview [24]

PtG can support the flexibility of the grid and the integration of intermit-
tent renewable energies by utilizing the existing fossil fuel storage infrastruc-
ture, especially for long term and seasonal storage in countries where there

15
is a huge difference in the availability of the intermittent renewable energies
among different seasons [28].

2.2.1 Power-to-hydrogen

Currently, hydrogen is mainly used in producing ammonia, methanol, hydro-


gen peroxide, and in oil refineries [29]. Hydrogen production nowadays is
highly dependent on fossil fuel, especially natural gas and coal, where only
2% of hydrogen is produced by water electrolysis [28]. Hydrogen production
is responsible for GHG emissions equal to 830 MtCO2/year which corre-
sponds to about 2.5% of the global GHG emissions [28], [30]. In the future,
higher demand for hydrogen is expected, to defossilize the energy sector, as
it can be used in different parts of the energy sector, such as electricity, in-
dustry, heating, and transportation [31].

For avoiding the emissions associated with hydrogen production, green


hydrogen can be produced via biomass gasification, artificial photosynthesis,
or water electrolysis with renewable electricity sources, where different types
of electrolyser technologies can be used, such as alkaline electrolysis, proton
exchange membrane electrolysis (PEM), and solid oxide electrolysis cells
(SOEC) [28], [32].

Green hydrogen can be used as a feedstock for producing synthetic gas and
synthetic liquid fuels [28], which can replace fossil fuels in different sectors.
Besides, it can act as an energy storage and demand response source for add-
ing more flexibility and improving the integration of intermittent renewable
electricity to the electricity grid. Demand response can be achieved by the
electrolysis process where alkaline electrolysis operates between 10-110 % of
its nominal power, PEM operates between 0-160 % of its nominal power, and
SOEC operates between 20-100 % of its nominal power [28]. While as an
energy storage, hydrogen can be stored efficiently for long term to overcome
the seasonal variability of intermittent renewable energy [33], [34] and
transported through natural gas pipelines with some limitations [28],
Melaina et al. estimated that 5%-15% of hydrogen by volume can be blended
into USA natural gas pipelines network without affecting the public safety or
the durability of the network [35]. Moreover, hydrogen can be used in trans-
portation via hydrogen fuel cells and hydrogen-fueled internal combustion
engine [31], [36].

2.2.2 Power-to-methane

In Power-to-methane, hydrogen and carbon dioxide are used to produce me-


thane via catalytic methanation or biological methanation [37]. Methane can
later be upgraded and converted to synthetic natural gas (SNG) [38]. SNG

16
can fully replace fossil-based natural gas in different sectors, it can be used
in transportation, power generation, heat generation, or as a raw material in
industry [27]. The main advantage of converting hydrogen into SNG is that
SNG has a higher chemical energy density [27] and can fully utilize the exist-
ing natural gas network infrastructure.

On the other hand, Götz et al. [37] showed that SNG is not yet economi-
cally competitive with fossil-based natural gas. Hydrogen production and
storage are the two dominant cost factors for Power-to-methane process. Hy-
drogen storage is needed to utilize more intermittent renewable energies
since electrolysis is more dynamic than the methanation process. It is worth
mentioning that Götz et al. did not include carbon certificates revenue, the
avoided costs of expanding the electricity grid infrastructure, and revenue
from the electricity balancing market in their calculations; when added, these
opportunities can increase the economic feasibility of SNG.

2.2.3 Power-to-syngas

Synthetic gas (syngas) is a mixture of hydrogen and carbon monoxide [39].


It is used to produce a large spectrum of chemicals and fuels, which are used
in transportation, chemical industry, and energy storage, through changing
the ratio between hydrogen and carbon monoxide in different chemical reac-
tions [26], [39]. Figure 5 shows the products produced from syngas, their
usage in different industrial sectors, the processes needed to produce each
product, and the ratio required between hydrogen and carbon monoxide in
each process.

17
Figure 5: Products produced from Syngas and their usage [26]

Traditionally, syngas is produced by partial oxidation and steam reform-


ing of hydrocarbons, or coal gasification where natural gas and coal are the
two main sources for producing syngas [26], [39]. To ensure the carbon neu-
trality of the end products, syngas should be produced in a sustainable green
method. This can be done by biomass gasification, or producing hydrogen
separately via water electrolysis and producing carbon monoxide by carbon
dioxide reduction or RWGS reaction then mix them with the desired ratio for
the upcoming process, or producing both of them at the same time by co-
electrolysis of water and carbon dioxide at high temperature [26], [40]. The
latter process is still under development due to the complexity of the system.
Figure 6 describes different methods for producing syngas.

18
Figure 6: Syngas Production methods [26]

2.3 Power-to-liquid

Liquid fuels have higher volumetric energy density than gaseous fuels, which
makes them more favorable in aviation, shipping, and heavy vehicle trans-
portation [41]–[44]. Besides, liquid fuels are easier to store and transport,
since no extensive safety precautions nor high pressure are needed [42].
Their usage in hard-to-defossilize sectors increased the interest in developing
carbon neutral liquid fuels to tackle the climate crisis [41].

Power-to-liquid (PtL) is one of the technologies directed to develop carbon


neutral liquid fuels. In carbon neutral PtL process, green hydrogen and car-
bon dioxide, captured directly from air or from large point sources, act as
feedstock for well-established chemical processes as shown in Figure 7 [41],
[42]. Syngas can act as feedstock in methanol, Fischer-Tropsch fuels, and di-
methyl ether production [41]. Currently, methanol production via PtL is
mainly done by direct hydrogenation of carbon dioxide, and dimethyl ether
is produced by methanol dehydration, while Fischer-Tropsch depends on
syngas as feedstock [41]. Moreover, more research is being done to utilize
carbon dioxide directly in the production of liquid fuels via Fischer-Tropsch
and compare between these two routes [41].

19
Figure 7: Power-to-liquid overview [41]

The production of Fischer-Tropsch fuels and dimethyl ether via PtL is still
economically uncompetitive with fossil-fuel based production, while metha-
nol production via PtL started to be competitive [41]. In Figure 8, we can see
that methanol prices are fluctuating. When methanol prices are at their peak,
methanol production via PtL becomes economically competitive with fossil-
based production [41]. It is worth mentioning that hydrogen prices, or elec-
tricity prices needed for producing hydrogen, are the dominant cost for pro-
ducing liquid fuels [41], [45]. Figure 9 shows the effect of electricity prices on
methanol price via PtL. As can be seen in Figure 9, increasing the electricity
price from 0.01 to 0.07 €/kWh can increase methanol price from about 0.06
to about 0.15 €/kWh. Therefore, decreasing the electricity prices increases
the economic competitiveness of liquid fuels via PtL with their fossil coun-
terparts.

20
Figure 8: Power-to-Methanol economically competitiveness [41]

Figure 9: Variation of methanol cost with electricity prices [42]

21
Flexible operation of PtL processes play an important role in integrating
intermittent renewable energies to the electricity grid, to avoid the curtail-
ment of available clean energy. Methanol synthesis can be operated in a flex-
ible manner while Fischer-Tropsch synthesis is sensitive to any changes in
H2/CO ratio [41]. For having more flexibility in Fischer-Tropsch synthesis,
intermittent hydrogen storage will be needed, if hydrogen production is done
on-site, which increases the cost of the Fischer-Tropsch fuels. Besides, Di-
eterich et al. suggests that more research on dimethyl ether synthesis is
needed to examine the flexibility of its operation whether it is produced in
direct one stage process or indirect two stage process [41]. In this thesis, more
focus will be given to production of diesel by PtL via FT route.

22
3 Carbon-neutral Fischer-Tropsch process
Fischer-Tropsch (FT) synthesis is a heterogeneous catalytic and polymeriza-
tion reaction for producing liquid hydrocarbon fuels and chemicals from syn-
gas [46], [47]. The overall process is named after the feedstock used for pro-
ducing the syngas, such as coal-to-liquid, gas-to-liquid, and biomass-to-liq-
uid [47]. The first attempt for coal liquefaction was done by Friedrich Bergius
in 1913 [48], [49]. He succeeded in producing petroleum-like-liquids by hy-
drogenation of coal at 200 atm and 673 K [48]. In 1926, Franz Fischer and
Hans Tropsch introduced their process to convert coal to liquid hydrocar-
bons at 1 to 10 atm and 453 to 473 K. In this process coal gasification took
place for producing syngas which reacted later, in the presence of cobalt-
based catalyst, to produce the desired liquid hydrocarbons [48]. FT synthesis
can be optimized to produce different range of hydrocarbons, such as syn-
thetic naphtha, diesel fuel, kerosene, lubricating oils, synthetic waxes, and
olefins [8], [9].

There are mainly two modes of operation in FT, low temperature Fischer-
Tropsch (LTFT) synthesis and high temperature Fischer-Tropsch (HTFT)
synthesis. In LTFT the operating temperature ranges from 180 to 250 0C
while in HTFT it ranges from 300 to 350 0C [50]. The higher temperature
increases the selectivity of shorter hydrocarbons and 1-alkenes [50]. Conse-
quently, LTFT is used to produce diesel, naphtha, and wax while HTFT is
used to produce gasoline and chemicals [47]. Product upgrading is required
in both modes for maximizing the desired output, such as hydrocracking in
LTFT and isomerization, alkylation, and oligomerization in HTFT [51]. Table
1 summarizes the main differences between LTFT and HTFT.

Table 1: Comparison between LTFT and HTFT


LTFT HTFT

Temperature 180 – 250 0C 300 – 350 0C

Catalyst Iron or Cobalt based Iron based


Longer Hydrocarbon Shorter Hydrocarbon chain +
Favors
chain Olefins
Isomerization + Alkylation +
Product Upgrading Hydrocracking
Oligomerization

Since FT synthesis is a catalytic reaction [46], the choice of suitable cata-


lyst is critical to the process. Cobalt, iron, ruthenium, and nickel based cata-
lysts can activate the reaction [52]. In practical applications, only cobalt and

23
iron-based catalysts are used since ruthenium is very expensive and rarely
available while nickel has high selectivity towards methane and lighter prod-
ucts [52]. Cobalt-based catalysts are active, stable, and have long lifetime, but
they have high methane selectivity in high temperatures, so they are only
used in LTFT. Besides, they have low activity to water gas shift reaction, so
the H2/CO ratio needs to be high to ensure high CO conversion [52], [53].
On the other hand, iron-based catalysts are cheaper than cobalt, active in
broader temperature range, and have higher water gas shift reaction activity,
so H2/CO ratio used is lower compared to cobalt-based catalysts. Nonethe-
less, they have lower lifetime and less CO conversion than cobalt-based cata-
lysts [52], [53].

Since FT synthesis is a polymerization reaction [46], increasing the pro-


duced hydrocarbon chain length is done via stepwise polymerization of CHx
polymer (x=0-3), the probability of increasing the chain length of the pro-
duced hydrocarbons or terminating the reaction is determined by a factor
called growth probability (α-value) [54]. The growth probability is affected
by the operating conditions, such as temperature, pressure, and H 2/CO rati0.
However, in ideal distribution, the produced hydrocarbon spectrum follows
Anderson-Schulz-Flory (ASF) distribution and α-value is constant and inde-
pendent on the number of carbon atoms, while in reality it differs and can
be experimentally determined for each hydrocarbon in the product spectrum
as shown in Figure 10 [54]. Van de Loosdrecht et al. [47] explained the rela-
tion between growth probability and selectivity of hydrocarbons in FT syn-
thesis in different FT synthesis types. As can be seen in Figure 11, in LTFT,
when growth probability is between about 0.85 and 0.95, high selectivity of
diesel and waxes can be achieved. While in HTFT, when growth probability
is between about 0.7 and 0.75, high selectivity of gasoline can be achieved.

24
Figure 10: Relation between growth probability (α) and FT-product chain
length (n) when ASF distribution is assumed (dotted red line) and when ex-
perimental data is measured (black line) [54]

Figure 11: Relation between carbon atom selectivity in FT-product spec-


trum (y-axis) , growth probability (x-axis), and type of FT-synthesis (high-
lighted in red and blue) [47]

Near carbon-neutral FT synthesis can be achieved when having captured


carbon dioxide and green hydrogen as feedstock. Producing near carbon-
neutral FT products can be done through two approaches. The first one is
separately producing syngas from green hydrogen and carbon dioxide using
RWGS reaction then FT synthesis takes place [55], [56]. This approach is
shown in Figure 12. It is also analyzed and modelled later in this thesis using

25
Aspen Plus. It is worth mentioning that the technology readiness level (TRL)
of this two-step approach is 6, because RWGS reaction has TRL 6, while FT
synthesis and product upgrading have TRL 9 [43]. The second approach is
direct hydrogenation of carbon dioxide to produce C5+ liquid fuels, where
RWGS reaction and FT synthesis are activated by the same catalyst in the
same reactor simultaneously [57], [58]. The direct hydrogenation of CO2 ap-
proach is less mature than the two-step approach. Many research attempts
are being currently done to develop catalysts that can improve the selectivity
of C5+ liquid fuels in the case of direct hydrogenation of carbon dioxide. Choi
et al. [58] said they have developed a catalyst which gives C5+ fuels selectivity
in the direct hydrogenation method similar to conventional FT synthesis.

Figure 12: Carbon-neutral FT synthesis

3.1 Potential global market of Fischer-Tropsch products

In 2020, the refined petroleum products global market was 1788 billion USD
[59]. This value was lower than usual due to the Covid-19 pandemic, and it is
expected that the demand will return to the pre-pandemic levels by the end
of 2022 [60]. The FT products can directly replace their equivalent of petro-
leum products which gives FT synthesis huge market potentials.

In 2020 in EU-28, the total oil products demand was 567 Mtonnes/year,
this was 12 % lower than 2019 level due to the Covid-19 pandemic [61]. Diesel
and gasoil demand accounted for about half of the total oil products demand
[61]. Figure 13 shows the historical oil products demand in EU-28 from 2009
to 2020.

26
Figure 13: Oil products demand in EU-28 from 2009 to 2020 [61]

Energy security has always been an important question when assessing


energy sources [62]. EU-28 has excess production of gasoline, but it suffers
deficiency in kerosene and diesel/gasoil production where Middle East and
Russia are the main exporters of these fuels to EU-28 respectively [61]. FT
synthesis is one of the technologies that can tackle this energy security prob-
lem since it can be optimized to produce the needed kerosene and die-
sel/gasoil [8]. As shown in Figure 14 and Figure 15, a huge percentage of ker-
osene consumption in EU-28 is imported, while the percentage of imported
diesel/gasoil is not that high. So, the need for producing kerosene via FT syn-
thesis, from energy security perspective, is higher than the need for produc-
ing diesel/gasoil.

27
Figure 14: Relation between kerosene production and consumption in EU-
28 from 2000 to 2019 [61]

Figure 15: Relation between diesel/gasoil production and consumption in


EU-28 from 2000 to 2019 [61]

28
3.2 Environmental impact

When assessing the environmental impact of FT synthesis, several environ-


mental footprints should be taken into consideration, such as GHG emis-
sions, land use, water demand, and particulate matter emissions [43]. The
main environmental footprints considered when evaluating different tech-
nologies are shown in Figure 16.

Figure 16: Main Environmental Footprints[63]

According to Schmidt et al. [43] GHG emissions of near carbon-neutral


FT synthesis is about 1 gco2eq/MJfuel, when emissions related to construction
of power plants and production facility today is not considered, and from 11
to 28 gco2eq/MJfuel, when construction emissions today is considered, while
GHG emissions of crude oil is 87.5 gco2eq/MJfuel. On the other hand, Biomass-
to-liquid (BtL) GHG emissions can be as low as -2 gco2eq/MJfuel when land use
change is considered.

Although BtL can have lower GHG emissions than near carbon-neutral FT
synthesis, it has higher land use and water demand [43]. Schmidt et al. [43]
compared land use and water demand of jet fuel produced by different pro-
duction pathways. They found that from 470 to 1070 GJ/ha.yr can be pro-
duced by PtL via FT synthesis while BtL can produce from 47 to 171 GJ/ha.yr
and biofuel can produce from 15 to 402 GJ/ha.yr. Figure 17 shows a compar-
ison of land use between PtL via FT synthesis and biofuels. Moreover, the
water demand of PtL via FT synthesis is 0.04 m3/GJ, while it is 112 m3/GJ
29
from BtL and from 14 to 574 m3/GJ from biofuel. Figure 18 shows a compar-
ison between the water demand of PtL via FT synthesis and different biofuels.

Figure 17: Land use comparison between PtL via FT and biofuels [43]

Figure 18: Water demand comparison between PtL via FT and biofuels [43]

FT products are free of sulfur and nitrogen and have low aromatics [43],
[64]. These characteristics reduce the particulate matter emission produced
during combustion, when compared with conventional fuels, and improve air
quality [43]. Conventional fossil fuel combustion is responsible for 85% of
the particulate matter emissions worldwide, and almost all sulfur dioxide and
30
nitrogen oxide emissions [65]. Air pollution causes severe health problems
to humans and kills about 7 million people every year [66].

31
4 Techno-economic analysis of Fischer-Tropsch
synthesis
In this thesis, a near carbon neutral FT synthesis process is modelled using
Aspen Plus. Hydrogen and carbon dioxide are the feedstock of this process.
Then, a techno-economic analysis is conducted using both Aspen Plus and
Microsoft Excel. Hydrogen electrolysing and carbon dioxide capturing are
not included in this model or in the economic analysis, but they’re included
when calculating the efficiency of the overall process. The model consists of
three main parts: syngas production, FT synthesis, and FT-products upgrad-
ing.

Different techno-economic analyses were done earlier on PtL via LTFT


synthesis route by different papers, such as Selvatico et al. [50], Zang et al.
[50], and Vázquez et al. [67]. Syngas is assumed to be the feedstock in the
work done by Selvatico et al. [50], while H2 and CO2 were the feedstock in the
work done by Zang et al. [45]. In SOLETAIR pilot described by Vázquez et al.
[67], renewable electricity from PV powerplant is used to produce hydrogen
via water electrolysis and CO2 via direct air capture. Then, the two gases un-
dergo RWGS reaction to produce syngas which passes by FT synthesis to pro-
duce liquid fuels and paraffin wax. The energy efficiency of the overall system
reported by Selvatico et al. [50] ranged from 39% to 40%, while Zang et al.
[50] reported overall energy efficiency equals to 57.5 %, and Vázquez et al.
[67] reported overall energy efficiency equals to 47%.

4.1 Methodology

In this thesis, the flowsheet simulation model is based on the work done by
Selvatico et al. [50], Dimitriou et al. [56], and Lox and Froment [68]. In the
model developed in this thesis, optimizing the production of diesel via FT is
targeted. So, LTFT synthesis is modelled, since it produces longer hydrocar-
bon chain length than HTFT [47]. For examining the effect of H 2/CO ratio on
the economic feasibility of the process, precipitated promoted iron catalyst
was assumed, since it stimulates both FT reactions and water gas shift (WGS)
reaction in the FT reactor which can make the optimal H2/CO ratio lower
than the ratio when using cobalt based catalyst [52], [53], [68]. In addition
to that, FT product spectrum is assumed to follow ASF distribution, this ap-
proach was taken by different papers such as Vázquez et al. [67], and Santos
et al. [69]. This model includes the production of paraffins and olefins until
molecules with 30 carbon atoms, aromatics and oxygenates are neglected
since they occur in small quantities in LTFT synthesis [50], [56]. SRK method
was used to develop this model in Aspen Plus.

32
4.2 Technical Modelling of Fischer-Tropsch synthesis

The Aspen Plus process model flowsheet is shown in Figure 19 and it consists
of these main stages:

 Syngas production
 FT synthesis
 FT-products upgrading

4.2.1 Syngas production

H2 and CO2 are the feedstock of this process at 25 0C and 1 bar. They are
mixed together with a ratio of 3:1 respectively on molar basis to prevent car-
bon deposition on the catalyst surface in the RWGS reactor [56]. The H2/CO2
mix is preheated by HX1 to 650 0C before entering the RWGS reactor [56].
RWGS reactor is modelled as an adiabatic Gibbs reactor (RGibbs) where re-
action (1) is assumed to be in equilibrium state and takes place at 800 0C and
1 bar [50], [56]. The flow rate of CO2 and H2 are chosen to achieve about 500
tonne/h of syngas at the entrance of the FTS reactor (SYN stream), which is
equivalent to the typical values from constructed FT plants [50]. In RWGS
reactor, high temperature increases the formation of products of RWGS re-
action while low pressure decreases the formation of methane, which is pro-
duced by methanation reaction [67]. Methanation reaction is a side reaction
that occurs in the RWGS reactor and is shown in equation (2) [70].

CO2 + H2 CO + H2O (1)


CO + 3H2 CH4 + H2O (2)

The product from the RWGS reactor is then cooled down by HX2 to 5 0C
and passes by adiabatic flash separator (F1) to separate water from the
stream. Then, the stream passes by pressure swing adsorption system (PSA1)
which separates H2 to control the ratio of H2/CO entering the FTS reactor
using Split1. While the excess H2 is mixed with excess H2 from other parts of
the process in MIX5 where its majority is recycled back to the beginning of
the process (H2-R stream). The PSA system has the ability to achieve 85%
hydrogen recovery and purity 99.999 mol-% [56]. PSA1 is modelled using a
separator which separates 85% of H2 from the stream with 100% purity. After
that, CO2 is separated from the stream with 80% recovery and 95% purity
[71]. PSA and CO2 separators are not modelled in detail, they are modelled
using component separator where the purity and the ratio recovered are de-
termined based on the references mentioned earlier.

33
4.2.2 FT synthesis

In the next step, syngas (SYN0 stream) is compressed to 25 bar and heated
to 220 0C in COMP1 before entering the FT reactor [50]. FT slurry reactor
(FTS) is modelled using RCSTR reactor where reactions (3), (4), and (5) takes
place [50], [68], [72]. The reactor volume is assumed to be 1325 m3 and pres-
sure loss is assumed to be 1 bar [50]. The catalyst assumed in the reactor is
precipitated promoted iron catalyst with particle density 3100 kg/m3 and bed
voidage 25% [68], [73].

n CO + (2n+1) H2 CnH2n+2 + nH2O (3)


n CO + 2n H2 CnH2n + nH2O (4)
CO + H2O CO2 + H2 (5)

Langmuir–Hinshelwood–Hougen–Watson (LHHW) kinetic models are


used for simulating reactions (3) and (4) until C30 paraffins and olefins, while
reaction (5) is assumed to be in equilibrium. The reaction rates used in this
model for simulating production of paraffins and olefins are shown in equa-
tions (6) and (7), where k1, k5, and k6 are rate constants, pH2 is the partial
pressure of hydrogen, pCO is the partial pressure of carbon monoxide, n is the
number of carbon atoms in each hydrocarbon molecule, and α is the growth
probability [68]. In this model, FT products are assumed to follow ASF dis-
tribution, so the growth probability is assumed to be constant and equals 0.9,
which is the mean of the values range of growth probability in LTFT [47]. The
detailed values of the parameters used for calculating the reaction rates can
be seen in Appendix B.

RCnH2n+2= (6)
1 𝐶𝑂 𝑛−1
1 2
1 𝐶𝑂 + 5 2
1 𝐶𝑂 1
1+
1 𝐶𝑂 + 5 2 1−

RCnH2n = (7)
1 𝐶𝑂 𝑛−1
6
1 𝐶𝑂 + 5 2
1 𝐶𝑂 1
1+
1 𝐶𝑂 + 5 2 1−

4.2.3 FT-products upgrading

FT-products are cooled down in HX3 to 60 0C to separate water and majority


of FT-gas products from FT-liquid products in adiabatic flash separator (F2)
[50]. Then, FT-liquid products (CU-feed stream) are fed into atmospheric
crude fractionation unit to divide it into gases, gasoline, diesel, and wax. The
atmospheric crude fractionation unit (DIST1) is modelled using PetroFrac
column [50]. DIST1 parameters are shown in Table 3. Gasoline and diesel are

34
end-products, while wax is further hydrocracked to produce more diesel, and
gases are combusted to produce the heat required for the process.

Gases produced from DIST1 are mixed with gases produced from F2 and
F3, which is the flash separator in the hydrocracking stage, in MIX2. From
the output 85% of H2 is separated with purity 100% in PSA2 [56]. Then, 80%
of CO2 is separated with 95% purity in CO2-SEP2 [71]. The recovered H2 from
PSA2 mixes with that from PSA1 in MIX5, where part of it is sent to the hy-
drocracking phase and the rest is recycled back to the beginning of the pro-
cess. The recovered CO2 from CO2-SEP2 mixes with that from CO2-SEP in
MIX6 where it is recycled back to the beginning of the process. The remaining
gases enters combustion chamber (CC), which is modelled as RStoic reactor,
where they’re fully combusted with air to produce heat required for the whole
process and improve the efficiency of the system.

Wax produced from DIST1 is cooled down to 340 0C in HX4 and pressur-
ized to 40 bar in PUMP [50], then it mixes with recovered H2 from SPLIT2
in MIX3 (CR stream). Recovered H2 was also heated to 340 0C and pressur-
ized to 40 bar in COMP2. The mass of recycled H2 (H2-R-CR stream) is set
to be 10.5% of the mass of wax needed to be hydrocracked [74], [75]. The
cracking reactor (CRACK) is modelled as RPlug reactor with catalyst loading
equals to 3 kgwax/kgcatalyst.h and bed voidage equals to 40% [50], [75], [76].
The hydrocracking reactions are explained by LHHW kinetic models.

As explained by Selvatico et al. [50] and Pellegrini et al. [75], wax hydroc-
racking process consists of two consecutive reactions, hydroisomerization
and hydrocracking of alkanes, as shown in reaction (8). The direct hy-
drocracking of alkanes can be neglected. In this model as a simplified hypoth-
esis, isomer alkanes are assumed to crack from the middle [75]. The reaction
rates of hydroisomerization and hydrocracking are shown in reactions (9)
and (10). The parameters in LHHW kinetic rates are calculated using the
equations shown in Table 2 [50], [75].

n-C iso-C cracking products


(8)

(9)
−𝐸 𝑛
1

Risom(n) =
0
𝑛 𝑒 𝑓𝑢𝑔𝑛−𝐶 𝑛 − 𝑓𝑢𝑔 −𝐶 𝑛
𝐾𝑒 𝑛
𝐷𝑆

Rcr(n) = (10)
−𝐸 𝑛
0 𝑛 𝑒 𝑓𝑢𝑔 𝑛
−𝐶
𝐷𝑆

ADS = 𝑓𝑢𝑔 2
∑301 𝐾 𝑛−𝐶
𝑛 𝑓𝑢𝑔𝑛−𝐶 𝑛 ∑304 𝐾 −𝐶
𝑛 𝑓𝑢𝑔 −𝐶 𝑛 (11)

35
Table 2: Equations used for calculating parameters used in hydroisomeriza-
tion and hydrocracking LHHW kinetic rates
Parameter Equation Unit
KLn-c (n) 𝑒 0 4𝑛 1/Pa
KLiso-c (n) 𝑒 0 1𝑛 1/Pa
k0cr(n) 𝑛7 63 ∗ 16 kmol/(kgcat.h)
k0isom(n) 𝑛7 7 ∗ 17 kmol/(kgcat.h)
Ecr(n) 𝑛 ∗ 4 kJ/kmol
Eisom(n) 𝑛 ∗ 4 ∗ 5 kJ/kmol
Keq(n) 2
𝑛 − 𝑛∗ −2
∗ 3 -

where k0 is the pre-exponential factor of isomerization of n-alkanes and


hydrocracking of isomers, E is the activation energy of isomerization and hy-
drocracking, Keq is the equilibrium constant between isomerization direct
and backward reactions, fug(n) is the fugacity of n-alkanes and iso-alkanes
with n carbon atoms, and KL is Langmuir constants for the adsorption of al-
kanes and isomers [50].

The products of hydrocracking phase (CR-OUT stream) are cooled down


in HX6 to 50 0C to separate gases in adiabatic Flash separator F3. Finally,
hydrocracking liquid products passes by distillation unit (DIST2) which is
modelled as RadFrac to separate wax and diesel. DIST2 parameters are
shown in Table 3.

Some calculation blocks were used to ensure the smoothness of operation


of the model. The used calculation blocks are CR-CTLYS, DIST1-D, DIST1-
W, DIST2-W, FT-T, and H2-CR. CR-CTLYS is used to ensure that catalyst
loading in CRACK equals to 3 kgwax/kgcatalyst.h. DIST1-D and DIST1-W are
used to calculate reference mass for diesel and wax respectively used in
DIST1. While DIST2-W is used to calculate reference mass for wax used in
DIST2. FT-T is used to ensure that the temperature of FTS equals to the tem-
perature of SYN stream. Finally, H2-CR is used to ensure that the mass of
recycled H2 going to hydrocracking process (H2-R-CR stream) equals to
10.5% of the mass of wax needed to be hydrocracked in WAX-HP.

The data of compounds and elements used in the model was retrieved
from Aspen Plus library if it is existing. If not, data was retrieved from differ-
ent sources such as Wessel and Jurs [77], Chemeo [78], and ChemSpider [79]
databases. Besides, Table 3 summarizes the main parameters of the main
blocks used in the model.

36
Figure 19: Aspen Plus model flowsheet

37
Table 3: Parameters of the main blocks in the model
Block Parameters

CC Rstoic reactor, Duty=0, Δp=0

Rplug reactor, diameter=2m, height=6m,


CRACK
isothermal, Δp=0, bed voidage=0.4

PetroFrac column, 54 stages, reflux ra-


tio=1.8, column Δp=1 bar, Feed on stage
DIST1 30, rebolier power=1 MWth, lateral strip-
per=12 stages, draw=stage 45, re-
turn=stage 35
RadFrac column, 15 stages, reflux ra-
DIST2 tio=1.8, Feed on stage 10, column Δp=1
bar
RCSTR reactor, isothermal, volume=1325
FTS
m3, Δp=1 bar

RWGS RGibbs reactor, isothermal

For examining the technical performance of the overall process and FT


synthesis, selectivity, CO2, CO, and H2 conversion percentages, and efficiency
of the overall process when considering electricity as input and FT-products
as output are calculated. In selectivity calculation, the FT paraffins products
selectivity is calculated and compared with their normalized selectivity when
the product follows ASF distribution and growth probability is constant and
equals 0.9 . In ASF distribution, the molar fraction of hydrocarbon products
(Mn) depends only on carbon number and growth probability, as shown in
equation (12) [80]. It is worth mentioning that there is a difference between
the experimental FT product spectrum and the product spectrum when fol-
lowing ASF distribution. As can be seen in Figure 20, in real life process, rel-
atively higher methane and less C2 products will be produced [54]

Mn = (1-α) αn-1 (12)

38
Figure 20: Relation between log of molar fraction of FT-products (LogMn)
and FT-product chain length (n) when ASF distribution is assumed (dotted
red line) and when experimental data is measured (black line)[54]

In the other parameters, gasoline, diesel, and wax are considered as end
products. CO2 and H2 conversion percentages were calculated twice. First
time when heat produced in the combustion chamber (CC), to supply the
heating requirement of the process, is considered as an end-product, and the
second time when it is not considered as an end-product. For calculating
conversion percentages, the mass flow of carbon or hydrogen atoms in the
considered output was divided by the mass of carbon or hydrogen atoms in
the considered input as shown in equations (13), (14), (15), (16), (17), and
(18). Full combustion of H2 and hydrocarbons (HC) in the CC is assumed
when heat energy is included in the conversion percentages. On the other
hand, efficiency of the overall process is calculated from electricity to FT-
products by dividing the end products (gasoline, diesel, and wax) flow rate
multiplied by their low heating value (LHV) by the electricity needed for pro-
ducing hydrogen, capturing carbon dioxide, supplying cooling needs of the
process, and supplying electric needs of the process, as shown in equation
(19). LHV of hydrogen, gasoline, diesel, and wax are assumed to be 120
MJ/kg , 43.3 MJ/kg, 42.6 MJ/kg, and 41.5 MJ/kg respectively [81]. For esti-
mating the electricity needed for producing hydrogen, PEM electrolysis is as-
sumed to be used with efficiency 80% [82]. While carbon dioxide capture is
assumed to consume 300 kWh of electricity per tonne of captured carbon
dioxide [25]. Moreover, cooling needs is assumed to be fulfilled by
wastewater source heat pump (WWSHP) which has a coefficient of operation
(COP) equals to 2.23 [83].

Table 4 shows the values used for calculating the efficiency of the overall
process.

39
CO2 conversion (heat included) = (13)
̇𝐶 𝑛𝑒
+ ̇ 𝐶𝐷 𝑒 𝑒 + ̇ 𝐶 + ̇ 𝐶 𝐶 𝑛 𝐶𝐶
̇ 𝐶𝐶𝑂 − 𝑛
2

CO2 conversion (heat not included) =


̇𝐶 𝑛𝑒
+ ̇ 𝐶𝐷 𝑒 𝑒 + ̇ 𝐶 (14)
̇ 𝐶𝐶𝑂 − 𝑛
2

(15)
H2 conversion (heat included) =
̇ + ̇ + ̇ + ̇ + ̇ 2 𝑛 𝐶𝐶
𝑛𝑒 𝐷𝑒 𝑒 𝐶 𝑛 𝐶𝐶
̇ 2− 𝑛

(16)
H2 conversion (heat not included) =
̇ 𝑛𝑒
+ ̇ 𝐷𝑒 𝑒 + ̇
̇ 2− 𝑛

(17)
CO conversion (FT synthesis) =
̇𝐶 𝐶 𝑛
̇ 𝐶𝐶𝑂 𝑛 𝑆

(18)
H2 conversion (FT synthesis) =
̇ 𝐶 𝑛
̇ 2 𝑛𝑆

(19)
Efficiency =
̇ 𝑛𝑒 ∗ 𝑛𝑒 + ̇ 𝐷𝑒 𝑒 ∗ 𝐷𝑒 𝑒 + ̇ ∗
𝐸𝑒 2 +𝐸 𝑒 𝐶𝑂2 +𝐸 𝑒 𝑛𝑔 +𝐸 𝑒 𝑢

Table 4: Parameters used in efficiency calculation


Efficiency Calculation
Cooling Electric
H2 CO2 Gasoline Diesel Wax
Utility Utility
Mass Flow rate
71.7 879.2 28.3 94.2 1.2
(tonne/hr)
LHV (MJ/kg) 120.0 43.4 42.6 41.5
LHV (MWh/tonne) 33.3 12.1 11.8 11.5
Power (MW) 2389.7 855.6 341.0 1114.7 13.5
Power (MWe) 2987.1 263.8 383.7 96.1

In the end of the technical analysis, sensitivity analysis is conducted to


examine how the FT products will change with changing different operating
conditions, such as pressure, temperature, and H2/CO ratio.

4.3 Economic analysis of Fischer-Tropsch synthesis

In the economic analysis, the levelized cost of fuel (LCOF) of FT-products is


calculated and compared with the prices of their equivalent fossil counter-
parts. In this model, hydrogen and carbon dioxide are considered as feed-
stock for the process, so the capital costs related to electrolysis and carbon
dioxide capture is not in the scope of this thesis. The purchased equipment

40
cost (Ce,i) for all process components, except for PSA and CO2 separation
blocks, are estimated by Aspen Process Economic Analyzer (APEA) using
United States dollars (USD) based on the prices in the 1st quarter of 2018.
The inflation rate between the 1st quarter of 2018 and the 1st quarter of 2021
in the United States is calculated using equation (20), where Consumer Price
Index (CPI) values are extracted from U.S. Bureau of Labor Statistics [84].
Then, all prices are converted to Euros with a conversion rate of 1 USD equal-
ing to 0.85 EUR.

Inflation Rate = (20)


𝐶 𝑛𝑢 2021 −𝐶 𝑛𝑢 2018

𝐶 𝑛𝑢 2018

APEA estimates the total capital cost (CTCC) using Icarus evaluation en-
gine, which is a mathematical model includes designs and prices for more
than 250 solid, liquid, and gas handling and process equipment, 60 plant
bulk items, 70 site preparation work, and about 12 types of buildings [85].
APEA sizes equipment used and estimates the costs based on the dimensions
[86] . CTCC estimated by APEA includes both direct and indirect costs, which
includes purchased equipment, installation, piping, instrumentation and
controls, electrical systems, service facilities, land, yard improvement, build-
ings including services, engineering and supervision, legal expenses, con-
struction expenses, contractor’s fees, and contingency [87]. CTCC is calculated
using equations (21) and (22) [86].

CISBL = 𝛴 1 𝐶𝑒 𝑓 𝑓 𝑓𝑒 𝑓𝑒 𝑓 𝑓 𝑓 𝑓 (21)

CTCC = 𝐶 𝑆 𝑂𝑆 𝐷𝐸 𝑋 (22)

where CISBL is the inside battery limits investment cost, fp, fm, fer, fel, fic, fc,
fs, and fl are the corrections factors for piping, material, equipment erection,
electrical, instrumentation and control, civil work, structures and buildings,
and lagging and paint respectively. Also, OS is the percentage of outside bat-
tery limit costs, DE is the design and engineering percentage, and X is the
contingency percentage.

The purchased equipment cost of PSA blocks are estimated based on the
work done by Marcoberardino et al. [88]. While CO2 separation blocks pur-
chased equipment costs are calculated based on the work done by Zang et al.
[45] and Rezaei and Dzuryk [89]. The costs were calculated based on the
prices in the 1st quarter of 2018 to be compatible with the prices used in
APEA. Both prices were calculated using equation (23) [88].

Ci,2018 = 𝐶 (23)
𝑓
𝑆 𝐶𝐸 𝐶 2018
0 𝑆 𝐶𝐸 𝐶 0
0

41
where Ci,2018 is the purchased equipment cost in 2018, Ci,0 is the reference
cost, Si is the actual size of equipment, Si,0 is the reference size, f is the scale
factor, and CEPCI0 is the chemical engineering plant cost index of the refer-
ence year, while CEPCI2018 is that of year 2018, which can be found in [90].

Parameters shown in Table 5 are used to calculate the high and low values
of the operating expenses (OPEX). Cooling water and refrigerant – Freon 12
costs are estimated by APEA, maintenance percentages from the total capital
cost are estimated from maintenance percentages used by Hamelinck et al.
[91] and Snehesh et al. [92], while the rest of the parameters were suggested
by Wärtsilä to unify the basis of economic calculation.

Table 5: OPEX calculations parameters


Parameter Value low Value high
Discount Rate (r) 6% 8%
Annual Running hours (h) 5000 8300
Economical Lifetime (Years) 20
Electricity Cost (€/MWh) 30 50
Cooling Water (€/m )3
0.032
Refrigerant - Freon 12 (€/tonne) 0.187
CO2 cost (€/tonne) 30 100
H2 cost (€/kg) 2 4
Maintenance 3% 5%

After calculating the total capital cost and the OPEX of the plant, the lev-
elized cost of fuel (LCOF) for low and high values of FT products are calcu-
lated and compared with the price of their fossil based equivalent products
without adding tariff or taxes [93], [94]. The ratio in prices between fossil
gasoline, diesel, and wax act as a reference for determining the levelized cost
of each independent product from the FT process. LCOF is calculated using
equation (24).

(24)
𝐶 𝐶𝐶 +𝐶 +𝐶 2 +𝐶𝐶𝑂2 +𝐶 𝑛𝑔 +𝐶𝑒 𝑒

LCOF =
𝛴𝑛 0
1+ 𝑛
̇ 𝑛𝑒 + ̇ 𝐷 𝑒 𝑒 + ̇
𝛴𝑛 0
1+ 𝑛

Furthermore, the effect of different FT operating parameters, such as tem-


perature, pressure, and H2/CO ratio, on the revenue is investigated, while
assuming fixed capital cost and fixed price for the produced fuel. Moreover,

42
a sensitivity analysis of the ratio between the prices of FT products and fossil
fuel is conducted to examine how it changes by varying H2, CO2, and total
capital costs.

43
5 Results
The technical performance and economic feasibility of the Aspen model are
presented in detail in this chapter.

5.1 Technical performance

After calculating the selectivity of FT paraffins products, it was found that, as


shown in Figure 21, the selectivity of the paraffins produced in the model is
equal to the normalized ideal ASF selectivity. Therefore, the molar selectivity
of FT-products produced in this model follows ASF distribution, as assumed
in calculating LHHW reaction rates of FT synthesis. From these calculations,
we can find that the C5+ molar selectivity is 64.09%.

12%

10%
Selectivity of products (%)

8%

6%

4%

2%

0%
0 5 10 15 20 25 30 35
Number of carbon atoms (n)

Aspen Plus Model Normalized Ideal ASF

Figure 21: Relation between carbon selectivity of FT products in Aspen


Plus model and in normalized ideal ASF distribution

Considering other parameters calculations, Hannula et al. [95] mentioned


that CO2 to FT products carbon conversion when using recycling configura-
tion is between 65% and 89%, while the efficiency is from 33% to 41% when
using low temperature electrolysis and considering electricity to FT syncrude
route. Besides, Schmidt et al. [43] mentioned that electricity to FT-products
efficiency when using low temperature electrolysis is from 38% to 53%, de-
pending on the source of CO2 captured. Also, Dieterich et al. [41] mentioned
that CO carbon conversion of FT synthesis is about 75%. Table 6 summarizes
the different parameters calculated for assessing the technical performance

44
of the model. It can be seen that parameters calculated in this model are in
the same range as previously reported by other papers. The calculated effi-
ciency equals about 39% and is in the same range reported by both Hannula
et al. [95] and Schmidt et al. [43]. Besides, CO2 carbon conversion percent-
ages of the overall process, when heat is included or not, equal about 89%
and 75% respectively and lie in the same range reported by Hannula et al.
[95]. Moreover, CO carbon conversion percentage in FT-synthesis equals
about 75% and is equal to the value mentioned by Dieterich et al. [41].

Table 6: Parameters showing the performance of the Aspen Plus model


Parameter Percentage Scope
CO2 Carbon conversion (heat included) 88.69%
CO2 Carbon conversion (heat not in-
74.43%
cluded) Overall
H2 conversion (heat included) 35.76% Process
H2 conversion (heat not included) 26.24%
Efficiency (from electricity to FT-products) 39.38%
CO Carbon conversion 74.91%
FT Synthesis
H2 Conversion 41.78%

After calculating the main technical efficiencies of the model, a sensitivity


analysis was done to examine how different parameters are affecting FT C5+
productivity in FT reactor and the recovered and recycled hydrogen (H2R) in
the process. The main case is modelled at P=25 bar, T=220 0C, and H2/CO =
2, only one parameter is varied at a time to examine its independent effect.
Figure 22, Figure 23, and Figure 24 show the effect of pressure, temperature,
and H2/CO ratio respectively, the revenue part will be discussed in the eco-
nomic analysis. In addition to that, the effect of the operating conditions on
the growth probability is not considered since it is assumed to be constant.

Although FT C5+ productivity increases with increasing pressure, temper-


ature, until 235 0C, and H2/CO ratio in Figure 22, Figure 23, and Figure 24
respectively, the slope is not constant, also the amount of H2R decreases,
which will increase the operating cost of the process. So, the economic anal-
ysis of the process will play an important role in determining the optimal op-
erating conditions.

45
140 € 60

Millions
120 € 40
Mass Flow rate (tonne/hr)
100 € 20

80 €-

60 € (20)

40 € (40)

20 € (60)

0 € (80)
10 12 14 16 18 20 22 24 25 26 28 30
Pressure (bar)

H2R FTP-C5+ Revenue

Figure 22: Sensitivity analysis of FT synthesis pressure, T=220 0C,


H2/CO=2

160 € 40

Millions
140 € 20
Mass Flow rate (tonne/hr)

120 €-
100 € (20)
80 € (40)
60 € (60)
40 € (80)
20 € (100)
0 € (120)
180 190 200 210 220 230 240 250
Temperture (0C)

H2R FTP-C5+ Revenue

Figure 23: Sensitivity analysis of FT synthesis temperature, P=25bar,


H2/CO=2

46
160 € 50

Millions
140
€-
Mass Flow rate (tonne/hr)
120
€ (50)
100
80 € (100)
60
€ (150)
40
€ (200)
20
0 € (250)
0,50 0,75 1,00 1,25 1,50 1,75 2,00 2,25 2,50
H2/CO molar ratio

H2R FTP-C5+ Revenue

Figure 24: Sensitivity analysis of FT synthesis H2/CO ratio, T=220 0C,


P=25bar

5.2 Economic Feasibility

The LCOF of different FT-products produced in this model are summarized


in Table 7. It was found that their prices are higher than their equivalent fossil
fuels by 2.5 to 5 times. These values lie in the same range reported by Dieter-
ich et al. [41], where FT diesel price was reported to be higher than fossil die-
sel by 2 to 7 times.

Table 7: LCOF of FT-products


Value low Value high
Gasoline (€/tonne) 1616.63 € 3180.43 €
Diesel (€/tonne) 1415.02 € 2783.81 €
Wax (€/tonne) 2687.68 € 5287.53 €
Ratio with fossil
2.5 5.0
product

The main parameters affecting the FT-products prices were investigated.


It was found that H2 cost has the highest contribution in the prices, about
80%, then followed by CO2 cost and total capital costs, as shown in Figure 25.
Furthermore, sensitivity analysis was conducted on the three main contribu-
tors, to examine how the change in their prices will change the cost ratio be-
tween FT-products and fossil products. Figure 26, Figure 27, and Figure 28

47
show the relation between changing the prices of H2, CO2, and Capex respec-
tively, in comparison with the main case scenario prices, and the ratio be-
tween FT-products prices and the prices of their equivalent fossil based prod-
ucts. We can see in Figure 26 that hydrogen prices can decrease the low value
ratio to be about 1, when hydrogen cost equals 0.5 €/kg, and increase the
high value ratio to be about 6, when hydrogen cost equals 2.5 €/kg. On the
other hand, Figure 27 and Figure 28 show that CO2 price and Capex devia-
tions from the prices used in the main case scenario have small effect on the
low and high values of FT end-products. Therefore, decreasing the price of
hydrogen is a must to have FT-products economically competitive with their
fossil counterparts.

100%

80%

60%

40%

20%

0%
Low Value High Value

Capital Cost O&M Hydrogen Cost CO2 cost Cooling cost Electricity Cost

Figure 25: Contribution percentages of cost sources

Figure 26: Relation between FT-products cost divided by fossil-based cost


and change in H2 prices

48
Figure 27: Relation between FT-products cost divided by fossil-based cost
and change in CO2 prices

Figure 28: Relation between FT-products cost divided by fossil-based cost


and change in Capex

For examining the effect of changing FT parameters on the revenue, FT


products prices are considered constant. Besides, only H2 and CO2 variability
are considered, while Capex, maintenance, and utilities costs are considered
constant. The change of revenue with changing pressure, temperature, and
H2/CO ratio are shown in Figure 22, Figure 23, and Figure 24 respectively.
We can see in Figure 22 that increasing the operating pressure increases the
revenue, since it increases both CO conversion rate and selectivity of paraffin
over olefin, so more liquid hydrocarbons can be produced [96] .While Figure
23 shows that increasing temperature increases the revenue, due to the in-
crease of CO conversion rate, until it reaches a specific point, 225 0C in this
model, where after it the increase in selectivity towards shorter chain hydro-
carbons outweigh the benefits from increasing the CO conversion rate [96].
Similarly, Figure 24 shows that revenue increases with increasing H2/CO mo-
lar ratio, until about 2.3 in this model, then it starts to decrease, since in-
creasing H2/CO molar ratio increases CO conversion but decreases the selec-
tivity towards longer chain hydrocarbons [97]. In this model the revenue
slightly changes, whether in a positive or negative direction, after H2/CO mo-
lar ratio of 2 is reached. Moreover, it is worth mentioning that, in real life,

49
growth probability is not constant. So, increasing the temperature and
H2/CO molar ratio will increase the reaction rates but it will shift the prod-
ucts spectrum more into shorter hydrocarbons, which will change the opti-
mal operating conditions for producing liquid fuels.

50
6 Conclusion
Renewable energy, with solar and wind energies in its core, can pave the road
for a defossilized energy and transport sector with PtX technologies as one of
the keys required to achieve the climate targets. PtX can improve the integra-
tion of the intermittent renewable energies by adding more flexibility to the
grid and providing inexpensive energy storage options. Moreover, PtG and
PtL based hydrocarbons can directly replace fossil hydrocarbons and utilize
their existing infrastructure. In this thesis, a techno-economic analysis of CO2
and H2 based FT synthesis was carried out.

FT synthesis is a process where syngas, which is a mixture of CO and H2,


is converted to a large range of hydrocarbons. Among these hydrocarbons are
gasoline, diesel, kerosene, naphtha, wax, and olefin. FT synthesis isn’t a dy-
namic operation since the output hydrocarbons are highly affected by the op-
eration conditions. Because of that, if green hydrogen production is done in
the same facility where FT synthesis takes place and dynamic operation is
needed for providing demand response and producing hydrogen only when
electricity prices are low, hydrogen storage has to be added which will in-
crease the capital cost of the plant. Although carbon neutral FT hydrocar-
bons’ costs are higher than their equivalent fossil counterparts, from 2.5 to 5
times more in the model described in this thesis, they can directly replace
fossil hydrocarbons in different sectors from transportation to electricity
generation and industry with utilizing the same infrastructure, which highly
decreases GHG emissions. The GHG emissions from PtL via FT is currently
lower than fossil fuel by 68-87.4%.

For making PtL via FT synthesis ready for large scale production, testing
RWGS in operational conditions is needed to increase its TRL, since its TRL
is currently 6. While FT synthesis and product upgrading are mature tech-
nologies and already used in large scale production facilities. Therefore, more
efforts should be given to increase the TRL of PtL via FT synthesis, since FT
fuels can make immediate reduction in GHG emissions without sacrificing
the stability or the reliability of the energy and transport sector.

Moreover, FT synthesis is affected by the operating conditions, such as


pressure, temperature, and H2/CO molar ratio. Increasing the pressure in-
creases the produced liquid fuels, but, at the same time, increases OPEX and
Capex, since more electricity is needed to compress syngas, and more expen-
sive pipes and reactor tanks are needed for operating in higher pressure.
Moreover, increasing temperature and H2/CO molar ratio increase the pro-
duced liquid fuels, but, at the same time, increase the selectivity towards
shorter chain hydrocarbons. Besides, since carbon-neutral FT synthesis
mainly consists of three different parts, which are RWGS, FT synthesis, and
51
product upgrading, optimizing the operating conditions of the whole system
with respect to each other is required to maximize the efficiency, which ena-
bles having the highest production of liquid fuels with the lowest possible
cost. The efficiency calculated in the model developed in this thesis was about
39%

In addition to that, the key factor for reducing the cost of FT hydrocarbons
is the price of hydrogen, which contributes to about 80% of the LCOF of FT
hydrocarbons. According to the model in this thesis, if hydrogen prices
reached 0.5 €/kg, FT hydrocarbons will be economically competitive with
their fossil equivalent hydrocarbons. In general, green hydrogen production
can be optimally carried out in locations where excellent renewable sources
exist and low electricity prices via solar and wind energies can be achieved,
to have currently the lowest possible hydrogen prices which decreases FT
fuels’ costs and make it as economically attractive as possible especially when
carbon emissions prices are added to the prices of fossil fuels. In addition to
that, more research is needed to decrease the cost of hydrogen production via
electrolysers.

To conclude, carbon-neutral FT synthesis has the potential to contribute


to defossilizing the energy and transport sector, in short and medium terms,
by producing different near carbon-neutral hydrocarbons needed in this sec-
tor. However, more work needs to be done to make this technology ready for
large-scale production and decrease FT-fuels cost so that FT-fuels can be
more economically competitive with their fossil counterparts.

52
References
[1] V. Masson-Delmotte, P. Zhai, A. Pirani, S. L. Connors, C. Pean,
S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell,
E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, Ö. Yelekci, R. Yu,
and B. Zhou, Eds., “Summary for policymakers,” in Climate Change 2021:
The Physical Science Basis. Contribution of Working Group I to the Sixth
Assessment Report of the Intergovernmental Panel on Climate Change,
Cambridge University Press, 2021.
[2] “The Paris Agreement | UNFCCC.” https://unfccc.int/process-
and-meetings/the-paris-agreement/the-paris-agreement (accessed Dec. 12,
2021).
[3] “Summary report 31 October – 12 November 2021,” IISD Earth
Negotiations Bulletin. https://enb.iisd.org/Glasgow-Climate-Change-Con-
ference-COP26-summary (accessed Dec. 12, 2021).
[4] H. Ritchie and M. Roser, “CO₂ and Greenhouse Gas Emissions,”
Our World in Data, May 2020, Accessed: Dec. 09, 2021. [Online]. Available:
https://ourworldindata.org/co2-emissions
[5] “Grid stability and 100% renewables,” pv magazine Interna-
tional. https://www.pv-magazine.com/2021/12/08/grid-stability-and-100-
renewables/ (accessed Dec. 12, 2021).
[6] J. Burre, D. Bongartz, L. Brée, K. Roh, and A. Mitsos, “Power-to-
X: Between Electricity Storage, e-Production, and Demand Side Manage-
ment,” Chemie Ingenieur Technik, vol. 92, no. 1–2, pp. 74–84, 2020, doi:
10.1002/cite.201900102.
[7] “Power-to-Liquids - Fraunhofer ISE,” Fraunhofer Institute for
Solar Energy Systems ISE. https://www.ise.fraunhofer.de/en/business-ar-
eas/hydrogen-technologies-and-electrical-energy-storage/thermochemical-
processes/power-to-liquids.html (accessed Dec. 12, 2021).
[8] A. Krylova, “Products of the Fischer-Tropsch synthesis (A Re-
view),” Solid Fuel Chemistry, vol. 48, Jan. 2014, doi:
10.3103/S0361521914010030.
[9] Y. Liu, H. Kamata, H. Ohara, Y. Izumi, D. S. W. Ong, J. Chang,
C. K. Poh, L. Chen, and A. Borgna, “Low-Olefin Production Process Based on
Fischer–Tropsch Synthesis: Process Synthesis, Optimization, and Techno-
Economic Analysis,” Ind. Eng. Chem. Res., vol. 59, no. 18, pp. 8728–8739,
May 2020, doi: 10.1021/acs.iecr.0c00542.
[10] “Transport Fuels.” https://studentenergy.org/distribu-
tion/transport-fuels/ (accessed Dec. 12, 2021).
[11] J. C. Koj, C. Wulf, and P. Zapp, “Environmental impacts of
power-to-X systems - A review of technological and methodological choices
in Life Cycle Assessments,” Renewable and Sustainable Energy Reviews,
vol. 112, pp. 865–879, Sep. 2019, doi: 10.1016/j.rser.2019.06.029.

53
[12] D. Welle (www.dw.com), “Power-to-X: The secret to a 100% re-
newable energy system? | DW | 19.12.2019,” DW.COM.
https://www.dw.com/en/power-to-x-the-secret-to-a-100-renewable-en-
ergy-system/a-51662014 (accessed Jun. 22, 2021).
[13] International Renewable Energy Agency (IRENA), Innovation
landscape for a renewable-powered future: Solutions to integrate variable
renewables. International Renewable Energy Agency, Abu Dhabi, 2019.
[14] I. B. Sperstad, M. Z. Degefa, and G. Kjølle, “The impact of flexible
resources in distribution systems on the security of electricity supply: A liter-
ature review,” Electric Power Systems Research, vol. 188, p. 106532, Nov.
2020, doi: 10.1016/j.epsr.2020.106532.
[15] Frontier Economics, “INTERNATIONAL ASPECTS OF A
POWER-TO-X ROADMAP.” Weltenergierat - Deutschland, 2018.
[16] R. Daiyan, I. MacGill, and R. Amal, “Opportunities and Chal-
lenges for Renewable Power-to-X,” ACS Energy Lett., vol. 5, no. 12, pp.
3843–3847, Dec. 2020, doi: 10.1021/acsenergylett.0c02249.
[17] C. Schnuelle, J. Thoeming, T. Wassermann, P. Thier, A. von
Gleich, and S. Goessling-Reisemann, “Socio-technical-economic assessment
of power-to-X: Potentials and limitations for an integration into the German
energy system,” Energy Research & Social Science, vol. 51, pp. 187–197, May
2019, doi: 10.1016/j.erss.2019.01.017.
[18] International Renewable Energy Agency (IRENA), Innovation
landscape brief: Renewable power-to-heat. 2019.
[19] A. Pollard and B. Berg, Heat pump performance. 2019. doi:
10.13140/RG.2.2.25116.13449.
[20] J. M. Cullen and J. M. Allwood, “Theoretical efficiency limits for
energy conversion devices,” Energy, vol. 35, no. 5, pp. 2059–2069, May
2010, doi: 10.1016/j.energy.2010.01.024.
[21] C. Luerssen, O. Gandhi, T. Reindl, K. W. D. Cheong, and C. Sek-
har, Levelised Cost of Thermal Energy Storage and Battery Storage to Store
Solar PV Energy for Cooling Purpose. 2018, p. 10. doi: 10.18086/eu-
rosun2018.04.09.
[22] A. R. Dehghani-Sanij, E. Tharumalingam, M. B. Dusseault, and
R. Fraser, “Study of energy storage systems and environmental challenges of
batteries,” Renewable and Sustainable Energy Reviews, vol. 104, pp. 192–
208, Apr. 2019, doi: 10.1016/j.rser.2019.01.023.
[23] B. Elmegaard, F. Holm, and F. Bühler, Potentials for the electri-
fication of industrial processes in Denmark. 2019.
[24] C. Wulf, J. Linssen, and P. Zapp, “Chapter 9 - Power-to-Gas—
Concepts, Demonstration, and Prospects,” in Hydrogen Supply Chains, C.
Azzaro-Pantel, Ed. Academic Press, 2018, pp. 309–345. doi: 10.1016/B978-
0-12-811197-0.00009-9.
[25] S. Jackson and E. Brodal, “Optimization of the Energy Con-
sumption of a Carbon Capture and Sequestration Related Carbon Dioxide

54
Compression Processes,” Energies, vol. 12, no. 9, Art. no. 9, Jan. 2019, doi:
10.3390/en12091603.
[26] S. R. Foit, I. C. Vinke, L. G. J. de Haart, and R.-A. Eichel, “Power-
to-Syngas: An Enabling Technology for the Transition of the Energy Sys-
tem?,” Angewandte Chemie International Edition, vol. 56, no. 20, pp. 5402–
5411, 2017, doi: 10.1002/anie.201607552.
[27] K. Ghaib and F. Z. Ben Fares, “Power-to-Methane: A state-of-
the-art review,” Renewable and Sustainable Energy Reviews, vol. 81, pp.
433–446, Jan. 2018, doi: 10.1016/j.rser.2017.08.004.
[28] International Energy Agency (IEA), The Future of Hydrogen:
Seizing today’s opportunities. 2019.
[29] J. Laurikko, J. Ihonen, J. Kiviaho, O. Himanen, R. Weiss, V. Saa-
rinen, J. Kärki, and M. Hurskainen, “NATIONAL HYDROGEN ROADMAP,”
p. 48, Nov. 2020.
[30] “Global CO2 emissions in 2019 – Analysis,” International En-
ergy Agency (IEA). https://www.iea.org/articles/global-co2-emissions-in-
2019 (accessed Dec. 15, 2021).
[31] I. Staffell, D. Scamman, A. V. Abad, P. Balcombe, P. E. Dodds,
P. Ekins, N. Shah, and K. R. Ward, “The role of hydrogen and fuel cells in the
global energy system,” Energy Environ. Sci., vol. 12, no. 2, pp. 463–491, Feb.
2019, doi: 10.1039/C8EE01157E.
[32] I. Dincer, “Green methods for hydrogen production,” Interna-
tional Journal of Hydrogen Energy, vol. 37, no. 2, pp. 1954–1971, Jan. 2012,
doi: 10.1016/j.ijhydene.2011.03.173.
[33] A. Amid, D. Mignard, and M. Wilkinson, “Seasonal storage of
hydrogen in a depleted natural gas reservoir,” International Journal of Hy-
drogen Energy, vol. 41, no. 12, pp. 5549–5558, Apr. 2016, doi:
10.1016/j.ijhydene.2016.02.036.
[34] A. M. Elberry, J. Thakur, A. Santasalo-Aarnio, and M. Larmi,
“Large-scale compressed hydrogen storage as part of renewable electricity
storage systems,” International Journal of Hydrogen Energy, vol. 46, no.
29, pp. 15671–15690, Apr. 2021, doi: 10.1016/j.ijhydene.2021.02.080.
[35] M. W. Melaina, O. Antonia, and M. Penev, “Blending Hydrogen
into Natural Gas Pipeline Networks: A Review of Key Issues,” National Re-
newable Energy Lab. (NREL), Golden, CO (United States), NREL/TP-5600-
51995, Mar. 2013. doi: 10.2172/1068610.
[36] S. E. Hosseini and B. Butler, “An overview of development and
challenges in hydrogen powered vehicles,” International Journal of Green
Energy, vol. 17, no. 1, pp. 13–37, Jan. 2020, doi:
10.1080/15435075.2019.1685999.
[37] M. Götz, J. Lefebvre, F. Mörs, A. McDaniel Koch, F. Graf, S. Ba-
johr, R. Reimert, and T. Kolb, “Renewable Power-to-Gas: A technological and
economic review,” Renewable Energy, vol. 85, pp. 1371–1390, Jan. 2016,
doi: 10.1016/j.renene.2015.07.066.

55
[38] T. Schaaf, J. Grünig, M. R. Schuster, T. Rothenfluh, and A. Orth,
“Methanation of CO2 - storage of renewable energy in a gas distribution sys-
tem,” Energy, Sustainability and Society, vol. 4, no. 1, p. 2, Dec. 2014, doi:
10.1186/s13705-014-0029-1.
[39] I. Wender, “Reactions of synthesis gas,” Fuel Processing Tech-
nology, vol. 48, no. 3, pp. 189–297, Sep. 1996, doi: 10.1016/S0378-
3820(96)01048-X.
[40] J. G. Speight, “6 - Production of syngas, synfuel, bio-oils, and bi-
ogas from coal, biomass, and opportunity fuels,” in Fuel Flexible Energy
Generation, J. Oakey, Ed. Boston: Woodhead Publishing, 2016, pp. 145–174.
doi: 10.1016/B978-1-78242-378-2.00006-7.
[41] V. Dieterich, A. Buttler, A. Hanel, H. Spliethoff, and S. Fendt,
“Power-to-liquid via synthesis of methanol, DME or Fischer–Tropsch-fuels:
a review,” Energy Environ. Sci., vol. 13, no. 10, pp. 3207–3252, Oct. 2020,
doi: 10.1039/D0EE01187H.
[42] A. Varone and M. Ferrari, “Power to liquid and power to gas: An
option for the German Energiewende,” Renewable and Sustainable Energy
Reviews, vol. 45, pp. 207–218, May 2015, doi: 10.1016/j.rser.2015.01.049.
[43] P. Schmidt, W. Weindorf, A. Roth, V. Batteiger, and F. Riegel,
Power-to-Liquids – Potentials and Perspectives for the Future Supply of Re-
newable Aviation Fuel. Umweltbundesamt, 2016. Accessed: Jul. 17, 2021.
[Online]. Available: https://www.umweltbundesamt.de/en/publika-
tionen/power-to-liquids-potentials-perspectives-for-the
[44] A. L. Dicks, “PEM Fuel Cells: Applications,” in Reference Module
in Earth Systems and Environmental Sciences, Elsevier, 2020. doi:
10.1016/B978-0-12-819727-1.00004-2.
[45] G. Zang, P. Sun, A. A. Elgowainy, A. Bafana, and M. Wang, “Per-
formance and cost analysis of liquid fuel production from H2 and CO2 based
on the Fischer-Tropsch process,” Journal of CO2 Utilization, vol. 46, p.
101459, Apr. 2021, doi: 10.1016/j.jcou.2021.101459.
[46] W. Gao, Q. Zhu, and D. Ma, “Nanostructured Catalyst for
Fischer–Tropsch Synthesis,” Chinese Journal of Chemistry, vol. 36, no. 9,
pp. 798–808, 2018, doi: 10.1002/cjoc.201800146.
[47] J. van de Loosdrecht, F. G. Botes, I. M. Ciobica, A. Ferreira, P.
Gibson, D. J. Moodley, A. M. Saib, J. L. Visagie, C. J. Weststrate, and J. W.
Niemantsverdriet, “7.20 - Fischer–Tropsch Synthesis: Catalysts and Chem-
istry,” in Comprehensive Inorganic Chemistry II (Second Edition), J. Reed-
ijk and K. Poeppelmeier, Eds. Amsterdam: Elsevier, 2013, pp. 525–557. doi:
10.1016/B978-0-08-097774-4.00729-4.
[48] H. Mahmoudi, M. Mahmoudi, O. Doustdar, H. Jahangiri, A.
Tsolakis, S. Gu, and M. LechWyszynski, “A review of Fischer Tropsch synthe-
sis process, mechanism, surface chemistry and catalyst formulation,” Biofu-
els Engineering, vol. 2, no. 1, pp. 11–31, Dec. 2017, doi: 10.1515/bfuel-2017-
0002.

56
[49] A. Stranges, “A History of the Fischer-Tropsch Synthesis in Ger-
many 1926–45,” Studies in Surface Science and Catalysis - STUD SURF SCI
CATAL, vol. 163, pp. 1–27, Dec. 2007, doi: 10.1016/S0167-2991(07)80469-1.
[50] D. Selvatico, A. Lanzini, and M. Santarelli, “Low Temperature
Fischer-Tropsch fuels from syngas: Kinetic modeling and process simulation
of different plant configurations,” Fuel, vol. 186, pp. 544–560, Dec. 2016, doi:
10.1016/j.fuel.2016.08.093.
[51] A. de Klerk, “Fischer–Tropsch refining: technology selection to
match molecules,” Green Chemistry, vol. 10, no. 12, pp. 1249–1279, 2008,
doi: 10.1039/B813233J.
[52] A. Sahir, Y. Zhang, E. Tan, and L. Tao, “Understanding the role
of Fischer–Tropsch reaction kinetics in techno‐economic analysis for co‐con-
version of natural gas and biomass to liquid transportation fuels,” Biofuels,
Bioproducts and Biorefining, vol. 13, Jul. 2019, doi: 10.1002/bbb.2035.
[53] M. Gupta and J. J. Spivey, “Chapter 5 - Catalytic Processes for
the Production of Clean Fuels,” in New and Future Developments in Cataly-
sis, S. L. Suib, Ed. Amsterdam: Elsevier, 2013, pp. 87–126. doi:
10.1016/B978-0-444-53870-3.00005-8.
[54] J. Cheng, P. Hu, P. Ellis, S. French, G. Kelly, and C. M. Lok, “A
DFT study of the chain growth probability in Fischer–Tropsch synthesis,”
Journal of Catalysis, vol. 257, no. 1, pp. 221–228, Jul. 2008, doi:
10.1016/j.jcat.2008.05.006.
[55] M. Marchese, G. Buffo, M. Santarelli, and A. Lanzini, “CO2 from
direct air capture as carbon feedstock for Fischer-Tropsch chemicals and
fuels: Energy and economic analysis,” Journal of CO2 Utilization, vol. 46, p.
101487, Apr. 2021, doi: 10.1016/j.jcou.2021.101487.
[56] I. Dimitriou, P. García-Gutiérrez, R. H. Elder, R. M. Cuéllar-
Franca, A. Azapagic, and R. W. K. Allen, “Carbon dioxide utilisation for pro-
duction of transport fuels: process and economic analysis,” Energy Environ.
Sci., vol. 8, no. 6, pp. 1775–1789, Jun. 2015, doi: 10.1039/C4EE04117H.
[57] Z. He, M. Cui, Q. Qian, J. Zhang, H. Liu, and B. Han, “Synthesis
of liquid fuel via direct hydrogenation of CO2,” PNAS, vol. 116, no. 26, pp.
12654–12659, Jun. 2019, doi: 10.1073/pnas.1821231116.
[58] Y. H. Choi, Y. J. Jang, H. Park, W. Y. Kim, Y. H. Lee, S. H. Choi,
and J. S. Lee, “Carbon dioxide Fischer-Tropsch synthesis: A new path to car-
bon-neutral fuels,” Applied Catalysis B: Environmental, vol. 202, pp. 605–
610, Mar. 2017, doi: 10.1016/j.apcatb.2016.09.072.
[59] The Business Research Company, “Global Refined Petroleum
Products Market Data And Industry Growth Analysis,” Dec. 2020. Accessed:
Oct. 26, 2021. [Online]. Available: https://www.thebusinessresearchcom-
pany.com/report/refined-petroleum-products-global-market-report-2020-
30-covid-19-impact-and-recovery

57
[60] International Energy Agency (IEA), “Oil Market Report - June
2021 – Analysis,” IEA. https://www.iea.org/reports/oil-market-report-june-
2021 (accessed Oct. 26, 2021).
[61] FuelsEurope, “FuelsEurope Statistical Report 2021,”
FuelsEurope. https://www.fuelseurope.eu/publication/fuelseurope-statisti-
cal-report-2021/ (accessed Oct. 26, 2021).
[62] D. Yergin, “Ensuring Energy Security,” Foreign Affairs, vol. 85,
no. 2, pp. 69–82, 2006, doi: 10.2307/20031912.
[63] L. Čuček, J. J. Klemeš, and Z. Kravanja, “Chapter 5 - Overview of
environmental footprints,” in Assessing and Measuring Environmental Im-
pact and Sustainability, J. J. Klemeš, Ed. Oxford: Butterworth-Heinemann,
2015, pp. 131–193. doi: 10.1016/B978-0-12-799968-5.00005-1.
[64] T. Takeshita and K. Yamaji, “Important roles of Fischer–Trop-
sch synfuels in the global energy future,” Energy Policy, vol. 36, no. 8, pp.
2773–2784, Aug. 2008, doi: 10.1016/j.enpol.2008.02.044.
[65] F. Perera, “Pollution from Fossil-Fuel Combustion is the Lead-
ing Environmental Threat to Global Pediatric Health and Equity: Solutions
Exist,” Int J Environ Res Public Health, vol. 15, no. 1, p. 16, Jan. 2018, doi:
10.3390/ijerph15010016.
[66] “How air pollution is destroying our health.”
https://www.who.int/news-room/spotlight/how-air-pollution-is-destroy-
ing-our-health (accessed Oct. 21, 2021).
[67] F. V. Vázquez, J. Koponen, V. Ruuskanen, C. Bajamundi, A.
Kosonen, P. Simell, J. Ahola, C. Frilund, J. Elfving, M. Reinikainen, N. Heik-
kinen, J. Kauppinen, and P. Piermartini, “Power-to-X technology using re-
newable electricity and carbon dioxide from ambient air: SOLETAIR proof-
of-concept and improved process concept,” Journal of CO2 Utilization, vol.
28, pp. 235–246, 2018, doi: 10.1016/j.jcou.2018.09.026.
[68] E. S. Lox and G. F. Froment, “Kinetics of the Fischer-Tropsch
reaction on a precipitated promoted iron catalyst. 2. Kinetic modeling,” Ind.
Eng. Chem. Res., vol. 32, no. 1, pp. 71–82, Jan. 1993, doi:
10.1021/ie00013a011.
[69] G. R. S. Santos, O. M. Basha, R. Wang, H. Ashkanani, and B.
Morsi, “Techno-economic assessment of Fischer-Tropsch synthesis and di-
rect methane-to-methanol processes in modular GTL reactors,” Catalysis
Today, vol. 371, pp. 93–112, Jul. 2021, doi: 10.1016/j.cattod.2020.07.012.
[70] M. Seemann and H. Thunman, “9 - Methane synthesis,” in Sub-
stitute Natural Gas from Waste, M. Materazzi and P. U. Foscolo, Eds. Aca-
demic Press, 2019, pp. 221–243. doi: 10.1016/B978-0-12-815554-7.00009-
X.
[71] “CO2 Removal from Syngas,” Membrane Technology and Re-
search. https://www.mtrinc.com/our-business/refinery-and-syngas/co2-
removal-from-syngas/ (accessed Nov. 02, 2021).

58
[72] L. Sehabiague and B. I. Morsi, “Modeling and Simulation of a
Fischer–Tropsch Slurry Bubble Column Reactor Using Different Kinetic Rate
Expressions for Iron and Cobalt Catalysts,” International Journal of Chemi-
cal Reactor Engineering, vol. 11, no. 1, pp. 309–330, Jul. 2013, doi:
10.1515/ijcre-2012-0017.
[73] J. M. Fox, B. D. Degen, G. Cady, F. D. Deslate, R. L. Summers, A.
Akgerman, and J. M. Smith, “Slurry reactor design studies,” Bechtel Group,
Inc., San Francisco, CA (USA), DOE/PC/89867-T2, Jun. 1990. doi:
10.2172/6094135.
[74] S. Gamba, L. A. Pellegrini, V. Calemma, and C. Gambaro, “Liquid
fuels from Fischer–Tropsch wax hydrocracking: Isomer distribution,” Catal-
ysis Today, vol. 156, no. 1, pp. 58–64, Oct. 2010, doi: 10.1016/j.cat-
tod.2010.01.009.
[75] L. A. Pellegrini, S. Gamba, V. Calemma, and S. Bonomi, “Model-
ling of hydrocracking with vapour–liquid equilibrium,” Chemical Engineer-
ing Science, vol. 63, no. 17, pp. 4285–4291, Sep. 2008, doi:
10.1016/j.ces.2008.06.002.
[76] G. Towler and R. Sinnott, “Chapter 14 - Design of Pressure Ves-
sels,” in Chemical Engineering Design (Second Edition), G. Towler and R.
Sinnott, Eds. Boston: Butterworth-Heinemann, 2013, pp. 563–629. doi:
10.1016/B978-0-08-096659-5.00014-6.
[77] M. D. Wessel and P. C. Jurs, “Prediction of Normal Boiling
Points of Hydrocarbons from Molecular Structure,” J. Chem. Inf. Comput.
Sci., vol. 35, no. 1, pp. 68–76, Jan. 1995, doi: 10.1021/ci00023a010.
[78] “Cheméo - Chemical & Physical Properties by Cheméo.”
https://www.chemeo.com/ (accessed Nov. 04, 2021).
[79] “ChemSpider | Search and share chemistry.”
https://www.chemspider.com/ (accessed Nov. 04, 2021).
[80] K. Cheng, J. Kang, D. L. King, V. Subramanian, C. Zhou, Q.
Zhang, and Y. Wang, “Chapter Three - Advances in Catalysis for Syngas Con-
version to Hydrocarbons,” in Advances in Catalysis, vol. 60, C. Song, Ed. Ac-
ademic Press, 2017, pp. 125–208. doi: 10.1016/bs.acat.2017.09.003.
[81] “Fuels - Higher and Lower Calorific Values.” https://www.engi-
neeringtoolbox.com/fuels-higher-calorific-values-d_169.html (accessed
Dec. 15, 2021).
[82] S. Shiva Kumar and V. Himabindu, “Hydrogen production by
PEM water electrolysis – A review,” Materials Science for Energy Technolo-
gies, vol. 2, no. 3, pp. 442–454, Dec. 2019, doi: 10.1016/j.mset.2019.03.002.
[83] P. Im, X. Liu, and H. Henderson, “Operational Performance
Characterization of a Heat Pump System Utilizing Recycled Water as Heat
Sink and Heat Source in a Cool and Dry Climate,” Energies, vol. 11, no. 1, Art.
no. 1, Jan. 2018, doi: 10.3390/en11010211.
[84] “U.S. Bureau of Labor Statistics.” https://www.bls.gov/ (ac-
cessed Nov. 23, 2021).

59
[85] N. Bakhyt, “Aspen Icarus Reference Guide Icarus Evaluation En-
gine (IEE) V8.0”, Accessed: Nov. 23, 2021. [Online]. Available:
https://www.academia.edu/40036447/Aspen_Icarus_Refer-
ence_Guide_Icarus_Evaluation_Engine_IEE_V8_0
[86] J. Ruokonen, H. Nieminen, A. Rufai Dahiru, A. Laari, T.
Koiranen, P. Laaksonen, A. Vuokila, and M. Huuhtanen, “Modelling and Cost
Estimation for Conversion of Green Methanol to Renewable Liquid
Transport Fuels via Olefin Oligomerisation,” p. 1046, 2021, doi:
http://dx.doi.org/10.3390/pr9061046.
[87] J. Haydary, Chemical Process Design and Simulation: Aspen
Plus and Aspen Hysys Applications. John Wiley & Sons, 2019.
[88] G. D. Marcoberardino, D. Vitali, F. Spinelli, M. Binotti, and G.
Manzolini, “Green Hydrogen Production from Raw Biogas: A Techno-Eco-
nomic Investigation of Conventional Processes Using Pressure Swing Ad-
sorption Unit,” Processes, vol. 6, no. 3, Art. no. 3, Mar. 2018, doi:
10.3390/pr6030019.
[89] E. Rezaei and S. Dzuryk, “Techno-economic comparison of re-
verse water gas shift reaction to steam and dry methane reforming reactions
for syngas production,” Chemical Engineering Research and Design, vol.
144, pp. 354–369, Apr. 2019, doi: 10.1016/j.cherd.2019.02.005.
[90] “The Chemical Engineering Plant Cost Index,” Chemical Engi-
neering. https://www.chemengonline.com/pci-home/ (accessed Nov. 23,
2021).
[91] C. N. Hamelinck, A. P. C. Faaij, H. den Uil, and H. Boerrigter,
“Production of FT transportation fuels from biomass; technical options, pro-
cess analysis and optimisation, and development potential,” Energy, vol. 29,
no. 11, pp. 1743–1771, Sep. 2004, doi: 10.1016/j.energy.2004.01.002.
[92] A. S. Snehesh, H. S. Mukunda, S. Mahapatra, and S. Dasappa,
“Fischer-Tropsch route for the conversion of biomass to liquid fuels - Tech-
nical and economic analysis,” Energy, vol. 130, pp. 182–191, Jul. 2017, doi:
10.1016/j.energy.2017.04.101.
[93] “Fuel price breakdown,” FuelsEurope.
https://www.fuelseurope.eu/knowledge/refining-in-europe/economics-of-
refining/fuel-price-breakdown/ (accessed Nov. 24, 2021).
[94] “Argus Global Waxes | Argus Media.” https://www.argusme-
dia.com/en/oil-products/argus-global-waxes (accessed Nov. 24, 2021).
[95] I. Hannula, N. Kaisalo, and P. Simell, “Preparation of Synthesis
Gas from CO2 for Fischer–Tropsch Synthesis—Comparison of Alternative
Process Configurations,” C, vol. 6, no. 3, Art. no. 3, Sep. 2020, doi:
10.3390/c6030055.
[96] E. Makhura, J. Rakereng, O. Rapoo, and G. Danha, “Effect of the
operation parameters on the Fischer Tropsch synthesis process using differ-
ent reactors,” Procedia Manufacturing, vol. 35, pp. 349–355, Jan. 2019, doi:
10.1016/j.promfg.2019.05.051.

60
[97] A. Lillebø, E. Rytter, E. A. Blekkan, and A. Holmen, “Fischer–
Tropsch Synthesis at High Conversions on Al2O3-Supported Co Catalysts
with Different H2/CO Levels,” Ind. Eng. Chem. Res., vol. 56, no. 45, pp.
13281–13286, Nov. 2017, doi: 10.1021/acs.iecr.7b01801.

61
Compounds and elements used in the Aspen Plus
model
Table A 1: Components used in the Aspen Plus model
Non-hydrocarbon Molecules Paraffins Olefins Methyl-alkanes
H2 CH4 C2H4 C11H24
CO2 C2H6 C3H6 C12H26
CO C3H8 C4H8 C13H28
H2O C4H10 C5H10 C14H30
O2 C5H12 C6H12 C15H32
N2 C6H14 C7H14 C23H48
C7H16 C8H16 C24H50
C8H18 C9H18 C25H52
C9H20 C10H20 C26H54
C10H22 C11H22 C27H56
C11H24 C12H24 C28H58
C12H26 C13H26 C29H60
C13H28 C14H28 C30H62
C14H30 C15H30
C15H32 C16H32
C16H34 C17H34
C17H36 C18H36
C18H38 C19H38
C19H40 C20H40
C20H42 C21H42
C21H44 C22H44
C22H46 C23H46
C23H48 C24H48
C24H50 C25H50
C25H52 C26H52
C26H54 C27H54
C27H56 C28H56
C28H58 C29H58
C29H60 C30H60
C30H62

62
Reactions rates parameters for LHHW kinetic
models
Table B 1: Parameters used for calculating LHHW kinetic rates in FT synthe-
sis
Parameter Value Unit
A1 4.55E-01 kmol/(kg.s.Pa)
A5 4.32E-03 kmol/(kg.s.Pa)
A6 2.71E+06 kmol/(kg.s)
E1 1.05E+05 J/mol
E5 9.45E+04 J/mol
E6 1.32E+05 J/mol

Table B 2: Values of LHHW kinetic rates used for alkanes in FT synthesis

63
Table B 3: Values of LHHW kinetic rates used for alkenes in FT synthesis

64

You might also like