Energy Performance of Power-To-Liquid Applications

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Energy Conversion and Management: X 6 (2020) 100041

Contents lists available at ScienceDirect

Energy Conversion and Management: X


journal homepage: www.journals.elsevier.com/energy-conversion-and-management-x

Energy performance of Power-to-Liquid applications integrating biogas T


upgrading, reverse water gas shift, solid oxide electrolysis and Fischer-
Tropsch technologies
Marco Marchesea, , Emanuele Giglioa,b, Massimo Santarellia, Andrea Lanzinia

a
Energy Department, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
b
Department of Environmental Engineering, University of Calabria, Via Pietro Bucci, 87036 Rende, CS, Italy

ARTICLE INFO ABSTRACT

Keywords: Power-to-liquid (P2L) pathways represent a possible solution for the conversion of carbon dioxide into synthetic
Carbon capture and utilization value-added products. The present work analyses different power-to-liquid options for the synthesis of Fischer-
Biogas upgrading Tropsch (FT) fuels and chemicals. The FT section is integrated into a complete carbon capture and utilization
Reverse water gas shift route. The involved processes are a biogas upgrading unit for CO2 recovery, a reverse water gas shift, a solid
Solid oxide electrolysis
oxide electrolyser and a Fischer-Tropsch reactor.
Fischer-Tropsch synthesis
Power-to-Liquid
The upgrading plant produces about 1 ton/h of carbon dioxide. The recovered CO2 is fed to either a reverse
water gas shift reactor or a solid oxide electrolysis unit operating in co-electrolysis mode for the generation of
syngas. The produced syngas is fed to a Fischer-Tropsch reactor at 501 K and 25 bar for the synthesis of the
Fischer-Tropsch products, which are further separated into different classes based on their boiling point to yield
light gas, naphtha, middle distillates, light waxes and heavy waxes. The developed process model uses a detailed
carbide kinetic model to describe the formation of paraffins and olefins based on real experimental data. The
effect of Fischer-Tropsch off-gas recirculation has been studied against a one-through option. Finally, energy
integration of each configuration plant is provided. Results from process simulations show that the best model
configurations reach a plant efficiency of 81.1% in the case of solid oxide electrolyser as syngas generator, and
71.8% in the case of reverse water gas shift option, with a global carbon reduction potential of 79.4% and 81.7%,
respectively.

1. Introduction physical capture processes, upgrading of biogas, or even from the air
with direct air capture technologies (DAC) [4–6]. Furthermore, the
Since the Paris agreement of the COP21, increasing effort has been captured molecule can be transformed into synthetic products through
spent in implementing solutions that can reduce greenhouse gases thermochemical or electrochemical processes. These processes enable
emissions towards the environment. Thus, alongside the deployment of methane and syngas synthesis in the case of P2G, while P2L final out-
renewable energy technologies, the study of novel applications that can comes can include Fischer-Tropsch fuels and chemicals, methanol, di-
store such energy meanwhile utilizing CO2 of industrial processes as methyl ether (DME) and formic acid [7–9].
raw material has gained much attention [1]. In the present work, we focus on P2L chains presenting a Fischer-
Means for on-site carbon dioxide reuse include power-to-gas (P2G) Tropsch reactor (FT). The FT technology has risen in scientific interest
and power-to-liquid (P2L) applications. These concepts consider ex- as an effective application to be inserted into carbon capture and uti-
ploiting CO2 as a useful commodity, shifting it from a low-value re- lization (CCU) plants, capable of delivering the so-called syncrude [10]:
source to high-value product, meanwhile storing energy from renew- a broad mixture of synthetic hydrocarbons that can replace oil ex-
able electricity [2,3]. In the framework of P2G and P2L routes, different tracted from the ground. Syncrude accounts for hydrocarbons ranging
steps are required to transform carbon dioxide into further products. from carbon number C1 to C80+, in the form of n-paraffins, α-olefins,
CO2 can be captured from the flue gas of industrial and energy plants with a lower content of alcohols and aromatic compounds [11]. Hence,
burning fossil fuels through highly energy-intensive chemical or the great potential of the FT processes is the displacement opportunity


Corresponding author.
E-mail address: [email protected] (M. Marchese).

https://doi.org/10.1016/j.ecmx.2020.100041
Received 29 January 2020; Received in revised form 4 April 2020; Accepted 6 April 2020
Available online 11 April 2020
2590-1745/ © 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Nomenclature Abbreviations

A Area (cm2) ASR Area-specific resistance (Ωcm2)


F Faraday constant (C mole-1) ATR Auto thermal reactor
I Current (A) CCU Carbon capture and utilization
j Current density (A cm−2) CGO Gallium-doped ceria
K Equilibrium constant CPOX Partial oxidation
k Kinetic constant CRP Carbon reduction potential
n Carbon number DAC Direct air capture
ṅi-th Mole flow rate of component i-th (mol s−1) FT Fischer-Tropsch
ntot Exchange charge LHV Low Heating Value
P Pressure (bar) LSC Lanthanum strontium cobaltite
R Gas constant (kJ kmol−1 K−1) MEA Monoethanolamine
T Temperature (K) Ni Nickel
V Voltage (V) OCV Open circuit voltage
W Power (W) P2G Power to gas
α Hydrocarbon probability growth P2L Power to liquid
ΔG Gibbs free energy (kJ kmol−1) RR Recirculation rate
ΔH Enthalpy change RU Reactant utilization
Δp Pressure variation RWGS Reverse water gas shift
ΔT Temperature variation SMR Steam methane reforming
η Efficiency SOE Solid oxide electrolyser
ϴ Vacancy percentage SOEC Solid oxide electrolysis cell
Χ Conversion TN Thermoneutral
YSZ Yttria-stabilized zirconia

of wide-ranging hydrocarbons of fossil origin. Transportation sector As far as P2L routes are concerned, different solutions have been
utilizing gasoline, diesel and jet fuels, and the chemical industry uti- proposed in literature. Herz et al. [22] created a possible SOEC + FT
lizing long-chain hydrocarbons as a feedstock for chemical products can process model with SOE stacks operating in co-electrolysis mode, cal-
be inserted into a circular economy concept, where the use of fossil culating a maximum process efficiency of 68.1%. Rafiee et al. [23]
material is avoided in favour of recycled CO2 [12,13]. reached a carbon efficiency (i.e., efficiency of carbon utilization) of
The input to the Fischer-Tropsch process is synthesis gas, namely 68.2% for a system that captured CO2 from flue gases, combining it
syngas that is a mixture of CO2, H2, CO and H2O mainly. To obtain it, together with an ATR and an FT reactor. Vidal et al. [24] studied the
the captured carbon dioxide can be converted into CO through ther- integration of a DAC with low temperature electrolysis, RWGS and FT
mochemical or electrochemical devices. Thus, reverse water gas shift reactors, reaching 94% carbon efficiency and 47% plant efficiency. In
reactors (RWGS), autothermal reactors (ATR), partial oxidation re- fact, different studies on the integration of FT reactors with syngas
actors (CPOX), steam methane reformers (SMR) or solid oxide elec- generation units for P2L systems can be found. However, to the best of
trolysis cells (SOEC) can be employed. Specifically, RWGS and SOEC our knowledge, not many studies include a full process integrated with
can directly convert carbon dioxide to CO with the aid of hydrogen or the carbon capture one, and seldom focusing on a detailed products
steam, whereas CPOX, ATR and SMR use methane (or other hydro- separation analysis, too. For instance, Tagomori et al. [25] stated the
carbons) for the syngas generation [14,15]. need of properly designing and evaluating the process of distillation of
Once the syngas is fed to the Fischer-Tropsch reactor, the descrip- the FT products.
tion of the FT products distribution becomes a key aspect in evaluating In the present work, we provide a P2L system analysis from the
the whole system performance. Different kinetic approaches can be capture of CO2 to the generation and separation of the synthetic pro-
used. The overall FTS synthesis can be described by a single equation ducts. We investigate the coupling between a solvent-based biogas
like a modification of the Anderson-Flory-Shultz theory or power-law upgrading process and the Fischer-Tropsch reactor with two concurrent
kinetics [16]. Selectivity models can provide information on the re- technologies for syngas generation: one proven and commercially
actants consumption rates and specific groups of FT compounds [17]. available technology like the RWGS reactor; one less commercialized
Finally, mechanistic kinetic models allow for the identification of re- technology but with high CO2 conversion potential like the SOEC under
actants consumption rates as well as products generation rates. A recent co-electrolysis. To the authors knowledge, only one report (Comidy
review comprising the FT kinetics, beyond the scope of this work, is et al. [26]) is available in the open literature that provides some insight
provided by Santos et al. [18]. A different amount of information can be on the direct comparison of a P2L with RWGS against a P2L with SOEC
extracted from the PtL models, depending on the modelling approach technologies feeding a Fischer-Tropsch reactor. However, their analysis
employed for the FTS. For instance, Cinti et al. [19] applied in their focused on the production of light FT fuels for on-board consumption on
work on SOEC + FT the AFS distribution at a fixed chain growth aircraft carriers. Furthermore, they stated that no experimental vali-
probability α = 0.94, with a modification to account for olefins and dation of any of their employed technologies was assessed. In this work,
paraffins formation. In this regard, they could identify 5 main mole- we include a mechanistic kinetic description based on the carbide FTS
cules representative of the paraffins, 4 for the olefins and a clustered mechanisms of paraffins and olefins up to carbon number C80, experi-
molecule to account for C30+ waxes. Fazeli et al. [20] employed a one- mentally validated for the Fischer-Tropsch reactor and presented in
step reaction rate to describe the FT synthesis, with weight distribution another research work (Marchese et al. [27]). Specifically, the model
based on lumped species experimental data. Selvatico et al. [21], in- provides detailed information about the production rate of each of the
stead, used detailed kinetics for olefins and paraffins, but only up to heaviest FT fraction compounds, that are generally clustered into one
carbon number C30, thus excluding the heaviest FT fractions of interest single C30+ pseudo-component [21], from which exact FTS heat of
if targeting long-chain hydrocarbons like C30+ waxes. reaction is evaluated. Finally, our model includes in the analysis two

2
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

distillation towers: the first needed to separate waxes, the second to case of the RWGS solution, a low-temperature electrolysis unit was also
distillate the lighter FT products. With this work, we seek to maximize inserted to deliver the required hydrogen for the process. The resulting
the synthesis of middle distillates and waxes fractions, maximize the syngas was dehydrated and supplied to a Fischer-Tropsch reactor for
CO2 conversion and minimize the thermal requirements of the process. the syncrude synthesis. The syncrude was further upgraded into several
Each system is thermally integrated allowing a sensible reduction in the chemical classes, based on their boiling point: gas, naphtha, middle
thermal energy needs of the highly energy-intensive MEA process. distillates, light waxes and heavy waxes. To increase the targeted
Lastly, considerations about the technology readiness level of each Fisher-Tropsch products yield, each configuration also accounted for a
device are given. recirculation solution of the Fischer-Tropsch off-gases back to the
syngas generation unit inlet. The conditions at 0% recirculation rate
(0% RR) of the off-gases were considered as the reference. Finally, the
2. Methodology
effect of gas pressurization before the RWGS and the SOEC was ad-
ditionally included in the analysis. The model was developed on the
2.1. Plant layout
commercial tool for mass and energy balances AspenPLUS™. Fig. 1
depicts schematics of analysed processes.
A biogas upgrading unit based on chemical scrubbing and separ-
ating CO2 from CH4 was chosen as the source for carbon dioxide. Two
different process configurations were then studied, with variation in the 2.2. Biogas upgrading and CO2 capture unit
syngas generation unit. The recovered CO2 from the biogas upgrading
section was fed to either an RWGS reactor or a SOEC stack operating The first unit (common to all the plant configurations) consisted of a
under co-electrolysis conditions for the synthesis gas generation. In the medium-size biogas-upgrading system, which disposed of 1680 m3/hr

Fig. 1. Plant layouts from biogas inlet to FT products separation: top) Case A with RWGS reactor for syngas generation; bottom) Case B with SOEC stack for syngas
generation.

3
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

of clean biogas, with 34 mol-% CO2, 65 mol-% CH4 and 1 mol-% N2, Table 1
resulting in 1 ton/hr of CO2 captured [28]. The solvent selected for the Characteristic of the biogas and solvent fed to the CO2 separation unit.
absorption process was monoethanolamine (MEA). A detailed layout of Biogas Inlet Composition [%mol]
this biogas upgrading process is presented in Fig. 2. The absorber was
modelled as a rate-based multistage packing column, where the solvent
3
Flow rate [m /hr] 1680 CO2 0.34
T [K] 298 CH4 0.65
amine selectively bound with the CO2 in the biogas stream via an
p [bar] 1 N2 0.01
exothermic reaction. The stripper was a rate-based multistage packing
column, where the release of the CO2 took place by means of heat Lean Solvent Composition [%mol]

transfer through the reboiler bottom stage. The make-up section al- Flow rate [kg/s] 6.50 MEA 0.11
lowed for the reintegration of MEA and water lost during the process of T [K] 313 H2O 0.864
separation. The cross heat exchanger preheated the CO2-rich stream p [bar] 1.07 CO2 0.026
entering the stripper while precooling the one exiting the stripper. A
minimum temperature difference of 15 K between hot and cold stream
was set in this component. The recovered CH4 was compressed to 2H2 O H3 O+ + OH (1)
10 bar, dried and injected into the local gas grid, whereas the carbon CO2 + 2H2 O HCO3 + H3O+ (2)
dioxide was dried and sent to the syngas generator (Table 1).
The two columns required the definition of the loading capacity, the HCO3 + H2 O CO3 2 + H3 O+ (3)
number of stages as well as the insertion of the reaction mechanism to
simulate the CO2 capture process. The loading, i.e., the mole ratio of MEAH+ + H2 O MEA + HCO3 (4)
CO2 over the mole of amine, was set to 0.24, similarly to the works of MEACOO + H2 O MEA + HCO3 (5)
Raynal et al. [29] and Li et al. [30]: this value was used to evaluate the
lean solvent composition after CO2 separation. The columns were de- ln (Ki) = Ai + Bi /T + Ci ln(T) + Di T ( )
(6)
signed to reduce the thermal needs at the stripper reboiler and to
capture 98% of CO2 entering the system with the biogas flow. Both MEA + CO2 + H2 O MEACOO + H3 O+ (7)
columns presented 14 stages, with well-mixed flow both in the liquid MEACOO + H3 O+ MEA + CO2 + H2 O (8)
and vapour phases. Moreover, the stripper presented a kettle reboiler at
the bottom stage and a condenser at the top one, with a distillate-to- CO2 + OH HCO3 (9)
feed flow ratio of 0.04. Specific components information is listed in
section Appendix A. HCO3 CO2 + OH (10)
The solvent considered in this study was a mixture of 30 wt-% MEA ki = Aiexp( Eacti/RTi )( )
(11)
and 70 wt-% water. The solvent selection was related to its high se-
lectivity towards CO2, together with its widespread industrial applica- (*) Kinetic coefficients values are available in Appendix A.
tions despite corrosion and volatility issues. Furthermore, even though
MEA-based CO2 separation is highly energy-intensive, proper thermal 2.3. Syngas generation
integration allows reducing such needs. A carbamate formation me-
chanism was considered to simulate the process of carbon capture with In this study, a reverse water gas shift reactor and a solid oxide
specific kinetic steps applied as the one proposed by Moioli et al. [31]. electrolysis stack operating in co-electrolysis mode were selected as
Lastly, for this section, the Electrolyte-NRTL (Non-Random Two Liquid) possible ways to generate a synthesis gas.
and the Soave-Redlich-Kwong (SRK) state equations described the
thermodynamics of the Vapour-Liquid-Equilibrium for the liquid and 2.3.1. Reverse water gas shift reactor
vapour phase, respectively. The captured carbon dioxide was catalytically converted into

Fig. 2. Biogas upgrading unit schematics.

4
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

carbon monoxide inside the RWGS reactor according to: 2.3.2. Solid oxide electrolysis cell
As one possible alternative, syngas could be produced through a
CO2 + H2 CO + H2 O H0298K = 41 kJ/mol (12) solid oxide electrolysis stack (SOEC) operating under co-electrolysis
As demonstrated by Vidal Vazquez et al. [32], the RWGS can reach conditions (i.e., simultaneous splitting of H2O and CO2). The layout of
equilibrium conditions with the proper catalyst selection and the cor- this section is shown in Fig. 3.
rect reactor size (i.e. catalyst loading, reactor length and diameter). As The SOEC system operated at 1073 K and atmospheric pressure to
such, the RWGS reactor employed in this analysis was modelled at match the conditions of the RWGS. Solid oxide cells usually involve as
equilibrium conditions, minimizing the Gibbs free energy, and loaded state-of-the-art materials Ni/YSZ cathode, YSZ electrolyte, LSC/CGO
with a Ni/Al2O3 catalyst. It operated at 1073 K and ambient pressure, anode [35]. The electrochemical reactions for the conversion of CO2-to-
with an inlet gas flow preheated up to 833 K. At equilibrium conditions, CO and H2O-to-H2 followed this presented scheme at cathode and
also side reactions of methanation and carbon deposition through the anode sides:
Boudouard reaction were taken into account [32]: Cathode Side H2 O+ 2 e H2 + O2 (19)

CO + 3 H2 CH 4 + H2 O H0298K = 165kJ/mol (13) Cathode Side CO2 + 2 e CO + O2 (20)

CO2 + 4 H2 CH 4 + 2 H2 O 0
H298K = 206kJ/mol (14) Anode Side O2 0.5 O2 + 2 e (21)
Chemical reactions taking place at the cathode side were also con-
2 CO CO2 + C(s) H0298K = 172 kJ/mol (15) sidered. Due to the simultaneous presence of H2O, CO, CO2 and H2 at
the high operating temperature and reactant utilization, Ni also acts as
CO + H2 C(s) + H2 O H0298K = 131 kJ/mol (16) catalyst for the reactions of water gas shift (Eq. (12)) and methane
reforming (Eqs. (13) and (14)), which may take place at chemical
Recirculation of the Fischer-Tropsch off-gas to the RWGS reactor equilibrium conditions [36]. Also the Boudouard reaction (Eq. (15))
inlet was studied, too. Therefore, the equilibrium model considered also may represent a possible drawback, as it can lead to solid carbon de-
the side reactions of hydrocarbons decomposition and steam reforming position at the Ni cathode matrix, blocking the three-phase points of
of low molecular weight compounds produced during the Fischer- electrochemical reaction at too low or high reactant utilization [37].
Tropsch reaction entering the RWGS reactor [33]: Finally, recirculation of the Fisher-Tropsch off-gases could produce
cracking of light hydrocarbons at the stack level (Eqs. (15)–(18)).
Cn Hm n C+ (m/2) H2 e. g. H0298K = 791 kJ/mol CH 4 ; 5960 kJ
In this investigation, the electrochemical conversion of the reactants
/mol C8 H18 (17) was modelled using a stoichiometric reactor followed by an O2 se-
parator, as depicted in Fig. 3. The reactions taking place at equilibrium
Cn Hm + nH2 O n CO + (n+m/2)H2 e. g. H 0298K in the fuel electrode were inserted in the model using a reactor at the
= 206 kJ/mol CH 4 ; 1277 kJ/mol C8 H18 (18) equilibrium condition after the O2 removal (EQ-2). As demonstrated in
other studies, the conversion of CO2-to-CO could be accounted via the
All the captured carbon dioxide from the biogas upgrading unit was WGS reaction in addition to the electrochemical conversion [38]. Fi-
exploited and the desired H2/CO molar ratio of 1.9 was obtained by nally, in the case of RR values higher than 0%, unconverted CO2, CO,
adjusting the inlet water to the electrolysis unit. The alkaline electro- H2 and light hydrocarbons entered the SOEC unit, requiring the simu-
lyser delivered H2 at 1 bar and 293 K, with water consumption of 15 L/ lation of equilibrium reactions also before the O2 separator (EQ-1).
kgH2 and 51 kWh/kgH2 [34]. Moreover, steam could be additionally fed Even in this system configuration with the SOEC module, we as-
at the inlet of the RWGS reactor, if needed to shift the carbon–oxygen- sumed to exploit the full flow of CO2 exiting the biogas upgrading unit.
hydrogen equilibrium and avoid solid carbon formation. Lastly, the The targeted H2/CO molar ratio in the syngas was achieved by variation
RWGS applied the SRK equation of state. of the steam flow rate fed to the SOEC. Moreover, a syngas fraction was

Fig. 3. SOEC stack unit schematics.

5
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

recirculated back to the cathode inlet to obtain at least 10 mol-% of H2 FT general CO + 2 H2 (CH2) + H2 O h0298K = 165 kJ/mol
when feeding the SOEC fuel electrode to avoid nickel oxidation [37]. (29)
Finally, the effect of FT off-gases recirculation (RR) in the range be-
tween 0% and 90% was studied, with special attention to avoid carbon Paraffin: nCO + (2n + 1)H2 Cn H2n+ 2 + n H2 O (30)
deposition. As such, a reactants utilization (RU) of 75% at stack level Olefin: nCO + 2n H2 Cn H2n + n H2 O (31)
was set in the case of ambient pressure operation.
The Faraday’s law for co-electrolysis was expressed referring it to In this study, only paraffins and olefins were considered relevant
the O2 component to directly include the conversion of both CO2 and products, based on previous experimental results on the selected cata-
H2O. lyst where negligible amounts of side FT compounds were detected
[27]. Moreover, the reactants conversion and the generation of each
j Atot I product were modelled using a rate-based approach. The used kinetic
nO2 = =
n e,tot F n e,tot F (22) model allowed describing the synthesis of each FT hydrocarbon, de-
fining the product rates based on the syngas partial pressure, tem-
Being ne,tot the number of total electrons exchanged equal to 4, F the
perature and probability growth of the compounds, and to account for
Faraday constant (96485 C/mol). Furthermore, the SOEC was assumed
the ASF deviations. The detailed derivation of the kinetic model em-
to operate at thermoneutral condition, i.e., the thermal energy pro-
ployed in this analysis established on experimental results on a Co-
duced by the ohmic effect equalled the heat linked to electrochemical
based FT catalyst can be found in [27]. Here we present only the final
and chemical reactions: no external heat was required and the SOEC
formulation of the product rates.
system was thermally self-sustained [39].
g R CH4 = k 6M 1 (K 0PH2 ) PH2 [vac] (32)
VOCV =
I (23)
R C2 H4 = k7Ee2c 1 2 (K 0 PH2 ) [vac] (33)
h
VTN = n
I (24) R CnH2n +2 = k 6 (K 0 PH2 ) PH2 [vac]; for n 2
1 2 n
n=3 (34)
VTN VOCV
jTN =
ASR (25) n
R CnH2n = k7e nc 1 2 n (K 0PH2 ) [vac]; for n 3
I n =3 (35)
Atot =
jTN (26) * The definition αi-th and [vac] can be found in Appendix B.
The kinetic model is based on the carbide mechanism for the de-
The Area-Specific Resistance (ASR) accounting for all the internal
scription of the FT reaction, and both reactants and products rates
losses was set equal to 0.25 Ωcm2 at ambient pressure, and 0.2 Ωcm2 at
follow a LHHW formulation. Kith and kith are the rate and equilibrium
high-pressure conditions [40,41]. Once the thermoneutral operations
constants, respectively; αith the probability growth at each carbon
were ensured, the total electric power to be fed to the SOEC stack could
number; p the partial pressure of hydrogen and carbon monoxide. A
be evaluated.
plug flow reactor working at isothermal conditions was selected, with
Wel = VTNI (27) the application of an external kinetic subroutine for rates definition up
to C80 for paraffins and C40 for olefins (the detailed information about
The reactant utilization referred to the mole flow rate of reactants
the kinetic external subroutine implementation can be found in the
exiting the stack unit with respect to the feeding, accounting for both
supplementary material file).
electrochemical conversion and for reactions at equilibrium conditions.
From the molar balance of carbon monoxide between the reactor
The fractional conversion of electrochemical reactions was iteratively
inlet and outlet, it was possible to evaluate the CO conversion. This was
adjusted in the model to reach the targeted value.
set in all cases to a fixed per-pass value of XCO = 75% under no FT off-
n reactin n react out gases recirculation condition [44]. This value was selected to suppress
RU =
n reactin (28) both the water gas shift reaction over cobalt (relevant at XCO > 80%),
as well as to reduce the formation of methane and light hydrocarbons,
Lastly, the SRK equation of state was used for modelling this com- which were unwanted products compared to targeted long-chain hy-
ponent. Involved compounds are present as vapour phase only. drocarbons [45]. Finally, 35% catalyst bed porosity was assumed in the
reactor model.
2.4. Fischer-Tropsch synthesis unit nCOinFT nCOoutFT
Xco =
nCOinFT (36)
The Fischer-Tropsch reactor was modelled as a multitubular fixed-
bed reactor, operating at 501 K and 25 bar [42]. The operating pressure The thermodynamic model applied in the FT unit was the RKS
was reached via a 3-compressors block with intercooling. The involved equation of state, with Boston-Mathias alpha value modification (SRK-
catalyst was a Co-based one with alumina support, suitable for long- BM), accounting for hydrocarbon and petrochemical problems for non-
chain hydrocarbon synthesis (Co-Pt/γ-Al2O3) [27]. The reactor was polar mixtures, such as in the FT-synthesis. Petrochemical compounds
designed as one-through configuration, where no recirculation to its with carbon number higher than C30 were inserted inside the
inlet was considered. AspenPLUS™ environment from external databases.
In the general FT synthesis theory, the two main products are n-
paraffins and α-olefins, with possible formation of aromatic compounds 2.5. FT products distillation
and alcohols. Moreover, the products leaving the reactor follow a
modification of the Anderson-Flory-Shultz (AFS) distribution, where a The separation of the FT products consisted of three consecutive
higher methane formation and a lower ethane formation with respect to steps. A first three-phase separator operating at constant ambient
the ASF distribution are observed. Lastly, the FT product distribution temperature divided water from the gas and liquid FT fractions. The
has an exponential behaviour up to carbon number C15-20, after which a liquid compounds underwent a first distillation process separating
linear distribution of the products could be expected [43]. The FT re- naphtha from middle distillates, wax and residual gas fractions. A
action follows the scheme: second distillation tower provided the separation of light waxes from

6
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

heavy waxes. Both distillation towers were modelled with reference to 3. Results
the boiling temperature of each fraction cut operating at ambient
pressure. To evaluate the final tower size, a first estimation with the Both syngas generation options (SOEC and RWGS) account for two
Winn-Underwood-Gilliland method allowed finding the theoretical different solutions, without FT off-gas recirculation (0% RR) and 90%
minimum number of stages and minimum reflux ratio at ambient of the FT off-gas recirculation (90% RR). Moreover, the hypothesis
pressure required to separate each fraction [46]. Consequently, the considered for the comparison of the models are the following:
distillation towers were substituted and modelled with a more rigorous
PetroFrac component in AspenPLUS™, which accounted for the pre- • Use of all the captured CO from the biogas upgrading unit: 1 ton/
2
sence of kettle reboilers at the bottom stage, a condenser at the top hr;
stage and side strippers. The first distillation tower had a total number • Fix FT reactor size obtained at 75% CO conversion and 0% RR;
of 54 stages, with a reflux ratio of 1.8. It had a side stripper of 16 stages, • Avoid carbon deposition at the syngas generation unit;
with draw at stage 18 and overhead return at stage 12 of the main • Maximization of the CRP;
column. It presented a preheating up to 523 K at the inlet of the tower. • Maximization of the middle distillate and wax fraction throughput;
The second distillation column was composed of 58 stages and a reflux • Minimization of the energy demand.
ratio of 1.8. The tower separated the light wax fraction from the heavy
fraction. The FT products entering the second distillation tower were 3.1. CO2 separation unit
preheated to 728 K [46,47]. Hydrocarbons cuts temperatures selected
for each of the fractions are presented in Table 2. The biogas upgrading section is common to each different plant
Off-gases leaving the first distillation tower were mixed with the off- configuration. This section provides a capture rate of 98% of the total
gases coming from the 3-phase separator. The fraction not recirculated CO2 injected into the plant and 99.9% recovery of the CH4. The CO2
back to the inlet of the syngas generator was burnt. The burner oper- purity is about 98%, while the separated stream has a methane purity of
ated at atmospheric pressure, with air-excess conditions. The airflow 98.5%. Both values are in line with similar applications of biogas up-
needed for the combustion was iteratively changed to reach a com- grading with chemical absorption solutions [50–52]. Moreover, the
bustion temperature of 1273 K [48]. specification of the CH4 stream makes it suitable for a possible injection
in the gas grid, having a Wobbe Index of 50.6 MJ/Sm3 and a high
heating value of 37.9 MJ/Sm3 [53]. On the other hand, the great system
2.6. Energy integration and system efficiency evaluation
separation performance results in high specific heat duty for the solvent
regeneration at the stripper reboiler stage of 4.75 GJth/tonCO2, in good
Energy integration was carried out to minimize the external heat
agreement with literature values for MEA-based plants (3.5–5 GJ/
requirement. Pinch analysis methodology was applied to all thermal
tonCO2) [54,55]. Such value would result in high heat consumption for a
streams of each plant configuration, matching cold with hot fluids (the
biogas upgrading system. However, when combined with a power-to-
first being those streams that increase their temperature, the latter the
liquid technology with proper energy integration, such external thermal
ones in need of refrigeration). Both hot and cold composite curves were
needs can be reduced or avoided. Exact outlet composition and specifics
built considering a minimum difference of temperature at the pinch
of the main components are presented in Table 3.
point of 288 K. Finally, plant efficiencies and key performance in-
The results listed in Table 3 show that the purity of the CO2 stream
dicators were selected. In the biogas upgrading section, the specific
is only counteracted by the presence of inert N2 gas, while the amount
reboiler heat duty GJ/tonCO2, the CO2 capture rate and stream purity
of CH4 leaving the biogas upgrading unit from the carbon dioxide side is
and the section efficiency (ηBiogas) were considered.
negligible. This becomes very important when coupling such a device
nCH4 LHVCH4 with a Fischer-Tropsch reactor, where the presence of CH4 may inhibit
Biogas = the formation of high molecular weight hydrocarbons. Moreover, a
Welbiogas + Q net biogas + nbiogas LHVbiogas (37)
reintegration of the MEA and water looping in the system is required.
Additionally, the efficiency of the syngas generation and FT reaction This loss is connected to phenomena of entrainment, volatilization and
were evaluated: degradation in the columns and heat exchangers mainly [30,56]. Lastly,
the cross heat exchanger allows having an internal heat transfer on
FTsyncrude
ni LHVi 860 kW from the stripping column to the outlet flow of absorption
Syngas + FTR = column (Fig. 2).
WelSyngas +FTR + QnetSyngas + FTR (38)

Finally, the global system efficiency (ηGlob) and the overall carbon 3.2. FT product synthesis
reduction potential were estimated:
In all cases, a reference configuration can be selected at H2/CO
nCH4 LHVCH4 + FTsyncrude
ni LHVi molar ratio of 1.9, with a recirculation rate of the Fischer-Tropsch off-
Glob = gases at 0%.
Welnet + Qnet + nbiogas LHVbiogas (39)

3.2.1. Baseline configuration: 0% RR


nCO2Biogas nCO2Exhaust
CRP = Carbon dioxide is sent from the biogas upgrading unit to the syngas
nCO2Biogas (40) generator. Water is fed to the alkaline electrolyser and delivered in the

The total system efficiency ηGlob considered the whole process, from Table 2
the CO2 separation in the biogas-upgrading unit to the ventilation to- Different product classes obtained after the distillation of the FT products.
wards the atmosphere of the exhaust gases. To account for the energy
Fraction Carbon Number Boiling Temperature [K]
content of the FT products in [kJ/mol], a correlation for paraffins and
olefins derived by Stempien et al. [49] was used: Gas C1-C4 313
Naphtha C4-C11 396
LHVPar. = 608.44n + 213.31 (41) Middle Distillate C11-C20 498
Light Waxes C20-C35 656
Heavy Waxes C35-C70+ 793
LHVOlef. = 604.93n + 113.83 (42)

7
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table 3 by the biogas upgrading section, and 32.4% of the total plant electricity
Composition of the separated CH4 stream and CO2 stream and final parameters in case A and 29.5% in case B. Lastly, for both cases A and B the 3-
of the components obtained after the simulation. compressors unit needed before the FT reactor accounts for about 4% of
CO2 stream [mol/s] Composition [% mol] the total electrical needs. In this component, the higher electrical needs
Mole flow rate 18.8 CO2 97.96 of case B is again related to the marginally different gas composition
CH4 0.02 entering the FT reactor, affecting the compression (Table 6 provides the
H2O 2.03
composition of the gas entering the FT reactor).
N2 0.0
CO2 separation [CO2out/ 98.4% Considering the technical specifics of the main components, the
efficiency CO2in] Fischer-Tropsch reactor results in 19.25 m3 and 18.06 m3 in case A and
CH4 stream [mol/s] Composition [% mol]
B, respectively. Moreover, case A has an RWGS reactor that requires
309.4 kW to operate at isothermal conditions. Lastly, the SOEC con-
Mole flow rate 12.5 CO2 0.02 sumes 3.6 MWel at thermoneutral condition, corresponding to a voltage
CH4 98.47 of 1.35 V (OCV at 0.96 V) and a current density of 0.58 A/cm2, leading
H2O 0.06
to a total required area of 455.7 m2.
N2 1.51
Wobbe Index [MJ/Sm3] 50.56
HHV [MJ/Sm3] 37.88
3.2.2. Upgrading of the systems: 90% RR
CH4 separation [CH4out/ 99.9%
efficiency CH4in] The recirculation of the FT off-gases enables increasing the
throughput of useful products and providing a rise in the total con-
Components
version of CO2. A value of 90% recirculation rate has been selected as a
Stripper reboiler [kWth] 1313.2 trade-off between the increase in useful FT material, conversion of
Stripper condenser [kWth] −227.1 carbon dioxide and rise in the energy demand due to the higher amount
Stripper specific heat [GJth/ 4.75 of FT products that have to be processed by the distillation columns. In
duty tonCO2]
Cross Heat Exchanger [kWth] 860.6
addition, higher RR could result in accumulation of inert gases in the
duty system lines [59].
Biogas blower [kWel] 7.44 In both the configurations, the recirculation is exponentially bene-
Circulation pump [kWel] 0.69 ficial for the synthesis of the targeted waxes, middle distillates and
CH4 Compressor [kWel] 103.3
naphtha, whereas it reduces the generation of methane and light hy-
Make-up flow [mol/s] 17.87 MEA 2.5
H2O 97.5 drocarbons thanks to their cracking inside the syngas units. Specifically,
case A provides a slightly higher amount of FT syncrude than case B.
Possibly, the high conversion of CO2 in the SOEC results in a high
form of H2 in case A, or to the SOEC unit after vaporization as steam in amount of CO processed by the FT reactor and transformed into gas
case B. Table 4 provides a summary of the most relevant information of fraction. Therefore, the iterative recirculation of the off-gases of case B
the mass balances of the reference configuration at 0% RR. determines a consequently higher amount of material fed to the burner
Considering the synthesis of useful FT products, the configuration and not recycled compared to case A (Table 7).
with the solid oxide electrolyser allows reaching a greater production. In case A, the endothermicity of the RWGS reactor requires 1.1
This is related to the chemical and technical limitations associated with MWth to operate at isothermal conditions. With a similar demand rise
the RWGS reactor, which reaches a maximum carbon dioxide conver- with respect to the 0% RR case, the SOEC needs the supply of 5.3 MWel
sion lower than the one obtained with the SOEC (63% and 75% for case at thermoneutral conditions, reaching 1.50 V (OCV at 0.94 V), a ther-
A and case B, respectively). This directly impacts the synthesis of hy- moneutral current of 0.84 A/cm2 and a total area of 418.8 m2.
drocarbons in the Fischer-Tropsch reactor, given the lower CO avail- In terms of global process efficiency, each configuration presents a
ability for further processing. CO2 might act as inert gas under the FT similar value to the cases at 0% recirculation rate, with slightly lower
reaction, providing no useful reactant for the synthesis of the hydro- values at 90% RR. This is due to a rise in the energy demand (thermal
carbons [57]. Consequently, a lower amount of air to the combustor is and electrical) that counterbalances the beneficial increase of useful FT
required for case A. It is worth to notice that the reached fractional products. Specifically, a great supply of energy is required by the RWGS
conversion of CO2 in the reactor is confirmed by other studies that reactor to operate at 1073 K, by the SOEC steam generator and by the
report RWGS applications [24,58]. Moreover, the RU of the SOEC unit distillation towers processing the FT products. As expected from the
could be additionally increased to 80–85%, determining a further rise mass balance, in the latter case of separation columns, a slightly higher
in the CO2 consumption rate. However, the value has been kept to a amount of energy is required by case A (98.2 kWth against 96.7 kWth).
maximum of 75% to avoid possible solid carbon deposits. Concerning On the other hand, the recirculation increases both the local and global
the global CRP, case B reaches a higher value than case A, given the
higher conversion contribution of the syngas unit. Table 4
In terms of thermal and electrical energy consumptions and pro- Case A and Case B 0% RR mass balance.
cesses efficiencies, the results are listed in Table 5. The thermal needs of Mass Flow Rate [kg/h] Case A Case B
the system with SOEC stack are higher. Firstly, the generation of steam Biogas Inlet 1745 1745
is highly energy-intensive. Secondly, the higher amount of FT products Captured CO2 1002.8 1002.8
needs more heat to process the separation in the distillation towers Biogas Up. CH4 728 728
H2O for electrolysis 1297.4 774.1
(case A 39.2 kWth, case B 44.9 kWth). On the contrary, the electricity Naphtha C5-11 26.1 31.4
consumed by the alkaline electrolyser is higher than the one absorbed Middle distillate C11-20 31.3 37.0
by the SOEC, impacting more on the global efficiency ηGlob described by Light wax C20-35 20.9 24.9
Eq. (48). Moreover, in both cases, the most energivorous process is Heavy wax C35+ 9.9 11.8
Steam to avoid C-deposition 0.0 0.0
related to the separation of CO2 inside the stripper of the biogas up-
Total condensed water 809.0 761.7
grading unit. The electrical consumption of the different sections is also Combustion air 5022.1 5812.1
provided. In the case of the biogas upgrading unit only, the most energy Exhaust gas 5549.0 6266.3
demanding element is the compressor of the biomethane for grid in- CRP 27.5% 34.2%
jection (103.3 kWel). This corresponds to 93% of the electricity required CO2 Conversion at Syngas Section 63.0% 75.0%

8
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table 5 Table 7
Energy consumptions and process efficiencies at 0% RR. a Complete informa- Mass and energy balance of case A and case B at 90% RR. a SOEC working at the
tion of the thermal balance is provided in Appendix C. b SOEC working at the thermoneutral condition: no heat exchange with the environment is required. b
thermoneutral condition: no heat exchange with the environment is required. Complete information of the thermal balance is provided in Appendix C.
Thermal balance [kWth]a Case A Case B Mass Balance [kg/h] Case A Case B
Biogas Inlet 1745 1745
Total Heating Needs 2432.9 2659.5 Captured CO2 1002.8 1002.8
Biogas Up. CH4 728 728
Biogas Section 1346.6 1346.6 H2O for electrolysis 2002.0 1274.9
Steam Generation – 755.9 Steam to avoid C-deposition 0.0 0.0
RWGS/SOEC 309.4 -b Total Condensed Water 1103.1 1185.8
FT Products Distillation 39.3 44.9 Combustion Air 2349.0 2398.2
Others 737.6 512.1 Exhaust Gas 2458.7 2522.1
Cooling Needs (Towards the environment) −4678.2 −5165.8 Naphtha C5-11 79.9 77.3
Fischer-Tropsch Reactor [kJ/molCO] −167.4 −168.7 Middle distillate C11-20 93.7 93.4
Fischer-Tropsch Reactor Volume [m3] 19.25 18.06 Light wax C20-35 61.1 59.3
Heavy wax C35+ 24.9 23.7
Electrical balance [kWel]
Thermal Balance [kWth]a
Biogas Upgrading Section 111.5 111.5
Syngas + FT Section 207.8 236.1 Heating Need 3342.1 3639.1
B2 1.7 – Biogas Section 1346.6 1346.6
B3 61.9 67.4 Steam Generation – 1245.1
B4 69.3 84.2 RWGS/SOEC 1089.7 –b
B5 69.5 84.5 FT Products Distillation 98.2 96.7
B-Cdep – – Recirculation Heating 233.3 394.3
Electrolysis 4341.1 3569.4 Others 574.3 556.4
Cooling Need (Towards to environment) −5734.6 −6254.3
Energy content [kW]
Fischer-Tropsch Reactor [kJ/molCO] −169.8 −169.3
Fischer-Tropsch Reactor Volume [m3] 19.25 18.06
Biogas Feed 9868.5 9868.5
FT Products −1079.3 −1285.5 Electric Balance [kWel]
Biogas Up. CH4 −9867.3 −9867.3
Biogas Upgrading Section 111.5 111.5
Process Efficiencies
Syngas + FT Section 540.9 613.9
B2 2.6 –
ηBiogas 87.2% 87.2%
B3 168.7 174.9
ηSyngas+FT 19.2% 25.2%
B4 188.8 218.8
ηGlob 64.5% 67.8%
B5 179.6 219.6
B-Cdep – –
Electrolysis 6698.7 5280.9

Table 6 Energy content [kW]


Gas molar composition at the outlet of the RWGS (Case A) and SOEC (Case B).
Biogas Feed 9868.5 9868.5
[% mol] Case A 0% RR Case B 0% RR Case A 90% RR Case B 90% RR FT Products −3168.6 −3110.5
Biogas Up. CH4 −9867.3 −9867.3
CO 22.5% 25.8% 30.1% 28.9%
Process Efficiencies
H2 42.7% 49.1% 57.3% 54.9%
H2O 22.9% 16.5% 8.0% 10.5%
CRP 81.1% 79.6%
CH4 0.0% 0.1% 0.4% 0.3%
CO2 Conversion at Syngas Section 97.3% 97.3%
CO2 11.8% 8.5% 4.2% 5.4%
ηBiogas 87.2% 87.2%
ηSyngas+FT 34.4% 38.2%
ηGlob 63.4% 66.5%
conversion of carbon dioxide: a greater material flow is processed in-
side the RWGS or SOC systems. The effect of recirculation is depicted in
Fig. 4: the recirculation of the FT off-gases could be considered bene- and B at increasing recirculation rates at the outlet of the RWGS and
ficial. SOEC units. As both cases at ambient pressure do not reach any solid
carbon formation conditions, no additional steam is required by the
RWGS reactor nor a reduced RU in the SOC stack. Lastly, no carbon
3.2.3. Carbon deposition formation has been detected experimentally nor with modelling in the
Solid carbon formation can lead to catalyst deactivation, reducing Fisher-Tropsch reactor.
the performance of the entire process. It has been experimentally de-
tected both on RWGS and SOEC technologies [60,61]. To avoid carbon
deposition, different solutions can be applied. With RWGS reactors, a 3.3. Energy integration
lower reactant conversion or an increase in the steam-to-carbon ratio
can be carried out [60,62]. Similarly, coke deposition onto the SOC The optimal energy integration of the thermal streams can improve
stack layers can be avoided with a variation of the reactants utilization, system performance by reducing the external thermal need. Given the
addition of steam or modification of the electrode with dopants de- higher throughput of products that can be obtained in the 90% RR
position [36,63]. Moreover, with an increase of the recirculation rate, configurations, the energy integration results are presented for these
solid carbon formation could be stronger due to the higher presence of cases only, in this work.
carbon compounds dragged into the syngas unit generators with the Table 8 lists the main results derived from the energy integration of
recirculated stream, with a higher contribution of Eqs. (15)–(18). the streams, while the composite curves of the two cases are shown in
However, thermodynamic calculations show that no solid carbon de- Fig. 6. In terms of hot composite curves, cases A and B have a similar
posits at the RWGS nor SOEC systems at the given conditions. Fig. 5 shape. The majority of the thermal power at a high temperature can be
provides the evolution of the thermodynamic equilibrium of cases A recovered from the cooling of the exhaust gases vented towards the

9
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Fig. 4. Effect of the increased in the recirculation rate: top) FT products; bottom-left) Gas fraction; bottom-right) carbon reduction potential.

environment: case A 747 kWth, case B 760 kWth. Moreover, both con- for the vaporization of the water entering the stack themselves, and at
figurations present a plateau at 501 K, corresponding to the heat re- high temperature for the RWGS to keep the reactor at the operating
leased by the exothermic activity of the Fischer-Tropsch reactor, ac- temperature. As such, the presence of high temperature heat demand
counting for 25% and 23% of the total available heat of the systems for related to the RWGS determines an increased external energy con-
case A and B, respectively. Considering the cold composite curves, two sumption in case A, with a lowering of the global systems efficiency
plateaus can be identified for every configuration. The first is related to with respect to case B. As summarized in Table 8, the final thermal
the heat required by the reboiler of the biogas upgrading unit at 388 K. needs of the RWGS system correspond to 0.9 MWth. This could be
A second one can be identified at low temperature for the SOEC stack supplied by high-temperature electric heaters. In the case of the SOEC

Fig. 5. Ternary diagrams of solid carbon formation at increasing value of RR: left) Case A, outlet of the RWGS reactor; right) Case B, outlet of the SOE electrolyser.

10
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table 8
Effect of the energy integration of the total energy consumption and plant efficiencies at 90% RR.
Pinch Point Temperature [K] Case A Case B

1080.7 1265.7

Before Integration After Integration Before Integration After Integration

Total heating Demand [kWth] 3340.2 1000.5 3639.1 0.0


Total cooling demand [kWth] −5734.6 −3358.1 −6254.3 −2615.3
ηGlob 63.4% 71.5% 66.5% 81.8%

stack, all the thermal needs can be covered with internal system heating RR. The second-highest efficiency is reached at similar condition, but
generation. Under configuration A, heat recovery from internal flows employing the integrated RWGS in place of the SOEC stack (71.5%). On
corresponds to 70.1% of the total heat required by the system upon the one hand, the very high values can be related to the energy content
integration, whereas configuration B reaches 100% recovery, given the of the recovered methane in the biogas upgrading unit. On the other
lower demands of the SOEC against the RWGS at high temperature. For hand, the ideal integration of the stream sensibly reduces the energy
instance, referring to the most impacting heat sources and sinks, above needs coming from the outside. With respect to similar studies where
the pinch point of configuration A 145 kW can be transferred from the CO2 capture is inserted in the plant process, Vidal et al. [24] achieved a
exhausts cooling to the RWGS reactor, to partially sustain its en- system efficiency as high as 47% when coupling an FT reactor with an
dothermicity activity. Below the pinch point, the same heat source RWGS with direct air capture. With biomass gasification, Leibbrandt
transfers 233 kW to the recirculated FT gas fraction. The syngas cooling et al. [64] reached a value of 51%. Monaco et al. [65] calculated effi-
for water condensation can supply 460 kW to preheat the mixture en- ciency of 65.9% and 62% with CO2 recovered from biomass upgrade
tering the RWGS reactor and the FT products entering the first dis- and biogas upgrade, respectively. As such, the present model represents
tillation columns and to the reboiler of the column. Similarly, 80 kW are a highly efficient solution to combine a CO2 capture system with FT
transferred to the preheater of the FT material entering the second synthetic chemicals production.
distillation column and to the same column reboiler. The exothermicity Samavati et al. [66] and Cinti et al. [19] obtained 85% and 57.2%
of the FT can be exploited in the biogas upgrading unit stripper plant efficiency, operating with a solution SOEC + FT. Selvatico et al.
(1313 kW) and to preheat both the biogas and the make-up flow en- [21] of 43.7% with an RWGS + FT solution. In this study, considering
tering this unit. Similar considerations can be given in configuration B, only the section of CO2 upgrade to FT products, the highest ηSyngas+FTR
where all the heat exchanges take place below the pinch point tem- efficiencies correspond to 38.3% and 34.4% for a recirculation rate of
perature. The cooling of the exhaust gases can heat the recirculated gas 90% for case B and A, respectively. However, by optimally integrating
fraction, the SOEC sweep air and preheat the water up to zero vapour only the section after the CO2 separation, the efficiency ηSyngas+FTR
fraction (541 kW). The cooling of the syngas can preheat the mixture would reach a value of 38.7% for case A and of 52.7% for case B,
entering the SOEC, sustain the reboiler requirements for the distillation making this plant configuration without CO2 capture still competitive
towers and provide thermal power for the biogas stripper (716 kW). with other studies from the literature.
The exothermicity of the FT can sustain the vaporization of the water
for the electrochemical reaction and part of the heat required by the 4.1. H2/CO molar ratio effect
biogas stripper (1229 kW). Finally, for all the cases, excess heat is
available at low temperature. In this study, this heat has been con- A variation of the ratio of hydrogen-to-carbon monoxide at the inlet
sidered as a waste, however, it can be employed for steam generation or of the Fischer-Tropsch reactor can affect the synthesis of FT products.
low-temperature heating of local users. The Fischer-Tropsch process requires a feed gas with a molar ratio of
H2/CO around 2.0 to account for the stoichiometry of the reaction
4. Discussion taking place over the catalytic bed [67]. A rise in the H2/CO ratio re-
duces the synthesis of high molecular weight hydrocarbons for the
The overall highest plant efficiency corresponds to 81.8%. This stoichiometry of the Fischer-Tropsch synthesis, by means of lower
value can be reached with optimal energy integration for case B at 90% probability growth at Eqs. ((31)–(38)). This results in higher selectivity

Fig. 6. Composite curves derived by the pinch analysis at 90% RR: left) Case A; right) Case B.

11
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

of lighter hydrocarbons and lowers CRP values by means of more ma- production rises with an increment of the pressure up to 25 bar, de-
terial fed to the burner. Contrariwise, having lower H2/CO values termining a reduction of the CO available at the outlet of the reactors
provides αith towards unity, and thus increases the yield of wax pro- with respect to the ambient pressure case. However, this is an unwanted
duction. However, the heat duty of the distillation towers would con- product in power-to-liquid applications with FT reactors, counteracting
sequently rise. Moreover, this solution shifts the equilibrium of the FT the synthesis of long-chain hydrocarbons. Furthermore, a decrease in
reaction towards the reactants and reduces the total reaction rates and the value of CO2 conversion can be observed at increasing pressures,
CO conversion. For instance, reducing the H2/CO to 1.6 would result in possibly due to more favourable conditions for Eq. (13). Lastly, the drop
XCO of 67% in case A and of 66.5% in case B. Similarly, Campanario in the CO2 conversion for SOEC is further enhanced by the need for
et al. [48] found that reducing the H2/CO from 2.3 to 1.7 increases the reducing the reactants utilization value to avoid solid carbon formation
production of FT material but increases also the energy consumption. at pressures higher than 15 bar.
Lillebø et al. [68] experimentally demonstrated how reducing the H2/ Concerning the energy balance, pressure increase leads to a reduc-
CO ration reduces the CO conversion (Fig. 7). tion in the plants' thermal needs when operating at 0% RR. On the
contrary, at 90% RR, high-pressure configurations present a higher
material content at the recirculated stream that needs to be heated up to
4.2. Effect of pressure variation the operating temperature, thus increasing the thermal requirement of
the plants with respect to the atmospheric pressure solutions. Finally,
Given the need for high-pressure operation of the Fischer-Tropsch the configurations operating at ambient pressure provides higher plant
reactor, a solution with RWGS and SOEC at high pressure can be efficiency (global and limited to syngas + FT sections) given the same
evaluated. Mass and energy balances are summarized in Table 9. performance of the biogas upgrading unit: a lower amount of FT pro-
With reference to the synthesis of the useful FT products, it is pos- ducts at the system outlet results in a lower energy content for the
sible to note how the high-pressure configurations allow reaching a configurations at high pressure, i.e., lower plant efficiency.
lower yield of production with respect to their low-pressure counter- Furthermore, the SOC system requires the compression of a high
parts, regardless of the recirculation rate. This is related to the nature of amount of air to ensure the electrochemical conversion of the reactants,
the reactions involved in the syngas generator units, which further in- making it the least performing configuration on both the ηSyngas+FT and
fluences the performance of the Fischer-Tropsch reactor. For both the ηgl. However, the compression of the reactants before the RWGS or
RWGS reactor and the SOEC stack, Eq. (12) is equimolar, meaning that SOEC requires less electrical energy than the case of compression before
a change in pressure does not influence the equilibrium of the reaction. the FT reactor. Finally, pressure increase implies a stronger control to
However, the reactions of methanation (eq. (13)–(14)) are shifted to- avoid coking. In the case of the RWGS, this is avoided by adding steam
wards the products at high pressure by means of the Le Chatelier when operating at 90% RR (no C-deposition is detected at 0% RR and
principle [69]. This behaviour is depicted in Fig. 8, where the effect of 25 bar). In the case of the SOEC, the RU needs to be reduced from 75%
pressure is presented. For both configurations, the methane yield of

Fig. 7. Effect of H2/CO molar ratio in the syngas over: top) FT products; bottom-left) Heat duty for FT products separation; bottom-right) CRP.

12
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table 9
Mass and energy balance with syngas generators operating at high pressure (25 bar) at 0% and 90% RR.
Mass Balance [kg/h] Case A 0% RR Case B 0% RR Case A 90% RR Case B 90% RR

Biogas Inlet 1745 1745 1745 1745


Captured CO2 1002.8 1002.8 1002.8 1002.8
Biogas Up. CH4 728 728 728 728
H2O for electrolysis 1312.6 786.4 1858.1 2125.4
Steam to avoid C-deposition 0.0 0.0 840.5 0.0
Total Condensed Water 786.5 707.9 1851.8 2045.3
Combustion Air 6539.4 7697.0 4594.2 4043.9
Exhaust Gas 7126.5 9060.6 4859.9 4325.5
Naphtha C5-11 15.2 17.1 52.8 58.2
Middle distillate C11-20 18.7 17.4 72.5 67.6
Light wax C20-35 12.1 12.3 43.1 43.2
Heavy wax C35+ 5.6 5.7 16.2 16.2
Fischer-Tropsch Reactor [kJ/molCO] −167.96 −167.93 −169.07 −169.09
Fischer-Tropsch Volume [m3] 19.12 18.98 19.12 18.98

Thermal Balance [kWth]

Heating Need 2096.3 3005.2 5075.9 5593.9


Biogas Section 1346.6 1346.6 1346.6 1346.6
Steam Generation – 759.7 836.3 2050.9
RWGS/SOEC 61.2 – 1296.5 –
FT Products Distillation 22.9 23.1 71.6 75.9
Recirculation Heating – – 617.4 938.7
Others 665.5 875.8 907.4 1181.7
Cooling Needs 4842.1 5333.5 7591.1 8042.8

Electric Balance [kWel]

Biogas Upgrading Section 111.5 111.5 111.5 111.5


Syngas + FT Section 93.7 958.8 110.9 1125.5
B1-HP 69.3 69.3 69.3 69.3
B2 1.7 – 2.4
B2-HP 22.7 – 32.2 –
B3-HP – – 0.6 0.3
B4-HP – 1.8 – 4.8
B-Cdep – – 2.3 –
Air comp. SOEC – 888 – 1051.4
Electrolysis 4391.9 3400.5 6217.1 4742.9

Energy content [kW]

Biogas Feed 9868.5 9868.5 9868.5 9868.5


FT Products −631.9 −640.5 −2257.9 −2264.1
Biogas Up. CH4 −9867.3 −9867.3 −9867.3 −9867.3

Process Efficiencies

CRP 16.1% 18.1% 58.5% 58.6%


CO2 Conversion at Syngas Section 60.5% 70.0% 86.9% 84.6%
ηBiogas 87.2% 87.2% 87.2% 87.2%
ηSyngas+FT 12.1% 10.2% 22. % 22.2%
ηGlob 63.4% 60.6% 56.7% 56.6%

Fig. 8. Pressure effect on the syngas generators over CO2 conversion, CH4 yield, CO yield and H2/CO molar ratio: left) Case A; right) Case B.

13
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Fig. 9. Ternary diagrams of solid carbon formation at increasing value of RR with syngas generators working at 25 bar: left) Case A, outlet of the RWGS reactor; right)
Case B, outlet of the SOE electrolyser.

Table 10 further economic assessment will be carried out in order to identify the
Energy integration effect on the total energy consumption and plant efficiencies most proficient plant configuration from an energy and economic point
at 90% RR at high pressure operation of the syngas generators. of view.
Pinch Point Case A Case B
Temperature [K]
1080.7 493.9 4.4. On the catalyst selection
Before After Before After
Integration Integration Integration Integration In the present analysis, a Co-based catalyst was employed to syn-
thetize long-chain hydrocarbons. Even though Co is more expensive
Total Heating 5075.9 1000.5 5593.9 153.0 than Fe (the other largely employed catalyst in FTS), it shows better
Need [kWth]
activity and enhanced selectivity towards linear C5+ FTS compounds
Total Cooling Need 7591.1 3535.0 8042.8 6749.1
[kWth]
[18]. In order to make PtL routes more appealing, not only optimal
ηGlob 56.7% 70.1% 56.6% 75.8% process modelling is required, but also studying the synthesis of per-
formance-enhanced catalysts becomes a key aspect to target specific
products. Rytter et al. [72] suggested that large pores on catalyst sup-
to 70% at 0% RR and to 40% if recycling 90% of the FT off-gases ports would influence positively the selectivity towards C5+. Moreover,
(Figs. 9 and 10). the presence of structural promoters such as K, Ce, Cs, Mn, Pt can im-
prove the catalysts activity, shift the selectivity and reduce their de-
gradation. For instance, Yang et al. [45] showed how K-promotion
4.3. Technology readiness level would decrease the activity of their Fischer-Tropsch Co catalyst. How-
ever, Mn-promotion would reduce the production of methane in favour
Schmidt et al. [70] identified the technology readiness levels (TRL) of longer chain products. Gavrilovic et al. [73] determined optimal Pt
of several devices for carbon capture and utilization for P2L. Conse- effect on different-sized Co particles. However, the cost associated to
quently, the TRL of different P2L routes with Fischer-Tropsch reactor the involved rare metal con negatively affect both installation, and
has been included. They could determine that the overall TRL of a P2L operation costs [74]. Also innovative and recent catalyst synthesis
route is downgraded to the lowest TRL of the production chain. Both techniques like the atomic layer deposition (ALD) are proving to be an
the biogas upgrading with MEA and the Fischer-Tropsch reactor are attractive solution to design resilient and precise catalysts, with en-
considered highly developed systems, with TRL 9. On the contrary, the hanced performance and desired morphology. For instance, O’Neill
RWGS and SOEC devices present lower TRLs. For instance, the RWGS et al. [74] demonstrated how ALD can used to effectively control the
presents a TRL of 6, while the SOEC has been estimated at TRL ranging active sites distribution and size, and influence the catalyst selectivity.
from 3 to 5 whether it accounts for CO2 electrolysis or not [71]. Eskelinen et al. [75] provided experimental data showing the effect of
In our study, a solution employing a solid oxide electrolyser under ALD Pt-promotion on Co-catalyst and variation in light products se-
co-electrolysis at ambient pressure seems to be the best fit to produce lectivity under FTS against traditionally impregnated catalysts. How-
FT products from industrial CO2 in terms of plant efficiency. This so- ever, if the ALD seems to be a very attracting technological solution, its
lution can replace the RWGS reactor and the low-temperature electro- cost for catalyst production is sensibly higher than traditional techni-
lyser into one single component and allow for high plant efficiency ques due to a low deposition rate [76]. As such, PtL systems require
thanks to thermoneutral operation. However, this device is not yet precise analysis and trade-off between installation and operation costs,
available at a commercial level for the required capacities, whereas the the first ones related to technologies and materials employed, the
RWGS is a proven technology. As a matter of fact, a solution as case A second ones linked, among other factors, to optimal system integration
could be considered at TRL 6, while a solution as case B would fall to before their implementation.
TRL much lower, making its deployment more expensive. On the con-
trary, a solution with the RWGS provides slightly lower ηGl but allows
for higher FT products synthesis when operating at 90% RR. As such, a

14
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

5. Conclusions recommended option, whereas pressurized case A would need


proper design and economic evaluation.
The present study provides a complete analysis of two systems for • Energy integration of the plant streams can drastically reduce the
the recovery of waste industrial carbon dioxide from a biogas upgrading energy input required from the outside, diminishing also the impact
plant, and its upgrade to Fischer-Tropsch synthetic hydrocarbons. Case that a high energy-intensive process like the biogas upgrade can
A uses an RWGS reactor for the generation of syngas, case B a SOEC have. Case A at 90% RR thermal needs reduce from 3.3 MWth to 0.9
operating under co-electrolysis conditions. The Fischer-Tropsch reactor MWth, while in case B the reduction is from 3.5 MWth to 0 MWth.
is modelled using a kinetic model that describes the formation of each • Given the TRL of the different technologies, fast and effective
FT product based on experimental results. commercialization of this specific process concept would need the
The analysis shows that: implementation of a P2L with an RWGS reactor. An economic as-
sessment including a real heat exchangers network is required to
• The biogas upgrading unit can separate 98% of CO 2 content in the identify the most proficient plant configuration.
biogas stream at the expenses of high energy consumption (4.75
GJth/tonCO2); CRediT authorship contribution statement
• When operating under a baseline condition of 0% RR, the solution
with the RWGS converts a lower amount of CO2 to useful products Marco Marchese: Methodology, Formal analysis, Investigation,
due to an intrinsic limitation of the technology at equilibrium con- Software, Data curation, Writing - original draft, Visualization.
dition. Moreover, case B reaches a slightly higher ηGlob of 67.8% Emanuele Giglio: Supervision, Formal analysis, Investigation,
against 64.5% of case A thanks to the higher energy content of the Resources, Writing - review & editing. Massimo Santarelli:
FT products; Supervision, Writing - review & editing. Andrea Lanzini: Supervision,
• Applying a recirculation of the FT off-gases to a rate as high as 90% Project administration, Writing - review & editing.
allows increasing exponentially the production of synthetic hydro-
carbons, the deployment CO2 and the overall plant efficiency.
Declaration of Competing Interest
Similarly, the configuration with the SOEC stack provides an effi-
ciency of 66.7% against 63.4% of case A. On the contrary, the re-
circulation allows reaching a higher production of syncrude when The authors declare that they have no known competing financial
applying an RWGS reactor; interests or personal relationships that could have appeared to influ-
• Regardless of the technology, increasing the operating pressure of ence the work reported in this paper.
the syngas generator to 25 bar (thus matching the pressure of the FT
reactor) reduces the production of useful syncrude. However, Acknowledgements
overall plant efficiency is only a few percentage points lower with
respect to the ambient pressure cases. Moreover, while RWGS re- This work has received funding from the European Union’s Horizon
actors operating at high pressure are commercially available, SOEC 2020 research and innovation programme under grant agreement No
systems under co-electrolysis are only present at the laboratory and 768543 (ICO2CHEM “From industrial CO2 streams to added value
demonstration scale. As such, a pressurized case B is not a Fischer-Tropsch chemicals”).

Appendix A

Table A1
Specifications of the biogas section components.
Preheater outlet 313 K Solvent Cooler outlet 313 K

Absorber Rate-based Stripper Rate-based


1.07 bar (Δp 0.5 bar) 1.6 bar (Δp 0.5 bar)
14 stages 14 stages (with condenser and reboiler)
1 m section diameter 1 m section diameter
Mixed Flow model Mixed Flow model
Interfacial area factor 1.2 Interfacial area factor 1.2
Reaction condition factor 0.9 Reaction condition factor 0.9
Film discretization ratio 5 Film discretization ratio 5
Distillate to feed ratio 0.05

Table A2
Reaction coefficients for the chemical reactions of the biogas upgrading unit Eq. (1)–(5) [31].
Reaction Coefficients

A B C D

(1) 132.89 −13445.9 –22.47 0


(2) 231.46 −12092.1 −36.78 0
(3) 216.05 −12431.7 −35.48 0
(4) −3.038 −7008.36 0 −0.0031
(5) −0.52 −2545.53 0 0

15
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table A3
Reaction coefficients for the chemical reactions of the biogas upgrading unit Eqs.
(7)–(10) [77].
Reaction Aj Eact [cal/mol]

(7) 9.77e + 10 9850


(8) 2.18e + 18 14138.4
(9) 4.32e + 13 13,249
(10) 2.38e + 17 29,451

Appendix B

k1PCO
1 =
k1PCO + k 6MPH2 (43)

k1PCO
2 =
k1PCO + k 6PH2 + k7Eec2 (44)

k1PCO
N = n 3
k1PCO + k 6PH2 + k7ecn (45)

[vac]
nc j nc j nc j
1 1 PH2O
= 1/{1 + (K 0 PH2) 1+ 1 + 1 2 + n + 1 + 1 2 + n n + +
j=3 n=3 j=3 n=3 j=3 n=3 K4 (K 0 PH2) K3K 4 (K 0PH2 )(1.5) K2K3K 4K5
1 PH2O
+ }
K5 (K 0PH2) (46)

Appendix C

Table C1
Information about the components employed in this study.
Component Description Technical information

Biogas upgrading section


B1 Biogas blower Target pressure 1.12 bar
Absorber Absorber of the biogas upgrading section 14 stages, 1.12 bar, Δp 0.05 bar, 313 K
Stripper Stripper of the biogas upgrading section, with condenser top stage 14 stages, 1.6 bar, Δp 0.1 bar, partial-vapour condenser (362.3 K), kettle reboiler
and reboiler bottom stage (388.8 K)
P1-b Circulation pump Target pressure 1.6 bar, ηel 0.95
B2-b Methane compressor Target pressure 10 bar; ηis 0.85, ηmech 0.95
H1 Biogas heater Target temperature 313 K
HX1-b CH4 heater Target temperature 293 K
Hx2-B Make-Up flow heater Target temperature 313 K
C1 CO2 cooler Target temperature 288 K
CX1-b CH4 cooler Target temperature 283 K
CX2-b CH4 cooler after compression Target temperature 558 K
CX3-b Looping cooler pre absorber Target temperature 313 K
Cross-HX Heat exchanger for heat recovery at the biogas upgrading unit ΔT 15 K
M1-b Mixer for make-up flow injection

Case A and B at reference condition

Alk. electrolyser Conversion of water to hydrogen in case A, accounting for 15 L/kgH2 consumed, 50 kWel/kgH2
auxiliaries consumptions
RWGS reactor Isothermal reactor working at equilibrium condition for the Operating at 1073 K, 1–25 bar, Δp 0.1 bar
generation of syngas
FT reactor Isothermal reactor working at rate-based conditions for the Operating at 501 K, 25 bar, Δp 0.1 bar
generation of FT products
SOEC stack Conversion of CO2 and H2O to syngas in case B Operating at 1073 K, 1–25 bar, Δp 0.1 bar
3-phase separator Separation of water from the gas and liquid FT fractions at constant Operating at 298 K
temperature
Col1 Petrochemical fractionations of the FT products, with one side 54 stages, 1 bar, reflux ratio 1.8, partial-vapour-liquid condenser (301 K), kettle
stripper: gas, naphtha, middle distillate, wax reboiler (670 K); Stripper (16 stage) draw stage 18, overhead return stage 12, top stage
condenser 456 K
Col2 Petrochemical fractionation of the FT wax fraction into light and 58 stages, 1 bar, reflux ratio 1.8, total condenser (656 K), kettle reboiler (793 K)
heavy waxes
Furnace Combustion of the FT off-gases and unconverted compounds Operating at 1273 K, ambient pressure
(continued on next page)

16
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

Table C1 (continued)

Component Description Technical information

V1 Lamination valve Target pressure 1 bar


V2 Lamination valve Target pressure 1 bar
V3 Lamination valve Target pressure 1 bar
V4 Lamination valve Target pressure 1 bar
B2 Pump for water entering in the alkaline electrolyser, only for case A Target pressure 15 bar
(consumption accounted in the alkaline electrolysers block)
B3 Intercooling compressor, only with ambient air operation of the Pressure ratio 2.924, from ambient pressure; ηis 0.85, ηmech 0.95
syngas units
B4 Intercooling compressor, only with ambient air operation of the Pressure ratio 2.924; ηis 0.85, ηmech 0.95
syngas units
B5 Intercooling compressor, only with ambient air operation of the Pressure ratio 2.924; ηis 0.85, ηmech 0.95
syngas units
C2 Condenser at the outlet of the syngas generator units Target temperature 293 K
C3 Intercooling cooler, only with ambient air operation of the syngas Target temperature 373 K
units
C4 Intercooling cooler, only with ambient air operation of the syngas Target temperature 373 K
units
C5 Intercooling cooler, only with ambient air operation of the syngas Target temperature 501 K
units
C6 Cooling simulation at the entrance of the 3-phase separator Target temperature 298 K
C7 Cooling the exhaust gases leaving the furnace Target temperature 323 K
C8 Cooling the naphtha FT fraction Target temperature 298 K
C9 Cooling the middle distillate FT fraction Target temperature 298 K
C10 Cooling the light waxes FT fraction Target temperature 298 K
C11 Cooling the heavy waxes FT fraction Target temperature 298 K
H2 Preheating of the gas mixture at the inlet of the syngas generator Target temperature 1073 K
units
H3 Preheating of the FT liquid entering the first distillation column Target temperature 523 K
H4 Preheating of the wax fraction entering the second distillation Target temperature 728 K
column
H5 Preheating of air entering the furnace Target temperature 473 K
H6 Preheating of the FT off-gases recirculated to the syngas generation Target temperature 1073 K
unit
H7 Heating for steam generation entering the SOEC, only for case B Target temperature 1073 K
H-Cdep Heater for steam generation needed to avoid carbon deposition, only Target temperature 1073 K
for case A
B-Cdep Pump for water needed to avoid carbon deposition, only for case A Target pressure 1–25 bar, ηel 0.95
M1 Mixer of the syngas generators reactants
M2 Mixer of FT off-gases
M3 Mixer of FT off-gases and air at the furnace inlet
S1 Splitter for recirculation of the FT off-gases

SOEC Section

Electr. Reactor Stoichiometric reactor accounting for electrochemical conversion of Operating at 1073 K, 1–25 bar, Δp 0.1 bar
CO2 and H2O, accounting for pressure drop
Ch. Eq. (1) Equilibrium reactions taking place inside the SOEC Operating at 1073 K
Ch. Eq. (2) Equilibrium reactions taking place inside the SOEC Operating at 1073 K
O2 Sep. Simulator of O2 generation inside the SOEC 100% recovery of O2
HX1-s Preheating of the inlet air at the oxygen electrode with recovery of ΔT 15 K with CX1-s
heat from the outlet of the oxygen electrode
HX2-s Heater of the air entering the oxygen electrode of the SOEC Target temperature 1073 K
CX1-s Cooling of the O2-rich air at the outlet of the oxygen electrode, with ΔT 15 K with HX1-s
heat recovery to the inlet air
M1-s
M2-s
S1-s

Components needed in the case of pressurised syngas generator units

B1-HP Compressor for CO2 Target pressure 25 bar; ηis 0.85, ηmech 0.95
B2-HP Compressor for H2, only case A Target pressure 25 bar; ηis 0.85, ηmech 0.95
B3-HP Compressor for recirculation Target pressure 25 bar; ηis 0.85, ηmech 0.95
B4-HP Pump for water, only case B Target pressure 25 bar, ηel 0.95
H1-HP Heater for syngas entering the FT reactor Target temperature 501 K

Appendix D. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ecmx.2020.100041.

17
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

References [26] Comidy LJF, Staples MD, Barrett SRH. Technical, economic, and environmental
assessment of liquid fuel production on aircraft carriers. Appl Energy
2019;256:113810. https://doi.org/10.1016/j.apenergy.2019.113810.
[1] Jones CR, Olfe-Kräutlein B, Naims H, Armstrong K, Jones CR. The Social acceptance [27] Marchese M, Heikkinen N, Giglio E, Lanzini A, Lehtonen J, Reinikainen M. Kinetic
of carbon dioxide utilisation: a review and research agenda. Front Energy Res study based on the carbide mechanism of a Co-Pt/γ-Al2O3 Fischer-Tropsch catalyst
2017;5:1–13. https://doi.org/10.3389/fenrg.2017.00011. tested in a laboratory-scale tubular reactor. Catalysts 2019;9:717. https://doi.org/
[2] Chehade Z, Mansilla C, Lucchese P, Hilliard S, Proost J. Review and analysis of 10.3390/catal9090717.
demonstration projects on power-to-X pathways in the world. Int J Hydrogen [28] Tjaden B, Gandiglio M, Lanzini A, Santarelli M, Järvinen M. Small-scale biogas-
Energy 2019;44:27637–55. https://doi.org/10.1016/j.ijhydene.2019.08.260. SOFC plant: technical analysis and assessment of different fuel reforming options.
[3] Cuellar-Franca RM, Azapagic A. Carbon capture, storage and utilisation technolo- Energy Fuels 2014. https://doi.org/10.1021/ef500212j.
gies: a critical analysis and comparison of their life cycle environmental impacts. J [29] Raynal L, Bouillon PA, Gomez A, Broutin P. From MEA to demixing solvents and
CO2 Util 2014;9:82–102. https://doi.org/10.1016/j.jcou.2014.12.001. future steps, a roadmap for lowering the cost of post-combustion carbon capture.
[4] Panahi M, Yasari E, Rafiee A. Multi-objective optimization of a gas-to-liquids (GTL) Chem Eng J 2011;171:742–52. https://doi.org/10.1016/j.cej.2011.01.008.
process with staged Fischer-Tropsch reactor. Energy Convers Manag [30] Li K, Leigh W, Feron P, Yu H, Tade M. Systematic study of aqueous mono-
2018;163:239–49. https://doi.org/10.1016/j.enconman.2018.02.068. ethanolamine (MEA)-based CO2 capture process: techno-economic assessment of
[5] Ye C, Dang M, Yao C, Chen G, Yuan Q. Process analysis on CO2 absorption by the MEA process and its improvements. Appl Energy 2016;165:648–59. https://doi.
monoethanolamine solutions in microchannel reactors. Chem Eng J org/10.1016/j.apenergy.2015.12.109.
2013;225:120–7. https://doi.org/10.1016/j.cej.2013.03.053. [31] Moioli S, Pellegrini LA, Gamba S. Simulation of CO2 capture by MEA scrubbing with
[6] Hombach LE, Doré L, Heidgen K, Maas H, Wallington TJ, Walther G. Economic and a rate-based model. Procedia Eng 2012;42:1651–61. https://doi.org/10.1016/j.
environmental assessment of current (2015) and future (2030) use of E-fuels in proeng.2012.07.558.
light-duty vehicles in Germany. J Clean Prod 2019;207:153–62. https://doi.org/10. [32] Vidal Vázquez F, Pfeifer P, Lehtonen J, Piermartini P, Simell P. V. Alopaeus, catalyst
1016/j.jclepro.2018.09.261. screening and kinetic modeling for CO production by high pressure and tempera-
[7] Hanaoka T, Miyazawa T, Shimura K, Hirata S. Jet fuel synthesis from Fischer- ture reverse water gas shift for Fischer-Tropsch applications. Ind Eng Chem Res
Tropsch product under mild hydrocracking conditions using Pt-loaded catalysts. 2017;56:13262–72. https://doi.org/10.1021/acs.iecr.7b01606.
Chem Eng J 2015;263:178–85. https://doi.org/10.1016/j.cej.2014.11.042. [33] Christensen KO. Steam Reforming of Methane on Different Nickel Catalysts.
[8] Kiss AA, Pragt JJ, Vos HJ, Bargeman G, de Groot MT. Novel efficient process for Norwegian University of Science and Technology; 2005Hdl.handle.net/11250/
methanol synthesis by CO2 hydrogenation. Chem Eng J 2016;284:260–9. https:// 248086.
doi.org/10.1016/j.cej.2015.08.101. [34] Tractebel, Engie, Hinicio, Study on Early Business Cases for H2 in Energy Storage
[9] Salomone F, Giglio E, Ferrero D, Santarelli M, Pirone R, Bensaid S. Techno-eco- and More Broadly Power To H2 Applications, 2017. http://www.hinicio.com/inc/
nomic modelling of a Power-to-Gas system based on SOEC electrolysis and CO2 uploads/2017/07/P2H_Full_Study_FCHJU.pdf%0Ahttp://www.fch.europa.eu/
methanation in a RES-based electric grid. Chem Eng J 2019;377. https://doi.org/ sites/default/files/P2H_Full_Study_FCHJU.pdf%0Ahttp://www.hinicio.com/file/
10.1016/j.cej.2018.10.170. 2018/06/P2H_Full_Study_FCHJU.pdf.
[10] De Klerk A. Can Fischer-Tropsch syncrude be refined to on-specification diesel fuel? [35] Hauch A, Brodersen K, Chen M, Mogensen MB. Ni/YSZ electrodes structures opti-
Energy Fuel 2009;75:4593–604. https://doi.org/10.1021/ef9005884. mized for increased electrolysis performance and durability. Solid State Ionics
[11] Van der Laan GP, Beenackers AACM. Kinetics and selectivity of the fischer-tropsch 2016;293:27–36. https://doi.org/10.1016/j.ssi.2016.06.003.
synthesis: a literature review. Catal Rev 1999;4940:255–318. https://doi.org/10. [36] Zhang X, Song Y, Wang G, Bao X. Co-electrolysis of CO2 and H2O in high-tem-
1081/CR-100101170. perature solid oxide electrolysis cells: recent advance in cathodes. J Energy Chem
[12] Rafati M, Wang L, Dayton DC, Schimmel K, Kabadi V, Shahbazi A. Techno-economic 2017;26:839–53. https://doi.org/10.1016/j.jechem.2017.07.003.
analysis of production of Fischer-Tropsch liquids via biomass gasification: the ef- [37] Giglio E, Lanzini A, Santarelli M, Leone P. Synthetic natural gas via integrated high-
fects of Fischer-Tropsch catalysts and natural gas co-feeding. Energy Convers Manag temperature electrolysis and methanation: Part I - Energy performance. J Energy
2017;133:153–66. https://doi.org/10.1016/j.enconman.2016.11.051. Storage 2015;2:64–79. https://doi.org/10.1016/j.est.2015.06.004.
[13] Kim HH, Mazumder M, Lee M-S, Lee S-J. Effect of blending time on viscosity of [38] Zheng Y, Wang J, Yu B, Zhang W, Chen J, Qiao J, et al. A review of high tem-
rubberized binders with wax additives. Int J Pavement Res Technol perature co-electrolysis of H2O and CO2 to produce sustainable fuels using solid
2018;11:655–65. https://doi.org/10.1016/J.IJPRT.2018.03.003. oxide electrolysis cells (SOECs): advanced materials and technology. Chem Soc Rev
[14] Mehran MT, Bin Yu S, Lee DY, Hong JE, Lee SB, Park SJ, et al. Production of syngas 2017;46:1427–63. https://doi.org/10.1039/c6cs00403b.
from H2O/CO2 by high-pressure coelectrolysis in tubular solid oxide cells. Appl [39] Kazempoor P, Braun RJ. Hydrogen and synthetic fuel production using high tem-
Energy 2018;212:759–70. https://doi.org/10.1016/j.apenergy.2017.12.078. perature solid oxide electrolysis cells (SOECs). Int J Hydrogen Energy
[15] Ma R, Xu B, Zhang X. Catalytic partial oxidation (CPOX) of natural gas and re- 2015;40:3599–612. https://doi.org/10.1016/j.ijhydene.2014.12.126.
newable hydrocarbons/oxygenated hydrocarbons - a review. Catal Today [40] Clausen LR, Butera G, Jensen SH. High efficiency SNG production from biomass and
2019;338:18–30. https://doi.org/10.1016/j.cattod.2019.06.025. electricity by integrating gasification with pressurized solid oxide electrolysis cells.
[16] Ostadi M, Rytter E, Hillestad M. Evaluation of kinetic models for Fischer-Tropsch Energy. 2019:1117–31. https://doi.org/10.1016/j.energy.2019.02.039.
cobalt catalysts in a plug flow reactor. Chem Eng Res Des 2016;114:236–46. [41] Graves C, Ebbesen SD, Mogensen M, Lackner KS. Sustainable hydrocarbon fuels by
https://doi.org/10.1016/j.cherd.2016.08.026. recycling CO2 and H2O with renewable or nuclear energy. Renew Sustain Energy
[17] Moazami N, Wyszynski ML, Mahmoudi H, Tsolakis A, Zou Z, Panahifar P, et al. Rev 2011;15:1–23. https://doi.org/10.1016/j.rser.2010.07.014.
Modelling of a fixed bed reactor for Fischer-Tropsch synthesis of simulated N2-rich [42] Saeidi S, Nikoo MK, Mirvakili A, Bahrani S, Saidina Amin NA, Rahimpour MR.
syngas over Co/SiO2: Hydrocarbon production. Fuel 2015;154:140–51. https://doi. Recent advances in reactors for low-temperature Fischer-Tropsch synthesis: process
org/10.1016/j.fuel.2015.03.049. intensification perspective. Rev Chem Eng 2015;31:209–38. https://doi.org/10.
[18] dos Santos RG, Alencar AC. Biomass-derived syngas production via gasification 1515/revce-2014-0042.
process and its catalytic conversion into fuels by Fischer Tropsch synthesis: a re- [43] Van De Loosdrecht J, Botes FG, Ciobica IM, Ferreira A, Gibson P, Moodley DJ, Saib
view. Int J Hydrogen Energy 2019. https://doi.org/10.1016/j.ijhydene.2019.07. AM, Visagie JL, Weststrate CJ, Niemantsverdriet JWH. Fischer-Tropsch Synthesis:
133. Catalysts and Chemistry Elsevier Ltd.; 2013. https://doi.org/10.1016/B978-0-08-
[19] Cinti G, Baldinelli A, Di A, Desideri U. Integration of solid oxide electrolyzer and 097774-4.00729-4.
Fischer-Tropsch: a sustainable pathway for synthetic fuel. Appl Energy [44] Tucker CL, van Steen E. Activity and selectivity of a cobalt-based Fischer-Tropsch
2016;162:308–20. https://doi.org/10.1016/j.apenergy.2015.10.053. catalyst operating at high conversion for once-through biomass-to-liquid operation.
[20] Fazeli H, Panahi M, Rafiee A. Investigating the potential of carbon dioxide utili- Catal Today 2018. https://doi.org/10.1016/j.cattod.2018.12.049.
zation in a gas-to-liquids process with iron-based Fischer-Tropsch catalyst. J Nat [45] Yang J, Ma W, Chen D, Holmen A, Davis BH. Fischer-Tropsch synthesis: a review of
Gas Sci Eng 2018;52:549–58. https://doi.org/10.1016/j.jngse.2018.02.005. the effect of CO conversion on methane selectivity. Appl Catal A Gen
[21] Selvatico D, Lanzini A, Santarelli M. Low temperature Fischer-Tropsch fuels from 2014;470:250–60. https://doi.org/10.1016/j.apcata.2013.10.061.
syngas: kinetic modeling and process simulation of different plant configurations. [46] Luyben WL, Yu CC. Reactive distillation design and control. Wiley; 2008. https://
Fuel 2016;186:544–60. https://doi.org/10.1016/j.fuel.2016.08.093. doi.org/10.1002/9780470377741.
[22] Herz G, Reichelt E, Jahn M. Techno-economic analysis of a co-electrolysis-based [47] Al Ashraf A, Al A. Aftab, Distillation Process of Crude Oil. Qatar University; 2012.
synthesis process for the production of hydrocarbons. Appl Energy [48] Campanario FJ, Gutiérrez Ortiz FJ. Fischer-Tropsch biofuels production from
2018;215:309–20. https://doi.org/10.1016/j.apenergy.2018.02.007. syngas obtained by supercritical water reforming of the bio-oil aqueous phase.
[23] Rafiee A, Panahi M, Khalilpour KR. CO2 utilization through integration of post- Energy Convers Manag 2017;150:599–613. https://doi.org/10.1016/j.enconman.
combustion carbon capture process with Fischer-Tropsch gas-to-liquid (GTL) pro- 2017.08.053.
cesses. J CO2 Util 2017;18:98–106. https://doi.org/10.1016/j.jcou.2017.01.016. [49] Stempien JP, Ni M, Sun Q, Chan SH. Thermodynamic analysis of combined solid
[24] Vázquez Francisco Vidal, Koponen Joonas, Ruuskanen Vesa, Bajamundi Cyril, oxide electrolyzer and Fischer-Tropsch processes. Energy 2015;81:682–90. https://
Kosonen Antti, Simell Pekka, Ahola Jero, Frilund Christian, Elfving Jere, doi.org/10.1016/j.energy.2015.01.013.
Reinikainen Matti, Heikkinen Niko, Kauppinen Juho, Piermartini Paolo. Power-to-X [50] Beil M, Beyrich W. Biogas upgrading to biomethane. Biogas Handb Sci Prod Appl
technology using renewable electricity and carbon dioxide from ambient air: 2013:342–77. https://doi.org/10.1533/9780857097415.3.342.
SOLETAIR proof-of-concept and improved process concept. J CO2 Util [51] Sun Q, Li H, Yan J, Liu L, Yu Z, Yu X. Selection of appropriate biogas upgrading
2018;28:235–46. https://doi.org/10.1016/j.jcou.2018.09.026. technology-a review of biogas cleaning, upgrading and utilisation. Renew Sustain
[25] Tagomori IS, Rochedo PRR, Szklo A. Techno-economic and georeferenced analysis Energy Rev 2015;51:521–32. https://doi.org/10.1016/j.rser.2015.06.029.
of forestry residues-based Fischer-Tropsch diesel with carbon capture in Brazil. [52] Brand CV. CO2 Capture Using Monoethanolamine Solutions: Development and
Biomass Bioenergy 2019;123:134–48. https://doi.org/10.1016/j.biombioe.2019. Validation of a Process Model Based on the SAFT-VR Equation of State. London:
02.018. Imperial College; 2013.

18
M. Marchese, et al. Energy Conversion and Management: X 6 (2020) 100041

[53] Economico MDS. D.M. 19 febbraio 2007 Approvazione, 2007. carbon dioxide and residual biomass. J Clean Prod 2018;170:160–73. https://doi.
[54] Huang Y, Zhang X, Zhang X, Dong H, Zhang S. Thermodynamic modeling and as- org/10.1016/j.jclepro.2017.09.141.
sessment of ionic liquid-based CO2 capture processes. Ind Eng Chem Res [66] Samavati M, Santarelli M, Martin A, Nemanova V. Thermodynamic and economy
2014;53:11805–17. https://doi.org/10.1021/ie501538e. analysis of solid oxide electrolyser system for syngas production. Energy
[55] Zacchello B, Oko E, Wang M, Fethi A. Process simulation and analysis of carbon 2017;122:37–49. https://doi.org/10.1016/j.energy.2017.01.067.
capture with an aqueous mixture of ionic liquid and monoethanolamine solvent. Int [67] De Klerk A. Fischer-Tropsch refining. Univ. Pretoria; 2012.
J Coal Sci Technol 2017;4:25–32. https://doi.org/10.1007/s40789-016-0150-1. [68] Lillebø AH, Rytter E, Blekkan EA, Holmen A. Fischer-Tropsch synthesis at high
[56] Moioli S, Nagy T, Langé S, Pellegrini LA, Mizsey P. Simulation model evaluation of conversions on Al2O3 supported Co catalysts with different H2/CO levels (2017).
CO2 capture by aqueous MEA scrubbing for heat requirement analyses. Energy https://doi.org/10.1021/acs.iecr.7b01801.
Procedia 2017;114:1558–66. https://doi.org/10.1016/j.egypro.2017.03.1286. [69] Liu ZK, Ågren J, Hillert M. Application of the Le Chatelier principle on gas reac-
[57] Storsæter S, Chen D, Holmen A. Microkinetic modelling of the formation of C1 and tions. Fluid Phase Equilib 1996;121:167–77. https://doi.org/10.1016/0378-
C2 products in the Fischer-Tropsch synthesis over cobalt catalysts. Surf Sci 3812(96)02994-9.
2006;600f:2051–63. https://doi.org/10.1016/j.susc.2006.02.048. [70] Schmidt P, Weindorf W, Roth A, Batteiger V, Riegel F. Power-to-liquids: potentials
[58] Dimitriou I, García-Gutiérrez P, Elder RH, Cuéllar-Franca RM, Azapagic A, Allen and perspectives for the future supply of renewable aviation. Umweltbundesamt
RWK. Carbon dioxide utilisation for production of transport fuels: process and 2016https://www.umweltbundesamt.de/en/publikationen/.
economic analysis. Energy Environ Sci 2015;8:1775–89. https://doi.org/10.1039/ [71] Jarvis SM, Samsatli S. Technologies and infrastructures underpinning future CO2
c4ee04117h. value chains: a comprehensive review and comparative analysis. Renew Sustain
[59] Maitlis PM, de Klerk A. Greener Fischer-Tropsch processes for fuels and feedstocks, Energy Rev 2018;85:46–68. https://doi.org/10.1016/j.rser.2018.01.007.
2013. https://doi.org/10.1002/9783527656837. [72] Rytter E, Borg Ø, Tsakoumis NE, Holmen A. Water as key to activity and selectivity
[60] Wolf A, Jess A, Kern C. Syngas production via reverse water-gas shift reaction over a in Co Fischer-Tropsch synthesis: Γ-alumina based structure-performance relation-
Ni-Al2O3 catalyst: catalyst stability, reaction kinetics, and modeling. Chem Eng ships. J Catal 2018;365:334–43. https://doi.org/10.1016/j.jcat.2018.07.003.
Technol 2016;298:1040–8. https://doi.org/10.1002/ceat.201500548. [73] Gavrilović Save, Blekkan. The effect of potassium on cobalt-based Fischer-Tropsch
[61] Deka DJ, Gunduz S, Fitzgerald T, Miller JT, Co AC, Ozkan US. Production of syngas catalysts with different cobalt particle sizes. Catalysts 2019;9:351. https://doi.org/
with controllable H2/CO ratio by high temperature co-electrolysis of CO2 and H2O 10.3390/catal9040351.
over Ni and Co-doped lanthanum strontium ferrite perovskite cathodes. Appl Catal [74] Neill BJO, Jackson DHK, Lee J, Canlas C, Stair PC, Marshall CL, Elam W, Kuech TF,
B Environ 2019;248:487–503. https://doi.org/10.1016/j.apcatb.2019.02.045. Dumesic JA, Huber GW. Catalyst design with atomic layer deposition (2015).
[62] Chein R-Y, Yu C-T. Thermodynamic equilibrium analysis of water-gas shift reaction https://doi.org/10.1021/cs501862h.
using syngases-effect of CO2 and H2S contents. Energy 2017;141:1004–18. https:// [75] Eskelinen P, Keskiväli L, Heikkinen N, Franssila S. Cobalt catalyst characterization
doi.org/10.1016/j.energy.2017.09.133. and modification by atomic layer deposition for Fischer-Tropsch synthesis. Aalto
[63] Gunduz S, Deka DJ, Ozkan US. Advances in high-temperature electrocatalytic re- University; 2019.
duction of CO2 and H2O. 1st ed. Elsevier Inc; 2018. https://doi.org/10.1016/bs. [76] Oviroh PO, Akbarzadeh R, Pan D, Coetzee RAM, Jen T-C. New development of
acat.2018.08.003. atomic layer deposition: processes, methods and applications. Sci Technol Adv
[64] Leibbrandt NH, Aboyade AO, Knoetze JH, Görgens JF. Process efficiency of biofuel Mater 2019;20:465–96. https://doi.org/10.1080/14686996.2019.1599694.
production via gasification and Fischer-Tropsch synthesis. Fuel 2013;109:484–92. [77] Abu Zahra MRM. Carbon dioxide capture from flue gas: development and evalua-
https://doi.org/10.1016/j.fuel.2013.03.013. tion of existing and novel process concepts, 2009. http://www.google.com/
[65] Monaco F, Lanzini A, Santarelli M. Making synthetic fuels for the road transpor- patents/US20110226010%5Cnhttp://repository.tudelft.nl/view/ir/
tation sector via solid oxide electrolysis and catalytic upgrade using recovered uuid:6d1689f3-7b1a-4355-81b0-af3d19fae469/.

19

You might also like