Production of Synthetic Natural Gas From Industrial Carbon Dioxide

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

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

net/publication/338388059

Production of synthetic natural gas from industrial carbon dioxide

Article in Applied Energy · February 2020


DOI: 10.1016/j.apenergy.2019.114249

CITATIONS READS

74 2,673

5 authors, including:

Remi Chauvy Lionel Dubois


National Chung Hsing University Université de Mons
13 PUBLICATIONS 631 CITATIONS 37 PUBLICATIONS 730 CITATIONS

SEE PROFILE SEE PROFILE

Paul Lybaert Diane Thomas


Université de Mons Université de Mons
41 PUBLICATIONS 407 CITATIONS 92 PUBLICATIONS 2,144 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Remi Chauvy on 03 March 2020.

The user has requested enhancement of the downloaded file.


To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Production of Synthetic Natural Gas from Industrial Carbon


2 Dioxide
3
4 Remi CHAUVY1, Lionel DUBOIS2, Paul LYBAERT3, Diane THOMAS2, Guy DE WEIRELD1*
5
6 University of Mons, 20 Place du Parc, 7000 Mons, Belgium
7 1
Thermodynamics and Mathematical Physics Unit
8 2
Chemical and Biochemical Process Engineering Unit
9 3
Thermal Engineering and Combustion Unit
10
11 * Corresponding author. Tel.: +32 65374203
12 E-mail address: guy.deweireld@umons.ac.be (Guy De Weireld)
13
14 Abstract
15

16 The Power-to-Gas strategy has become a mainstream topic for decarbonization and development
17 of renewables and flexibility in energy systems. One of the key arguments for decarbonizing the gas
18 network is to take advantage of existing network infrastructure, gradually transitioning to lower fossil
19 carbon sources of methane from Power-to-Gas.
20
21 This work proposes the techno-economic investigation of an integrated system considering an
22 advanced CO2 capture process, in terms of solvent and process configuration, to treat about 10% of a
23 cement plant’s flue gas and convert the captured CO2 into synthetic natural gas using renewable
24 hydrogen generated from a large-scale wind powered electrolyzer. An optimized heat recovery system
25 is proposed, drastically decreasing the external hot utility demand of the CO2 capture unit. In addition,
26 it leads to the production of complementary electricity (about 1.06 MW), reducing thus also the electrical
27 demand of the integrated process.
28
29 The synthetic natural gas produced has a composition (CH4 92.9 mol.%, CO2 3.7 mol.%, and H2
30 3.4 mol.%) and a Wobbe index (46.72 MJ/m3), corresponding to specification for gas grid injection at
31 50 bar in Germany. With an overall system efficiency of 72.6%, the process produces 0.40 ton synthetic
32 natural gas per ton of captured CO2. The cost of the synthetic natural gas produced is higher when
33 compared to the present natural gas market price, but cost reductions and possible commercial use of
34 coproducts like oxygen, represent a likely alternative. Costs are mainly driven by high capital
35 investments (the electrolyzer), and the price of renewable electricity, which is expected to decrease in
36 the coming years.
37
38
39 Keywords
40
41 Carbon Capture and Utilization; Power-to-Gas; Process simulation; Techno-economic evaluation; Heat
42 integration

1
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 1 Introduction
2
3 Following the 2016 Paris Agreement (COP 21) to keep warming below 2 °C above pre-industrial
4 levels, pursuing efforts to limit the temperature increase to 1.5° C, above pre- industrial levels, the
5 European Union (EU) reaffirmed its goal to reduce greenhouse gas emissions by 80-95% compared to
6 1990 levels by 2050. In order to achieve these goals, large-scale implementation of low-carbon
7 technologies, such as renewable energy sources (RES), and carbon capture utilization and storage
8 (CCUS), are necessary. Numerous EU countries are therefore considering these two families of
9 technologies, such as Norway, Germany, and Denmark. Zappa et al. recently conducted a study
10 investigating whether a 100% renewable European power system would be feasible by 2050 [1]. They
11 concluded that a 100% renewable power system would still require significant flexible zero-carbon
12 storage capacity to balance variable renewable energy generation, and CCUS may still be required to
13 achieved the EU’s ambitious climate objective.
14
15 Nowadays, global wind power capacity is growing continuously (310 TWh per year in 2015 in
16 European countries with an installed capacity of 136 GW [1]) as the cost of producing renewable
17 electricity is sometimes even lower than the cost obtained from fossil fuels. Penetration of wind power
18 in the electricity market is already high in several EU countries such as Denmark (51%), and Ireland
19 (24%). Germany also denotes a high share of solar-based electricity (6%) [2]. However, the penetration
20 of renewable sources into the energy system is often limited by the energy storage capability, mainly
21 due to high costs. Therefore, for continuous availability and reliability, development of other efficient
22 grid storage alternatives is needed.
23
24 In this context, the Power-to-Gas (PtG) strategy is of growing interest for decarbonization and
25 increasing development of renewables and flexibility in energy systems [3]. In PtG applications, when
26 abundant renewable energy is available, mainly (but not necessarily exclusively) excess electricity can
27 be used to produce a storable gas, such as hydrogen produced by water electrolysis. Currently, only 4%
28 of hydrogen is produced by water electrolysis (overwhelmingly by steam reforming), where the total
29 hydrogen production around the world is about 500 bill.Nm³/year [4],[5]. This renewable hydrogen is
30 either used directly in industrial applications or in fuel cells [6]. Direct injection of hydrogen into the
31 natural gas grid (up to 10%) or conversion into synthetic natural gas (SNG) by methanation and
32 subsequent injection into the gas grid are widely considered in the literature [2],[7],[8],[9]. In addition
33 to maintain the high heating value of the gas, methanation denotes several advantages, as it avoids
34 adaptation in the gas grid equipment necessary for higher hydrogen contents. Furthermore, compared to
35 pure hydrogen, SNG (87-97 wt.% methane) has fewer barriers to implementation, as it is safer, easier
36 to transport and store, and more suitable for industrial applications [10]. It overcomes the facility and
37 energy density issues related to the use of hydrogen, while providing other services [11]. It has to be
38 noticed that the conversion is limited in the term ‘Power-to-Gas’ to the production of hydrogen and
39 methane, the later in the form of SNG. Other converting forms of electrical energy to liquid energy
40 carriers (hydrocarbons such as methanol, dimethyl ether, and Fischer-Tropsch products) are more likely
41 to associate the term ‘Power-to-Fuel’ or ‘Power-to -Liquid’ [12]. Even though hydrocarbons denote a
42 higher volumetric density in comparison to gases, PtG allows the use of the existing natural gas
43 infrastructure as transport and storage medium, offering a market availability of all the system-relevant
44 components. Furthermore, wind-to-SNG offers an energy storage medium for the intermittent wind
45 resource, where the electricity generated by the renewable energy is merged into the grid providing a
46 more constant renewable energy supply, avoiding over-supplying power grids with high production
47 from renewable energy.
48
49 Concurrently, many industrial processes look for possibilities to decrease their CO2 emissions,
50 where CCUS is considered. Accounting for 2 to 2.5 GtCO2 per year (5 to 7% of the total anthropogenic
51 CO2 emissions) [13],[14], the cement sector is investigating ways to cut its emissions through different
52 levers such as modern dry-process technology, clinker substitution, free-carbon alternative fuels and
53 CCUS. It is estimated that only about one third of the CO2 released is due to combustion, and this makes
54 the cement industry a specific case compared with other combustion industries. Clinker, main
55 constituent of the cement, is made by heating a homogeneous mixture of raw materials at minimum

2
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 temperature of 1,450 °C. Two thirds of these released emissions, i.e. 550 kgCO2 per ton of clinker, come
2 from the limestone calcination during the decarbonation step in the clinker burning process, and are then
3 necessarily emitted [15],[16]. Hence, the capture of CO2 from cement plant’s flue gas and its conversion
4 are key issues. This CO2 can be a source for methanation.
5 Several technologies at different levels of maturity and performances can be envisaged for
6 capturing the CO2 from cement kilns (e.g. oxy-fuel combustion, chilled ammonia technology, adsorptive
7 processes, calcium looping, etc.) [17],[18],[19],[20]. The most mature technology is the absorption-
8 regeneration amine-based process, which reached a TRL (Technology Readiness Level) of 9 compared
9 to the lower TRLs of all other technologies. More precisely, a recent study from the IEA Greenhouse
10 Gas R&D Programme [21] pointed out that among the different liquid absorbents that can be used for
11 capturing CO2 (e.g. aqueous amine (TRL 6-9), amino acid or mixed salts (TRL 6), ionic liquids (TRL
12 4), water-lean absorbents (TRL 5), precipitating and demixing solvents (TRL 4-6), etc.), aqueous
13 solutions composed of amine blends (e.g. piperazine (PZ) mixed with another amine such 2-amino-2-
14 methyl-1-propanol (AMP) or methyldiethanolamine (MDEA)) can be considered as the new reference
15 solvents in place of monoethanolamine (MEA), long time considered as benchmark for such CO2
16 capture processes [21]. These amines blends and especially activated solutions allow to reduce
17 significantly the energy demand of the system. Regarding the other post-combustion CO2 capture
18 technologies, membrane systems (TRL 5-6) and solid sorbent processes (TRL 6) (e.g. Vacuum Pressure-
19 Swing Adsorption (VPSA), calcium looping, etc.) have also received a growing interest [22],[23],[24].
20 The widespread deployment of pilot and demonstration plants, including both large-scale demonstration
21 projects and commercial size amines-based post-combustion worldwide installations, such as Boundary
22 Dam (Canada), Petra Nova (USA), and Shengli (China), were reviewed by Idem et al. [25].
23 In the case of the post-combustion CO2 capture by absorption-regeneration processes applied in a
24 cement plant, the CO2 is separated from the exhaust gases of the system by adding a unit to the tail-end
25 of the clinker process. While this end-of- pipe mature technology achieves high absorption efficiency
26 (usually 90%), the main disadvantage is the high energy demand for the scrubbing liquid regeneration
27 (e.g. from 3.2 GJ/t CO2 to around 3.8 GJ/t CO2 with MEA 30 wt.% as solvent [26]). Nevertheless, the
28 development of new solvents (new solutions or blends) [21],[27], and new process configurations
29 [28],[29] are allowing the global performances of the absorption-regeneration technology to reduce its
30 energy consumption to be improved, so that this technology remains attractive and competitive.
31
32 The Power-to-Gas process chain was first proposed in Japan in the late 1980s. Recently, a growing
33 interest dealing with this technology has begun, especially in Europe. In a detailed review, Rönsch et al.
34 give an overview of methanation technology and research, focusing on projects with methane as a
35 product [30]. Chauvy et al. maintain that PtG is one of the best alternative processes for CO2 utilization
36 among examined cases in the sensitivity analysis [31]. The interest is mainly driven by the increasing
37 share of renewable energy (wind and solar power), where PtG may provide large-scale and long-term
38 energy storage [32]. Other field of applications include services to balance the loads in electricity
39 networks, a substantial source of fuel for heating and transportation, and a significant contribution to
40 emission reduction targets [33]. Pilot plants are under construction, or even in operation, in several
41 countries, including Germany, Switzerland, Denmark, France, and Japan. In particular, the European
42 Power-to-Gas Platform contains a database of past, current and planned PtG projects in Europe [34].
43 Numerous reviews of PtG projects present lab, pilot and demo plants for storing renewable energy and
44 CO2 [7],[35],[36]. Quarton and Samsatli is particularly recommended for more detail on real-life PtG
45 projects, economic assessments and systems modeling [7]. Bailera et al. detail worldwide existing
46 projects and present basic information together with qualitative descriptions of the plants, technical data,
47 budget and project partners [35]. Wulf et al. reference about 130 demonstration projects in operation or
48 planning in Europe (63 projects were in operation by end of May 2018), about 40% of which were
49 developed in Germany [36].
50 One example project that illustrates the concept of PtG is the e-Gas Project (Werlte, Germany)
51 [8],[35]. It demonstrates how renewable energy could be efficiently stored in the existing natural gas
52 network. It consists of methanation combined to an organic waste biogas plant, which produces about 1
53 kton methane-rich gas from concentrated CO2 and renewable H2. Three alkaline electrolyzer stacks of 2
54 MW supplied from a 14.4 MW capacity offshore wind park generate the H2. The annual SNG injection
55 to the grid from this process is approximately 3 million Nm3.

3
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 In the meantime, many studies focus on thermodynamic considerations, the development of new
2 catalysts, the characterization of the reaction kinetics and mechanisms of CO2 methanation [37],[38].
3 For instance, Su et al. proposed a review on the advances in catalytic CO2 hydrogenation to methane,
4 discussing the structure of the catalysts, the preparation methods, the pre-treatment conditions and their
5 components in detail [37].
6 In addition, Eveloy and Gebreegziabher recently reviewed projected PtG deployment scenarios at
7 regional and distributed scales [39]. They highlighted that even though many types of PtG
8 energy/material integrations are possible, few of them have been incorporated into PtG deployment
9 modelling works. Furthermore, they identified that heat recycling from catalytic methanation reaction
10 products, feed compression trains, and condensed water integrated with CO2 capture from cement
11 production and/or water electrolysis requirements have received limited attention. This lack of
12 experience with the whole PtG system was also highlighted by Quarton and Samsatli [7], Ghaib and
13 Ben-Fares [8], and Thema et al. [40]. It has to be pointed out that the development of PtG technology is
14 also subject to fundamental energy and climate policy decisions.
15
16 Therefore, only a few studies attempt to evaluate large-scale integrated PtG process performances,
17 showing a large potential for improvement, which helps the decision makers to justify their choices.
18 Becker et al. recently proposed a detailed SNG plant study case where the synergies between large-scale
19 reactors, thermal management and techno-economic considerations were provided [41]. Morosanu et al.
20 proposed a concept based on water electrolysis to produce hydrogen, CO2 capture from ambient air
21 using solid adsorption materials, catalytic CO2 methanation, gas separation, and a single mixed
22 refrigerant (SMR) methane liquefaction process, where mass and energy balances at demonstration scale
23 were provided [42].
24
25 To this extent, in-depth modelling and economic analysis of PtG systems are addressed in this
26 paper. The integration of SNG production with an advanced CO2 capture process, mainly in terms of
27 solvent and process configuration, and considering an optimized heat recovery leading to a
28 complementary electricity production, are especially innovative and correspond to the purpose of the
29 present study. The suggested PtG technology is flexible, easily up/down-scalable and modular in order
30 to allow an adjustment to any specific boundary conditions of a distinct application, such as end-product,
31 source of CO2, use of potential by-products, and main purpose of the system (utilization of excess
32 renewable energy, storage, etc.).

33 2 Materials and methods


34
35 The present PtG integrated process plant was developed to convert a part of CO2 coming from a
36 cement plant into SNG, and comprises four implementation steps: (i) an operational cement plant
37 equipped with the Best Available Technique (BAT), where an advanced CO2 separating unit for
38 scrubbing CO2 contained in the flue gas was considered; (ii) an electrolysis unit for splitting water by
39 electrolysis to produce renewable H2 and O2; (iii) a methanation unit for converting the CO2 and the
40 renewable H2 by a methanation reaction to a methane rich gas (SNG); and (iv) a SNG upgrading unit to
41 fulfil the requirements for further injection into the gas grid. Finally, the possibility of heat integration
42 through systematic process-to-process analysis is investigated to reach high energy efficiency and
43 minimize utility costs. It is worth noting that the produced SNG can also be converted into compressed
44 renewable natural gas or liquefied renewable natural gas to serve as a transport fuel.
45
46 The size of the PtG process is based on a 90 MW large-scale wind-based electrolyzer, generating
47 40 ton of renewable H2 per day. The flue gas to be treated comes from a conventional BAT cement plant
48 producing 3,000 tons of clinker per day. It corresponds to a total flow rate of 250,000 m³/h, at 1.20 bar
49 and 40 °C after conditioning (after desulfurization, denitrification, dedusting and cooling), with a CO2
50 content of 20 mol.%, generating 2,475 ton of CO2 per day [43]. Therefore, 9.8% of the cement plant flue
51 gas is treated where 90% of the CO2 is captured, representing a CO2 production of about 219 ton per
52 day. The CO2 capture unit can be upscaled in order to treat the entire cement plant’s flue gas, where the
53 captured CO2 can be geologically stored or converted into other products.
54

4
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Fig. 1 illustrates the conceptual flowsheet of this interconnected PtG system.


2
3
Other CCUS route(s)
4
Raw materials A
Cement kiln CO2-separation unit
5 Fuel Depleted flue gas
Flue gas Amine scrubbing

6
Cement
7 Scrubbing liquid
regeneration
8
CO2 Heat Excess heat for other
9 Excess RE B application
H2
Electrolysis Methanation
H2O
CO2/H2 Raw SNG
10
O2 H2O
Gas upgrading Grid injection
11 SNG

12 Fig. 1. Conceptual Flowsheet of the interconnected PtG system


13 comprising A: CO2 capture unit; B: CO2 methanation unit
14
15 2.1 Process modeling and design
16 2.1.1 Thermodynamics models and simulation tools
17
18 The units investigated in this study were implemented in AspenTech’s software. Concerning the
19 CO2 capture unit, it was simulated using the Acid Gas Property Package in Aspen HysysTM v10.0
20 software. More precisely, the thermodynamic models implemented in such package are the Electrolyte
21 Non-Random Two-Liquid (eNRTL) activity coefficient model for electrolyte thermodynamics in the
22 liquid phase [44] and the Peng-Robinson equation of state for the vapor phase [45]. The Acid Gas
23 Package also includes the physicochemical properties of acid gases (CO2 and H2S), water, amines alone
24 (e.g. MEA, DEA, MDEA, PZ, etc.) as well as several mixtures (e.g. MDEA+PZ, DEA+PZ, etc.). The
25 CO2 capture modelling is based on a rate-based calculation model and comprises a makeup unit
26 operation to automatically compensate water and amine losses, which was validated by Laribi et al. [46].
27
28 Aspen PlusTM v10.0 was used for both the CO2 methanation and gas upgrading units considering
29 the Peng Robinson equation of state for the calculation of gaseous properties. Estimates were considered
30 to model the block unit for hydrogen generation.
31
32 2.1.2 Renewable H2 production unit
33
34 Water electrolysis is mainly characterized by the source of electricity, i.e. high temperature steam,
35 solar or wind-based electrolysis, and the solution used as electrolyte, i.e. alkaline (AE), proton exchange
36 membrane (PEM) or solid oxide (SOEC) electrolyzers.
37 An essential block for PtG, electrolyzers need to fulfil some special requirements, including
38 dynamic behavior to follow fluctuating power inputs, high efficiency to avoid energy losses, elevated
39 pressure to reduce compressor costs, as well as long lifetimes. Even though alkaline electrolyzer
40 technology is fully mature, PEM water wind-based electrolysis, which is currently close to commercial
41 deployment, was chosen in this work as it is more compact and best suited to the dynamic load balancing
42 of electricity grids needed with use of intermittent renewable energy. Traditional alkaline electrolyzers,
43 on the other hand, are typically used for continuous steady-state industrial production of hydrogen
44 [47],[48]. The comparison of electrolyzer technologies is sum up in in the Supporting Information (See
45 SI.2). PEM electrolyzer produces hydrogen at 30 bar with a high purity level of 99.9 vol.%, whilst the
46 operating temperature is typically low (50-100 °C) [49]. The specific electrical power of these
47 electrolyzers is estimated at 5 kWh per Nm3, approximatively 55.6 kWh per kg H2. A production cost

5
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 of 4 to 5 € per kg H2, and a total system cost, including power supply and installation costs, of 1,200 €
2 per kW with an operation lifetime of 20 years are considered [41],[49]. The oxygen stream from the
3 water electrolysis is produced in high purity, and may be further used in the calcination process to
4 increase the CO2 partial pressure which facilitates CO2 separation (partial oxy-combustion)
5 [17],[19],[46]. This high purity by-product can also be traded for many high-tech applications.
6 The electrolyzers are assumed to be co-located with the sources of renewable electricity and CO2,
7 neglecting de facto the efforts to integrate the different units, such as infrastructures, and avoid transport
8 needs.
9
10 2.1.3 CO2 capture unit
11
12 The implemented CO2 capture process was based on the works of Dubois and Thomas [29],[50]
13 considering a Rich Vapor Compression (RVC) process combined with an Inter-Cooled Absorber (ICA)
14 and two Water-Wash (WW) sections, and the use of a MDEA 10 wt.% + PZ 30 wt.% blend as the
15 solvent. The selection of the configuration and the solvent are based on previous simulation results (see
16 Supporting Information for more details). Fig. 2 presents the implemented flow sheet.
17

18
19 Fig. 2. Aspen HysysTM flow sheet of the RVC+ICA+WW configuration with MDEA+PZ as solvent

20 The installation was designed taking the pilot unit used in the CASTOR/CESAR European Projects as
21 reference [51]. The CASTOR/CESAR unit was selected as a reference since all the design and operating
22 parameters in relation with the installation are available in literature [51]. In the current study, this unit
23 was upscaled (columns diameter) in order to treat 24,660 m³/h of flue gas with a CO2 absorption ratio
24 of 90 mol.%, corresponding to a production of 219 ton per day of CO2 (98 mol.% purity). The upscaling
25 was performed to keep the same gas and liquid velocities, 1.29 m/s and 6.70 10-3 m/s, respectively, such
26 as the optimum value for the (L/G)vol ratio, namely around 5.19 10-3 m³/m³.
27 The dimensions of the absorber and the stripper, and the operating conditions for each column are
28 provided in Table 1. Note that 𝑇𝑎𝑏𝑠 corresponds to the absorption temperature, at the top of the
29 absorption column, while 𝑇𝑟𝑒𝑔𝑒𝑛 refers to the regeneration temperature, at the bottom of the stripper.
30 PZ-based solvents being conventionally regenerated at higher temperature for optimal performances
31 [29], a regeneration pressure (𝑃𝑏𝑜𝑡𝑡𝑜𝑚 of the stripper) of 5 bar was considered, leading to a regeneration
32 temperature of 140.8 °C, in contrast to a temperature 𝑇𝑟𝑒𝑔𝑒𝑛 of around 120 °C with a conventional MEA
33 30 wt.% solvent.

6
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 1. Main process data equipment for the CO2 capture unit

Parameter Absorber Stripper


Diameter (m) 2.6 2.6
Packing height (m) 17 (17 x 1 m) 10 (10 x 1 m)
Random packing IMTP Norton Metal 50 mm
Packing type
MELLAPACK, SULZER, Standard, 125 X (for the water wash section)
𝑇𝑎𝑏𝑠 (°C) 40 -
𝑇𝑟𝑒𝑔𝑒𝑛 (°C) - 140.8
𝑃𝑏𝑜𝑡𝑡𝑜𝑚 (bar) 1.2 5
2

3 The RVC configuration (applied to the rich solution, i.e. after the absorption step) uses a heat pump
4 effect. As described by Le Moullec et al., the purpose of the RVC process modification is to increase
5 the heat quality provided to the system by enabling the valorization of heat available at a lower quality
6 level or when increasing the quality level is energetically interesting [28]. Thanks to the flashing of the
7 solvent, a gaseous stream mainly composed of CO2 and H2O is produced. This stream is compressed
8 and sent back to the regeneration column in order to reduce the steam demand at the reboiler. At the
9 same time, with the vapor coming from the compressor following the flash tank being quite hot (which
10 could lead to hot spot and degradation problems in the stripper’s bottom if it is reinjected at such
11 temperature level), a second internal heat exchanger is added in order to cool this vapor down to the
12 same temperature level at the bottom of the regeneration column (𝑇𝑟𝑒𝑔𝑒𝑛 ). Such an operation also has the
13 advantage of improving the preheating of the rich solution before entering the stripper.
14 Concerning Intercooling, this is a well-described technique which has already been applied in
15 several pilot or industrial units. It can also be considered in the “absorption enhancement” process
16 modification category. The principle of intercooling is to withdraw the solvent (partially or totally)
17 flowing in the absorber, at a stage 𝑛, to cool it down and to send it back to the absorber at the stage 𝑛 −
18 1. Such a method enables a shift in the thermodynamic gas-liquid equilibrium, and consequently
19 increases the rich loading at the absorber bottom (at a same liquid flow rate) [28]. However, a lower
20 temperature in the absorber leads to reduced chemical kinetics and diffusivities which is mostly
21 compensated by the increase of CO2 solubility leading globally to very little change to the overall mass
22 transfer coefficient. Globally, such a technical option is expected to decrease the reboiler duty, as
23 highlighted by Le Moullec et al. [28].
24 Regarding the water-wash sections, as described in the IEAGHG report prepared by CSIRO
25 (Australia) [52], scrubbers are used for extracting condensable or soluble vapor from gases. Amine
26 vapors and their degradation products are conventionally captured by water wash or a scrubbing stage,
27 or several stages with demineralized water, acidic water or with special reagents. In the present work,
28 and in accordance with the CASTOR/CESAR pilot unit considered as a reference, two water-wash
29 sections were added to the simulation flow sheet in comparison with the previous model described by
30 Dubois and Thomas [29]. As such, one water-wash section was used at the top of the absorber, consisting
31 of a packed column composed of three-stages of 1 m height per stage, and another wash section was
32 used directly inside the top of the stripper consisting of three stages of packing (same dimensional
33 parameters as for the main column section) below the condenser and above the rich solution inlet, in
34 which the vapor coming from the regeneration meets the liquid coming from the condensation. In the
35 absorber water-wash section, a freshwater make-up is added and also serves as water make-up for the
36 global flow sheet. Only a small complement is added into the make-up unit.
37 It is assumed that the pre-treated flue gas (after de-SOx, de-NOx, de-dusting and cooling) entering
38 the CO2 capture process (composition in Table 2) consists primarily of CO2, H2O, N2 and O2. This
39 composition is based on average values coming from the Brevik Cement plant (Norcem company) in
40 Norway. It is important to note that the amount of CO2 in the gas to be treated was maintained at 20.4
41 % for all the simulated cases. Nevertheless, the influence of the inlet gas CO2 content on the absorption-
42 regeneration performances has been discussed in previous works [29],[53].

7
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 2. Composition of the conditioned gas to be treated (G = 24,660 m³/h, 1.20 bar, 40 °C)

Component mol frac.


N2 64.7 %
CO2 20.4 %
O2 8.6 %
H2O 6.2 %
CO 1,330 ppm
SO2 111 ppm
NO 474 ppm
NO2 2 ppm
2
3 A summary of the simulation results of the implemented process highlighting its energetical benefit
4 in comparison with the conventional process using MEA 30 wt.% as solvent, is provided in the
5 Supporting Information, where the reactions that are included in the Acid Gas Package for PZ and
6 MDEA solvents reacting with CO2 are also presented (see SI.1).
7
8 2.1.4 CO2 methanation unit
9
10 Thermodynamics and kinetics. Methane production by catalytic CO2 hydrogenation (Sabatier
11 reaction) has been used since the beginning of the 20th century. The reaction scheme is given by [54]:
12
0
𝐶𝑂2 + 4𝐻2 ⇌ 𝐶𝐻4 + 2𝐻2 𝑂 ∆𝐻298 = −165 𝑘𝐽/𝑚𝑜𝑙 (1)
0
𝐶𝑂2 + 𝐻2 ⇌ 𝐶𝑂 + 𝐻2 𝑂 ∆𝐻298 = +41 𝑘𝐽/𝑚𝑜𝑙 (2)
0
𝐶𝑂 + 3𝐻2 ⇌ 𝐶𝐻4 + 𝐻2 𝑂 ∆𝐻298 = −206 𝑘𝐽/𝑚𝑜𝑙 (3)
13
14 High pressure and low temperature enhance the CO2 methanation process performance, as denoted
15 in Fig. 3. Below 600 °C, only species involved in the Sabatier reaction (see Eq.(1)) are present, with CO
16 appearing only for higher temperatures (see SI.3.1). Gao et al. underline that carbon deposition is not
17 observed under these simulation conditions for a H2/CO2 stoichiometric ratio of 4, as the water produced
18 during the methanation process reacts with solid carbon [55]. In this work, operating conditions are
19 chosen in such a way that CO and solid C are not produced. In addition, this would allow the
20 simplification of the following gas upgrading step before injection into the grid [56].
21

8
To cite : DOI: 10.1016/j.apenergy.2019.114249

1.0 100%

0.8 80%
Molar composition

H₂

CH4 yield (%)


0.6 60% CO₂
H₂O
0.4 40% CH₄
CO

0.2 20% CH₄ yield (10 bar)


CH₄ yield (1 bar)

0.0 0%
200 300 400 500 600 700 800
Temperature (°C)
1
2 Fig. 3. Molar concentrations of CO2, H2, H2O, CH4 and CO at stoichiometric ratio H2/CO2= 4, P =
3 10 bar, and CH4 yield at P = 1 and 10 bar)

4 The main detailed kinetic law was developed in 1989 by Xu and Froment on the Ni/MgAl2O4 (Ni
5 15 wt.%) commercial catalyst, defined for temperatures between 300 to 400 °C and pressures between
6 3 to 10 bar without dilution gas that is closer to the industrial implementation of CO2 methanation [57].
7 A lower temperature (below about 200 °C) may cause the formation of nickel carbonyl, while a high
8 temperature above 500-650 °C may cause carbon deposition (coking), which will deactivate the catalyst
9 [58]. Modelling catalyst reactions in a series of catalytic reactors has been considered, as it avoids the
10 coking and sintering of the catalyst. At least two adiabatic reactors have to be connected in series for a
11 good control of the reaction temperature [59]; the usual approach relying typically on a series of two to
12 five reactors [60]. In this work, the methanation unit consists of four adiabatic multi-tube fixed bed
13 reactors, featuring a multi-stage cooling system and gas recirculation, allowing the outlet specifications
14 to be achieved while minimizing costs. The Gas Hourly Space Velocity (GHSV) was maintained close
15 to 4,000 h-1.
16
17 The corresponding rate equations are therefore given by Eqs.(4)-(7), where the kinetic parameters
18 and their implementation are discussed in the Supporting Information (see SI.3.2).
19
𝑝4𝐻2 ×𝑝𝐶𝑂2
2
𝑘1 𝑝𝐶𝐻4 ×𝑝𝐻 2𝑂

𝑟1 = 3.5 ×
𝐾1 (4)
𝑝𝐻 𝐷𝐸𝑁2
2
𝑝𝐻2 ×𝑝𝐶𝑂
𝑝𝐶𝑂 ×𝑝𝐻2 𝑂 − 2
𝑘2 𝐾2
𝑟2 = × (5)
𝑝 𝐻2 𝐷𝐸𝑁2
𝑝3
𝐻2 ×𝑝𝐶𝑂
𝑘3 𝑝𝐶𝐻4 ×𝑝𝐻2 𝑂 −
𝐾3
𝑟3 = 2.5 × (6)
𝑝𝐻 𝐷𝐸𝑁2
2
20
21 where
𝐾𝐻2 𝑂 𝑝 𝐻 2 𝑂
𝐷𝐸𝑁 = 1 + 𝐾𝐶𝑂 𝑝𝐶𝑂 + 𝐾𝐻2 𝑝𝐻2 + 𝐾𝐶𝐻4 𝑝𝐶𝐻4 + (7)
𝑝 𝐻2
22
23 and with 𝑟𝑖 referring to the rate of reactions 𝑖 (reactions (1) to (3), respectively), 𝑘1 , 𝑘3 , the rate
24 coefficients of reactions (1) and (3) (kmol bar1/2/ kgcat h), respectively, 𝑘2 the rate coefficient of reaction
25 (2) (kmol/bar kgcat h), 𝐾1 , 𝐾3 the equilibrium constant of reactions (1) and (3) (bar2), respectively, 𝐾2
26 the equilibrium constant of reaction (2), 𝐾𝐶𝐻4 , 𝐾𝐶𝑂 , and 𝐾𝐻2 the adsorption constants of reaction for
27 CH4, CO and H2 (bar-1), respectively, 𝐾𝐻2 𝑂 the desorption constant of H2O, and 𝑝 the partial pressure
28 (bar).
29

9
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Process modelling. Fig. 4 shows the process flow sheet of the CO2 methanation unit including the
2 upgrading section.
3

4
5 Fig. 4. Aspen PlusTM flow sheet of the CO2 methanation unit including the upgrading section

6 The CO2 coming from the CO2 capture unit is fed at 5 bar with hydrogen coming at 30 bar from
7 the PEM water wind-based electrolysis, considering the stoichiometric ratio H2/CO2 of 4. A single-stage
8 CO2 compression step is considered technically appropriate to reach the working pressure of 10 bar,
9 while hydrogen is brought to the same pressure. They are then mixed with the recycle stream derived
10 from the upgrading unit and pre-heated to 350 °C before feeding the first adiabatic reactor (REA-1)
11 where the methanation reactions occur. A fraction of the effluent, set at 0.7, is recycled to the inlet of
12 REA-1 to lower the temperature below 650 ºC in order to avoid catalyst sintering [61]. Heat removal is
13 ensured to counteract thermodynamic limitations of the reaction as well as to prevent the catalyst bed
14 from sintering the methanation reactors. The reaction mixture is therefore cooled down to 350 °C after
15 each reactor to obtain high CO2 conversion. The heat of the methanation reactions can be partly used to
16 produce steam to regenerate the scrubbing liquid in the CO2 capture unit. After methanation, water vapor
17 has to be separated from the product gas. The outlet of the last reactor (REA-4) is therefore cooled to a
18 temperature of 25 °C to condense the water. A purity of 99.99 vol% is then achieved in the H2O stream.
19
20 Table 3 presents the main process data equipment for the CO2 methanation unit, while Table 4
21 introduces the specifications for the four multi- tubular adiabatic reactors.
22
23 Table 3. Main process data equipment for the CO2 methanation unit

Parameter Specification
Reactor operating Pressure (bar) 10
Pressure Loss (%) 2
Reactor operating Inlet Temperature (°C) 350
Catalyst Ni/MgAl2O4
Catalyst density (kg/m3) 2,350
Bed density (kg/m3) 790
Bed porosity 0.44
24

25

10
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 4. Main specifications for the four multi- tubular adiabatic reactors at a GHSV of 4,000 h-1

Parameter REA-1 REA-2 REA-3 REA-4


Reactor volume (m3) 15.60 8.32 7.71 7.47
Mass of catalyst (ton) 22.00 5.87 5.44 5.27
2
3 2.1.5 SNG upgrading unit
4
5 The raw-SNG, containing methane, CO2, water vapor and H2, is then upgraded to be injected into
6 the grid where it can be stored, transported, or used as fuel (e.g. for CNG cars). Even though the flash
7 unit removed most of the water (see Fig. 4), the raw-SNG must be further dehydrated to comply with
8 the water dew point requirements of the pipelines, to control corrosion and prevent formation of solid
9 hydrocarbon/water hydrates. Currently, glycol dehydrators are widely used for natural gas dehydration.
10 Despite their advantages, the system operation is complex and requires solvent storage, replacement and
11 disposal. Moreover, the vent streams from the dehydrators are a source of emissions of volatile organic
12 compounds. According to Lin et al., membrane technology appears to be an attractive alternative for
13 natural gas dehydration [62]. To this extent, a commercial Pebax®-based membrane (poly ethylene
14 oxide- based block copolymer) was considered to dry the raw-SNG [63],[64]. The performance is based
15 on literature values, using an estimated normalized permeability constant of 1,000 GPU (1 GPU being
16 3.35 10-10 mol/m2 s Pa) [62],[65], to determine the size required using Eq.(8). In the Aspen PlusTM model,
17 only the retentate/permeate pressures are specified for given feed conditions, as well as the split fractions
18 of one of the stream.
19 The feed gas mixture is compressed in a multi-stage raw-SNG compression step to reach a pressure
20 of 35 bar. A small one-stage system removes 90% of the water in the feed gas, producing a low-pressure
21 permeate gas representing 5-6% of the initial gas flow [66]. The permeate pressure containing the water
22 is set to 1.1 bar [62],[65].
23
𝑝𝑒𝑟𝑚𝑒𝑎𝑡𝑒 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒
𝐴𝑚𝑒𝑚𝑏 = (8)
𝑝𝑒𝑟𝑚𝑒𝑎𝑛𝑐𝑒 × 𝑑𝑟𝑖𝑣𝑖𝑛𝑔 𝑓𝑜𝑟𝑐𝑒
24
25 where the permeate molar flux (mol/s) through the membrane is driven by the partial pressure difference
26 (Pa) of the component of interest H2O, namely the driving force, and the permeance is expressed as the
27 ratio of the permeability with the thickness of membrane, leading to the normalized permeability
28 constant (mol/m2 s Pa) taken to be a measure of the membrane’s ability to permeate gas.
29
30 In addition, excess hydrogen also has to be separated in a gas upgrading downstream of the
31 dehydration membrane, and then fed back to the methanation reaction. A commercial membrane
32 technology (polysolfone-based membrane) for hydrogen separation was chosen to achieve a 90%
33 hydrogen separation [67]. An estimated normalized permeability constant of 100 GPU [66] was
34 considered to approximate the performance and determine the size required using Eq.(8). The stream is
35 therefore expanded to reach a working pressure of 20 bar. The permeate pressure containing the
36 component of interest H2 was set to 1 bar. Almost all H2 along with a certain amount of methane transport
37 across the membrane, was recirculated back to the reactor. The large amount of methane not crossing
38 the membrane forms the retentate at the identical pressure as the feed stream.
39
40 The product stream SNG was then supplied to the natural gas grid usually at elevated pressures
41 from 4 to 70 bar, by using a multistage compressor [68]. The number of stages was determined by a
42 pressure ratio of 4, with equal pressure ratio among all stages. The isentropic efficiency of each stage of
43 75%, and the inter-stage cooling temperature was set at 40 °C. In the frame of this study, it was assumed
44 that the SNG was delivered at 50 bar. Gas specification for gas grid injection in Germany is presented
45 in Table 5. The Wobbe index represents the interchangeability of fuel gases with respect to natural gas,
46 which is usually used to compare the combustion energy output of different composition fuels.
47

11
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 5. Gas specification for gas grid injection in Germany [56],[69],[70]

Parameter Specification
Wobbe Index (MJ/m3) 37.8 – 56.5
CO2 < 6 vol.%
H2 < 5 vol.%
Dust Technically free
water Traces (ppm)
2
3
4 2.1.6 Heat recovery and utilities
5
6 Systematic process-to-process heat recovery through data evaluation, pinch analysis, and
7 optimized heat exchanger network were performed on the system to reach high energy efficiency and
8 minimize utility costs. Appropriate heat exchangers areas were designed using Aspen PlusTM.
9 According to the present design, HX-1 was fed with pressurized boiled feed water at 110 °C and
10 125 bar. HX-2 was placed to extract heat from the outlet of the first reactor (REA-1). The steam
11 temperature and pressure at the outlet of HX-2 was 505 °C and 125 bar, respectively. A steam turbine
12 unit was used to generate process power to cover a portion of the electrical demand. This was expanded
13 to 10 bar with an isentropic efficiency of 0.85, generating 1,055 kWe . After the exit of the turbine, the
14 steam was fed to the condenser, where it released heat and was condensed to 110 °C. The condensate
15 returned to the system with a flow rate equal to 7.5 t/h.
16 As previously mentioned, five additional heat exchangers (from HX-3 to HX-7) were implemented
17 to cool down the outlet of each reactor to the working temperature of 350 °C, to obtain high CO2
18 conversion. They were fed with boiled feed water at 110 °C and 8 bar, to generate a medium-pressured
19 steam at 170 °C and 8 bar. This was done so that it could be brought to the reboiler heat duty of the CO2
20 capture unit, which requires heat at 140 °C and 5 bar.
21
22 2.2 Techno-economic analysis
23 2.2.1 Technical analysis
24
25 Technical indicators, including the mass balance of individual inputs and outputs, and the utilities
26 demand (mainly heat and electricity duties), are direct results from the process modelling using the
27 software Aspen HysysTM and Aspen PlusTM.
28
29 In addition, the methanation efficiency is calculated according to Eq.(9) using the general approach
30 suggested by Salomone et al. [9]:
31
𝜈𝐶𝐻4 × 𝑋𝐶𝑂2 × 𝑀𝐶𝐻4 × 𝐿𝐻𝑉𝐶𝐻4
𝜂𝐶𝐻4 = (9)
𝜈𝐻2 × 𝑀𝐻2 × 𝐿𝐻𝑉𝐻2
32
33 where 𝑋𝐶𝑂2 is the conversion of CO2 within the methanation unit, 𝑀𝐶𝐻4 and 𝑀𝐻2 are the CH4 and H2
34 molar weights (kg/kmol), 𝐿𝐻𝑉𝐶𝐻4 and 𝐿𝐻𝑉𝐻2 are the CH4 and H2 lower heating values (MJ/kg), 𝜈𝐶𝐻4
35 and 𝜈𝐻2 are the CH4 and H2 stoichiometric parameters, respectively. The conversion of CO2 within the
36 methanation unit is assumed to be superior to 98.9% to achieve the SNG quality [9].
37 To evaluate the process efficiency of the PtG plant, the PEM electrolyzer, methanation unit,
38 compressors, and heater are considered, according to Eq.(10) [9]:
39
𝑛𝐶𝐻4 × 𝑀𝐶𝐻4 × 𝐿𝐻𝑉𝐶𝐻4
𝜂𝑃𝑡𝐺 = (10)
𝐸𝑃𝐸𝑀 + 𝐸𝑐𝑜𝑚𝑝 + 𝐸ℎ𝑒𝑎𝑡
40
41 where 𝐸𝑃𝐸𝑀 is the electrical energy used by the electrolyzer (MJ), 𝐸𝑐𝑜𝑚𝑝 the energy spent for the
42 compression (MJ), and 𝐸ℎ𝑒𝑎𝑡 represents the total required heat (MJ).
43

12
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Finally, the overall energy utilization factor is calculated with Eq.(11).


2
𝐸ℎ𝑒𝑎𝑡
𝛷𝑃𝑡𝐺 = 1 − (11)
𝐸𝑃𝐸𝑀 + 𝐸𝑐𝑜𝑚𝑝
3
4 The following parameters are also required for the produced SNG stream: (i) composition, essential
5 to achieve the gas requirements for injection into the networks (i.e. percentage of hydrogen); (ii) the
6 Low Heating Value (LHV) and/or High Heating Value (HHV); and (iii) the Wobbe index.
7
8 2.2.2 Economic analysis
9
10 Both operating and investment costs were evaluated. Most of the capital investment for the
11 equipment was generated directly via AspenTech Economic AnalyzerTM. In addition, the cost of several
12 individual pieces of equipment was approximated when the cost of a similar item of a different size or
13 capacity was known, using expression (12). The costs were estimated with a nominal accuracy of ±
14 30%.
15
𝑆 𝛼
𝐶 = 𝐶0 ( ) (12)
𝑆0
16
17 where the incremental cost 𝐶 decreases with larger capacities 𝑆, 𝐶0 and 𝑆0 being the reference
18 characteristics values. The exponent 𝛼 takes the economy of scale effect into account, which is often
19 approximated to 0.6. The results were then updated from the original cost index to the year 2018 using
20 the annual chemical engineering plant cost index CEPCI (CEPCI2018 = 603.1) to account for the price
21 development effect (Chemical Engineering magazine for latest values).
22
23 The total capital investment (𝑇𝐶𝐼) was determined using the factorial methodology, according to
24 the recommended ratio factors for fluid processing plant by Peter and Timmerhaus [71], and the capital
25 investment (𝐼𝐸 ) of the equipment (see Eq.(13)).
26
𝑇𝐶𝐼 = 𝐼𝐸 × (1 + ∑𝑛𝑖=1 𝑅𝐹𝑖 ) (13)
27
28 where 𝑅𝐹 is the ratio factor for direct, indirect and working capital, 𝑖 being the items summarized in
29 Table 6.

13
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 6. Estimation of the total capital investment (𝑻𝑪𝑰) [71], [72]

Type Item (𝒊) Ratio Factor (𝑹𝑭)


1 Purchased equipment (delivered) 1.00
2 Purchased equipment installation 0.47
3 Piping 0.36
4 Instrumentation and Controls 0.68
Direct cost
5 Electrical systems 0.11
6 Buildings (including services) 0.18
7 Yard improvements 0.10
8 Service facilities 0.70
9 Engineering and supervision 0.33
10 Construction expenses 0.41
11 Legal expenses 0.04
Indirect cost
12 Contractor’s fee 0.22
13 Contingency 0.44
14 Working Capital 0.89
2
3 The total product cost (𝑇𝑃𝐶) is calculated by Eq.(14).
4
𝑇𝑃𝐶 = ∑𝑛𝑗=1 𝐶𝑗 (14)
5
6 with 𝐶𝑗 the cost of item 𝑗 based on Table 7.
7
8 Table 7. Economic assumptions for operating cost estimation (𝑻𝑷𝑪)

Item (𝒋) Assumptions


Solvent: 1 €/kg(1); Catalyst: 90 €/kg(2); Process water: 1
€/ton(3); Demineralized water: 3 €/ton(4); Renewable H2:
1 Raw material
onsite production(5); SNG selling price: 21 €/ton(6); O2
selling price: 87.4 €/ton(7); CO2 credit tax: 25.2 €/ton(8)
Cooling water: 0.025 €/ton(9); Electricity: 40 €/MWh(10);
2 Utilities
Steam: self-produced(11)
Operating labor and 60 labors/shift, 3 shift/day, 54,000 € per labor per year(12)
supervisory labor + 20% of operating labor
Operation and Maintenance and repairs 6% of capital investment
3
maintenance
Operating supplies 15% of Maintenance and repairs
Laboratory charges 15% of operating labor
Annual depreciation cost, with an assumption of a 20-year
4 Depreciation
recovery period and 0 € salvage value, linear
5 General expenses 15% of total product cost
9 (1)
Amine solvent considered: PZ 30% + MDEA 10% (+ 60% water): 2.16 €/kg PZ × 0.3 + 1.7 €/kg MDEA × 0.1 + 0.001 €/kg H2O × 0.6 =
10 0.818 €/kg ≈ 1 €/kg solvent cost estimated from de Medeiros et al. [73].
11 (2)
Catalyst cost considered from own estimation.
12 (3)
Process water cost considered from Michailos et al. [74].
13 (4)
Demineralized water cost considered from Li et al. [75].
14 (5)
Renewable H2 considered produced onsite. Only the cost of demineralized water is taken into account in the calculation of 𝑇𝑃𝐶.
15 (6)
SNG selling price estimated from natural gas price (≈ 4.56 €/1000 ft3 in April 2019).
16 (7)
Oxygen selling price from Michailos et al. [74].
17 (8)
CO2 credit tax (€/ton) in April 2019, according to European Emission Allowances (EUA).
18 (9)
Cooling water cost considered from Michailos et al. [74].
19 (10)
Electricity considered from renewable energy (wind power). Cost considered from Glenk & Reichelstein [76] supported by www.eex.com.
20 (11)
Steam considered self-produced. Only the cost of process water is taken into account in the calculation of 𝑇𝑃𝐶.
21 (12)
Operating labor and supervisory labor costs considered from Zhang et al. [72].

14
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 3 Results and discussions


2
3 3.1 Process performances
4 3.1.1 CO2 capture unit results
5
6 Table 8 summarizes the CO2 capture unit simulation results in terms of gas phase compositions.
7
8 Table 8. Composition of the gas to be treated, washed-gas and produced CO2 (mol frac. basis)

Component Gas to treat Washed-gas Produced CO2


N2 64.7 % 69.9 % 123 ppm
CO2 20.4 % 2.2 % 98.0 %
O2 8.6 % 9.3 % 29 ppm
H2O 6.2 % 18.5 % 1.98 %
CO 1,330 ppm 100 ppm 0.4 ppm
SO2 111 ppm 112 ppm 43 ppm
NO 474 ppm 512 ppm 3 ppm
NO2 2 ppm 2 ppm 1 ppm
MDEA 0 Traces Traces
PZ 0 Traces Traces
Total mol flow (kmol/h) 1,124 1,040 211
Total mass flow (t/h) 34.85 27.90 9.16
9
10 An absorption ratio of 90 mol.% leads to the production of 211 kmol/h of CO2 (9.16 t/h) with 98
11 mol.% purity. In addition to nitrogen (123 ppm) and oxygen (29 ppm), the produced CO2 flow still
12 contains some water (around 2 mol.%) and some other components initially present in the flue gas,
13 namely SO2 (< 50 ppm), NO (< 5 ppm), NO2 (1 ppm) and CO (< 0.5 ppm). As a result of the water-wash
14 sections, the amines (MDEA and PZ) are only present as traces in the washed-gas and the produced CO2
15 flow.
16 Concerning the energy consumptions, the main energy consuming unit is the reboiler of the
17 regeneration column (5.75 MWth), while the other thermal demands are for cooling (mainly condenser
18 cooling energy, -1.45 MWth, and lean solution cooling, -2.80 MWth). The cooling energy of the absorber
19 intercooling is less significant (-0.72 MWth) and the water-wash cooling energy is very low (-0.01
20 MWth). Regarding the electrical demands, while the total pumping energy is low (0.04 MWe ), the
21 electrical consumption of the compressor used in the RVC configuration is not negligible (0.91 MWe ).
22 These results were obtained for the optimum operating parameters given in Table 9, where the specific
23 energy consumption per ton of captured CO2 is also provided.

15
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 9. Optimum operating parameters and specific energy consumptions of the CO2 capture unit

Indicator Specification Value


(L/G)vol,opt (m³/m³) 5.19 10-3
Lopt (m³/h) 128
Intercooling temp. (°C) 40
Operating conditions Intercooling stage 8
Flash ∆p (bar) 4
𝛼𝐶𝑂2 ,𝑟𝑖𝑐ℎ (mol/mol) 0.98
𝛼𝐶𝑂2 ,𝑙𝑒𝑎𝑛 (mol/mol) 0.33
Energy consumption 𝐸𝑟𝑒𝑔𝑒𝑛 (GJ/tCO2) 2.28
2
3 The specific solvent regeneration energy (2.28 GJ/tCO2) is drastically reduced in comparison with
4 conventional CO2 capture processes using MEA 30 wt.% as a solvent (generally estimated around 3.5
5 GJ/tCO2 [26] and being precisely calculated at 3.36 GJ/tCO2 using the present simulation model), which
6 highlights the interest of using an advanced CO2 capture process such as the one considered in the
7 present study. All the simulation results in relation with the optimization of the CO2 capture unit are
8 provided in the Supporting Information (see SI.1), where comparisons are also performed in terms of
9 equivalent works and utilities costs both for conventional process configurations (without ICA) and for
10 three different solvents (MEA 30 wt.%, PZ 40 wt.% and MDEA 10 wt.% + PZ 30 wt.%).
11
12 3.1.2 CO2 methanation and SNG upgrading units’ results
13
14 The CO2 methanation unit simulation results in terms of outlet gas compositions are summarized
15 in Table 10.
16
17 Table 10. Composition of the outlet streams (mol frac. Basis)

Raw-
Component Inlet REA-1 REA-2 REA-3 REA-4 SNG
SNG
CO2 19.9 % 7.5 % 4.3 % 2.4 % 1.5 % 3.7 % 3.7 %
H2 79.9 % 31.8 % 19.0 % 11.5 % 8.1 % 21.3 % 3.4 %
H2O 0.2 % 40.5 % 51.1 % 57.4 % 60.3 % 0.3 % Traces
CH4 0% 20.2 % 25.6 % 28.7 % 30.2 % 74.8 % 92.9 %
CO 0% Traces - - - - -
Total mol flow (kmol/h) 1,038.12 745.23 692.67 664.94 653.05 268.54 214.22
Total mass flow (t/h) 10.84 10.79 10.79 10.79 10.79 3.78 3.60
18
19 Based on the simulations and the present process design, material and energy streams were
20 estimated and summarized in Table 11.

16
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 11. Material and energy consumptions of the PtG plant

Indicator Specification Value


H2 (t/h) 1.65
CO2 methanation inputs
CO2 (t/h) 9.09
CO2 compressor (MWe ) 0.14
Recycle compressors (MWe ) 0.19
Membrane compressor (MWe ) 0.36
Energy consumptions
SNG compressor (MWe ) 0.25
Pump (MWe ) 0.07
Turbine (MWe ) -1.06
SNG (t/h) 3.60
PtG outputs SNG Wobbe Index (MJ/m3) 46.72
Total Output HHV (MW) 50.4
Methanation efficiency (𝜂𝐶𝐻4 ) 83.4
Efficiency PtG plant HHV (𝜂𝑃𝑡𝐺 ) 72.6
Overall energy utilization factor (𝛷𝑃𝑡𝐺 ) 75.7
2
3 In order to meet the pipeline quality standards, the CH4 mole fraction in the raw SNG of about 70%
4 has to be increased to more than 95% in order to meet the natural gas grid requirements. Table 12
5 presents the technical features of the utilized membranes.
6
7 Table 12. Technical features of the membranes system

Parameter Dehydration membrane H2 separation membrane


Membrane area (m2) 62 244
Feed pressure (bar) 35 20
Permeate pressure (bar) 1.1 1
Temperature (°C) 40 90
Permeance (GPU) 1,000 100
Compressor (MWe ) 0.36 -
8
9 A membrane’s size of 62 m2 was therefore required to achieve a 90% water separation, using
10 Eq.(8). To achieve a 90% H2 separation, the size of the second membrane was estimated to 244 m2. The
11 electrical consumption of the compressor used to reach the working pressure of the dehydration
12 membrane is of importance (0.36 MWe ).
13
14 In addition, the SNG was brought to 50 bar to meet the pressure pipelines, leading to a consumption
15 of 0.25 MWe . It is worth mentioning that for some local or regional grid injections, this elevated pressure
16 is not necessary, as these pipelines mostly operate in the pressure range of 4 to 40 bar.
17

17
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 3.1.3 Process integration results


2
3 The methanation efficiency was estimated at 83.4 %, while the process efficiency of the PtG plant
4 (𝜂𝑃𝑡𝐺 ) equaled 65.5% (72.6% on HHV basis), based on Eq.(9) and (10), respectively, with a
5 comprehensive usage of excess heat and technical assumptions. The overall energy utilization factor
6 (𝛷𝑃𝑡𝐺 ) reached 75.7%. A Wobbe Index of 46.72 MJ/m3 was calculated, which is in the range of Wobbe
7 Indices in Germany (see Table 5). These results are summarized in Table 11.
8
9 Stream data (temperature, composition, heat capacities, etc.) computed by Aspen PlusTM are
10 available in the Supporting Information (see SI.4). A minimum pinch point temperature of 30 °C was
11 defined for the heat exchanger with gaseous streams [77].
12 The total recovered heat from the plant amounts to about 13.25 MWth, which is available between
13 25 to 615 °C (see Fig. 5). A fraction of the heat recovered at 170 °C can be brought to the reboiler of
14 the CO2 capture unit, accounting for 5.75 MWth, so that the heat integration compensates the hot utility
15 requirement of the CO2 capture unit.
16 In addition, the electricity produced in the steam turbine was used to cover internal demand (see
17 Table 11). Thus, no external electrical requirement was required for both the CO2 methanation and SNG
18 upgrading units. The total external electrical requirement of the whole suggested process designed is
19 therefore linked to the compressor used in the RVC configuration, equal to 0.90 MWe .

600

500
Temperature (°C)

400

300

200

100

0
0 2 4 6 8 10 12 14
Heat load (MW)

Hot composite curve Cold composite curve


20
21 Fig. 5. Hot and Cold Composite curves

22 3.2 Economic performances


23
24 Based on the simulation results, the capital investment and operating costs of the integrated process
25 plant were evaluated. Table 13 lists the estimated capital investment for the equipment.
26
27 Table 13. Cost estimation of main equipment (M€)

Main equipment CO2 Capture CO2 methanation Gas Upgrading


Columns (absorber/stripper) 1.24 - -
Reactors - 3.50 -
Compressors 3.36 2.28 3.91
Turbine - 0.45 -
Heat Exchangers (heaters/coolers) 2.40 9.35 0.15
Flash tanks 0.18 0.09 -
Membranes - - 0.68
Total Purchased equipment cost (delivered) 7.18 16.22 4.74

18
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 14 presents the main costs for all the units.


2
3 Table 14. Estimation of the total capital investment (M€)

Main cost H2 production CO2 Capture CO2 methanation Gas Upgrading


Total Direct Cost 387.36 (1) 25.84 58.40 17.05
Total Indirect Cost 154.94 10.34 23.36 6.82
Fixed Capital Investment (𝐹𝐶𝐼) 542.30 36.17 81.76 23.88
Total Capital Investment (𝑇𝐶𝐼) 638.07 42.56 96.20 28.09
4 (1)
Calculation based on specific electrical power of approximatively 55.6 kWh per kg H 2 and an investment cost of 1,200 € per kW
5 (installed) for the production of 40 ton per day of renewable H2 [49],[76], and Table 6 to estimate the Total Direct Cost.

6 The 𝑇𝐶𝐼 of the integrated process was therefore evaluated to 166.85 M€ (excluding the capital
7 investment of H2 production), where the CO2 conversion unit itself represents 58%, mainly due to the
8 complex network of heat exchangers that had been implemented to integrate the heat. The final 𝑇𝐶𝐼 was
9 estimated at 804.92 M€, where about 79% was dedicated to the renewable production of H2. Regarding
10 the total product cost, it was estimated at 61.60 M€ per year. Fig. 6 displays the operational expenses
11 distribution, based on the previously mentioned assumptions (see Table 7). The expected indirect
12 revenues were 407.9 € per ton SNG, considering SNG, oxygen and CO2 trading.
13

Operational Expenses (€/ton SNG)


-2,000 -1,600 -1,200 -800 -400 0 400

SNG selling

O₂ selling

CO₂ credit tax

Hydrogen Production

CO₂ Capture Unit


CO₂ Methanation Unit
(SNG upgrading Unit incl.)
Amortized CAPEX

Additional expenses

TOTAL

14
15 Fig. 6. Operational expenses in € per ton SNG

16 At present, the estimated cost of SNG is therefore 1,587 €/ton SNG, about 115 €/MWh, which is
17 in the cost range evaluated by Guilera et al. [2]. It is about 3.5 times higher than fossil natural gas (33
18 €/MWh in Germany [2]). The annual operating costs amount to around 7% of the investment, which is
19 in accordance with the values reported for catalytic processes [78]. It is worth noting that the amortized
20 CAPEX of the electrolyzers is taken into account in “Hydrogen production”.
21
22 A sensitivity analysis was carried out regarding the parameters that most influence the SNG cost,
23 derived from Table 15. One input variable at a time was changed to its low or high value, while the other
24 variables were maintained at their base values. Fig. 7 illustrates the sensitivity analysis.
25

19
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Table 15. Input variable for the sensitivity analysis

Input variables Low value Base value High value


Electricity price(1) (€/MWh) 0 40 120
Electrolyser efficiency(2) (%) 65 75 85
Additional expenses & CAPEX(3) -30% Nominal +30%
CO2 credit tax(4) (€/ton) 0 25.2 50.4
Oxygen price(5) (€/ton) 43.7 87.4 131.1
2 (1)
Electricity ranging from 0 €/MWh (free renewable electricity) to 120 €/MWh (own assumption).
3 (2)
Electrolyser efficiency commonly considered in the literature [74],[76].
4 (3)
Additional expenses and CAPEX accuracy with a ± 30% fluctuation.
5 (4)
Electricity considered CO2 credit tax in April 2019, according to European Emission Allowances (EUA).
6 (5)
Oxygen selling price from Michailos et al. [74], with a ± 50% fluctuation.
7

Electricity price

Additonal expenses & CAPEX

Electrolyser efficiency

O₂ selling price

CO₂ credit tax

0 50 100 150 200 250 300


SNG cost (€/MWh)

Low value High value


8
9 Fig. 7. Sensitivity analysis on the cost of SNG in €/MWh (base cost 115 €/MWh SNG)

10 The electricity price therefore has a huge influence on the cost of SNG, especially due to the fact
11 that H2 production derived directly from it, represents about 70% of the cost of SNG. The additional
12 expenses and CAPEX, the electrolyzer efficiency, and the O2 selling price have an influence higher than
13 10% on the final SNG cost. In particular, the income obtained from the sale of oxygen in the PtG plant
14 represents an interesting economic incentive, especially in the current context of low prices for carbon
15 emissions trading (CO2 credit tax) that have an almost negligible effect.

16 4 Conclusion and outlook


17
18 The present work has assessed the production of SNG from industrial CO2 in the context of PtG
19 technology. The PtG plant was implemented in AspenTech’s software, both in Aspen HYSYSTM and
20 Aspen PlusTM, to extract mass and energy balances in order to investigate its economic viability. An
21 advanced CO2 capture process was proposed to treat 10% of a BAT cement plant’s flue gas. Together
22 with renewable H2, they were converted to SNG through methanation reaction, producing 0.40 ton SNG
23 per ton of captured CO2. In the suggested process design, a dry methane mole fraction of 92.9 mol.% in
24 the final SNG was achieved, together with a CO2 and H2 content of 3.7 mol.% and 3.4 mol.%,
25 respectively, which are the most critical aspects of the SNG quality and compatibility.
26 Systematic process-to-process analysis was investigated to reach high energy efficiency and
27 minimize utility costs. Thermal integration led to an overall system HHV-based efficiency equal to
28 72.6%.
29 As seen in this work, SNG is currently not competitive with natural gas. The estimated cost of SNG
30 was about 115 €/MWh, about 3.5 times higher than fossil natural gas. Costs are mainly driven by the
31 high initial investment costs, as well as the production of H2, which depends directly on the price of
32 renewable electricity. Thus, a reduction of the electrolyzer investment costs combined with a higher
33 electrolyzer efficiency and higher revenues, could result in a profitable business case for large-scale
34 deployment of this PtG technology. In future scenarios, the cost of SNG can be therefore reduced.

20
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 Specific costs for PEM technologies are for instance expected to fall from about 1,200 €/kW in 2017 to
2 500 €/kW in 2050 [40]. Furthermore, when applying a time-of-use retail electricity price of
3 approximately 10 €/MWh, the costs of SNG decrease to about 40 €/MWh, which is really close to the
4 current price of fossil natural gas. Several studies also point out that with ongoing trends in 10 to
5 20 years, PtG might be economically profitable [33]. Moreover, in regions where a natural gas
6 infrastructure exists, the existing network for grid gas injection is a major advantage of the PtG
7 technology. By utilizing excess energy to produce hydrogen via water electrolysis, energy can be stored
8 and distributed in the existing system for use when and where it is needed. Despite the fact that the use
9 of (excess) renewable energy introduces a seasonal and daily component into H2 production, Gahleitner
10 highlighted that the use of such fluctuating power sources is satisfying, and is currently in use in a lot of
11 PtG pilot plants [79].
12
13 In the context of CCUS, the main drivers for methanation deployment are limited and/or unwanted
14 CO2 storage, high CO2 reduction targets, and high renewable energy availability and penetration in the
15 market. Additionally, several actions could be taken to foster the deployment of PtG technologies. A
16 key action is the development and deployment of cheap low carbon electricity, which is necessary to
17 reduce the costs for PtG. The transition to low-carbon depends on the cost for conventional fossil
18 choices, such as the gas price. Thus, additional taxes on fossil resources could promote a shift to PtG
19 technologies. Direct subsides is also a measure that can be used to improve the business case for private
20 investors. Creating a green hydrogen market would also be necessary, giving a premium in the selling
21 price of hydrogen. Finally, to overcome the seasonal component of PtG, nuclear, geothermal, biomass,
22 and hydro energy are necessary, also increasing the flexibility of the whole system.
23
24 By applying these actions, combined with a policy penalizing new investment in industries using
25 fossil resources, the suggested process will likely play a certain role for the transition to a renewable
26 network, integrating large fractions of renewables that require balancing power and seasonal energy
27 storage, and introducing considerable flexibility into the energy system. Furthermore, additional energy
28 and environmental policy challenges can be solved by the enlargement of the percentage of alternative
29 fuels in the mobility or heating sector. In a longer-term perspective, with markets characterized by cheap
30 electricity and high fuel prices, PtG shall gradually eliminate fossil fuels, transitioning to a low carbon
31 economy. PtG may also have higher acceptance in society, denoting advantages over the existing
32 alternative storage technologies as well as over network expansion, representing a non-negligible
33 component in the future prospects of the technology.
34
35

36 Acknowledgements
37
38 The authors gratefully acknowledge the European Cement Research Academy (ECRA) for its
39 technical and financial support. The authors thank Marc Frère for his help and scientific guidance.
40

41 References
42
43 [1] Zappa W, Junginger M, van den Broek M. Is a 100% renewable European power system feasible
44 by 2050? Appl Energy 2019;233–234:1027–50. doi:10.1016/j.apenergy.2018.08.109.
45 [2] Guilera J, Ramon Morante J, Andreu T. Economic viability of SNG production from power and
46 CO2. Energy Convers Manag 2018;162:218–24. doi:10.1016/j.enconman.2018.02.037.
47 [3] Wang L, Rao M, Diethelm S, Lin T-E, Zhang H, Hagen A, et al. Power-to-methane via co-
48 electrolysis of H2O and CO2: The effects of pressurized operation and internal methanation. Appl
49 Energy 2019;250:1432–45. doi:10.1016/j.apenergy.2019.05.098.
50 [4] Zu MY, Liu PF, Wang C, Wang Y, Zheng LR, Zhang B, et al. Bimetallic Carbide as a Stable
51 Hydrogen Evolution Catalyst in Harsh Acidic Water. ACS Energy Lett 2018;3:78–84.
52 doi:10.1021/acsenergylett.7b00990.
53 [5] Nagasawa K, Davidson FT, Lloyd AC, Webber ME. Impacts of renewable hydrogen production

21
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 from wind energy in electricity markets on potential hydrogen demand for light-duty vehicles.
2 Appl Energy 2019;235:1001–16. doi:10.1016/j.apenergy.2018.10.067.
3 [6] Mattioda RA, Antonucci V, Azzaro-Pantel C, Cao H, Cellura M, Felice F De, et al. Hydrogen
4 Economy: Supply Chain, Life Cycle Analysis and Energy Transition for Sustainability. 1st
5 Editio. Elsevier Ltd.; 2017.
6 [7] Quarton CJ, Samsatli S. Power-to-gas for injection into the gas grid: What can we learn from
7 real-life projects, economic assessments and systems modelling? Renew Sustain Energy Rev
8 2018;98:302–16. doi:10.1016/j.rser.2018.09.007.
9 [8] Ghaib K, Ben-Fares FZ. Power-to-Methane: A state-of-the-art review. Renew Sustain Energy
10 Rev 2018;81:433–46. doi:10.1016/j.rser.2017.08.004.
11 [9] Salomone F, Giglio E, Ferrero D, Santarelli M, Pirone R, Bensaid S. Techno-economic
12 modelling of a Power-to-Gas system based on SOEC electrolysis and CO2 methanation in a RES-
13 based electric grid. Chem Eng J 2018:1–20. doi:10.1016/j.cej.2018.10.170.
14 [10] Toro C, Sciubba E. Sabatier based power-to-gas system: Heat exchange network design and
15 thermoeconomic analysis. Appl Energy 2018;229:1181–90.
16 doi:10.1016/j.apenergy.2018.08.036.
17 [11] Gutiérrez-Martín F, Rodríguez-Antón LM. Power-to-SNG technologies by hydrogenation of
18 CO2 and biomass resources: A comparative chemical engineering process analysis. Int J
19 Hydrogen Energy 2018:1–10. doi:10.1016/j.ijhydene.2018.09.168.
20 [12] Lehner M, Tichler R, Steinmüller H, Koppe M. Power-to-Gas: Technology and Business Models.
21 Cham: Springer International Publishing; 2014. doi:10.1007/978-3-319-03995-4.
22 [13] Aresta M, Dibenedetto A, Angelini A. The changing paradigm in CO2 utilization. J CO2 Util
23 2013;3–4:65–73. doi:10.1016/j.jcou.2013.08.001.
24 [14] Benhelal E, Zahedi G, Shamsaei E, Bahadori A. Global strategies and potentials to curb CO2
25 emissions in cement industry. J Clean Prod 2013;51:142–61. doi:10.1016/j.jclepro.2012.10.049.
26 [15] Corcoran R. Technical Report - ECRA Project - Report about CO2 reuse from cement production
27 / MeOH and Methane Synthesis. Dublin: 2013.
28 [16] Meunier N, Laribi S, Dubois L, Thomas D, De Weireld G. CO2 capture in cement production
29 and re-use: First step for the optimization of the overall process. Energy Procedia 2014;63:6492–
30 503. doi:10.1016/j.egypro.2014.11.685.
31 [17] Granados DA, Chejne F, Mejía JM. Oxy-fuel combustion as an alternative for increasing lime
32 production in rotary kilns. Appl Energy 2015;158:107–17. doi:10.1016/j.apenergy.2015.07.075.
33 [18] Carrasco-Maldonado F, Spörl R, Fleiger K, Hoenig V, Maier J, Scheffknecht G. Oxy-fuel
34 combustion technology for cement production - State of the art research and technology
35 development. Int J Greenh Gas Control 2016;45:189–99. doi:10.1016/j.ijggc.2015.12.014.
36 [19] Ditaranto M, Bakken J. Study of a full scale oxy-fuel cement rotary kiln. Int J Greenh Gas Control
37 2019;83:166–75. doi:10.1016/j.ijggc.2019.02.008.
38 [20] Voldsund M, Gardarsdottir SO, De Lena E, Pérez-Calvo JF, Jamali A, Berstad D, et al.
39 Comparison of technologies for CO2 capture from cement production—Part 1: Technical
40 evaluation. Energies 2019;12. doi:10.3390/en12030559.
41 [21] IEAGHG. Further Assessment of Emerging CO2 Capture Technologies for the Power Sector and
42 their Potential to Reduce Costs. 2019.
43 [22] Abanades JC, Arias B, Lyngfelt A, Mattisson T, Wiley DE, Li H, et al. Emerging CO2 capture
44 systems. Int J Greenh Gas Control 2015;40:126–66. doi:10.1016/j.ijggc.2015.04.018.
45 [23] Younas M, Sohail M, Kong LL, Bashir MJK, Sethupathi S. Feasibility of CO2 adsorption by
46 solid adsorbents: a review on low-temperature systems. Int J Environ Sci Technol 2016;13:1839–
47 60. doi:10.1007/s13762-016-1008-1.
48 [24] Criado YA, Arias B, Abanades JC. Calcium looping CO2 capture system for back-up power
49 plants. Energy Environ Sci 2017;10:1994–2004. doi:10.1039/c7ee01505d.
50 [25] Idem R, Supap T, Shi H, Gelowitz D, Ball M, Campbell C, et al. Practical experience in post-
51 combustion CO2 capture using reactive solvents in large pilot and demonstration plants. Int J
52 Greenh Gas Control 2015;40:6–25. doi:10.1016/j.ijggc.2015.06.005.
53 [26] Van Der Spek M. Methodological improvements to ex-ante techno-economic modelling and
54 uncertainty analysis of emerging CO2 capture technologies. Utrecht University, 2017.
55 [27] Singh P, Swaaij WPM Van, (Wim) Brilman DWF. Energy Efficient Solvents for CO2 Absorption

22
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 from Flue Gas: Vapor Liquid Equilibrium and Pilot Plant Study. Energy Procedia 2013;37:2021–
2 46. doi:10.1016/j.egypro.2013.06.082.
3 [28] Le Moullec Y, Neveux T, Al Azki A, Chikukwa A, Hoff KA. Process modifications for solvent-
4 based post-combustion CO2 capture. Int J Greenh Gas Control 2014;31:96–112.
5 doi:10.1016/j.ijggc.2014.09.024.
6 [29] Dubois L, Thomas D. Comparison of various configurations of the absorption-regeneration
7 process using different solvents for the post-combustion CO2 capture applied to cement plant
8 flue gases. Int J Greenh Gas Control 2018;69:20–35. doi:10.1016/j.ijggc.2017.12.004.
9 [30] Rönsch S, Schneider J, Matthischke S, Schlüter M, Götz M, Lefebvre J, et al. Review on
10 methanation - From fundamentals to current projects. Fuel 2016;166:276–96.
11 doi:10.1016/j.fuel.2015.10.111.
12 [31] Chauvy R, Meunier N, Thomas D, De Weireld G. Selecting emerging CO2 utilization products
13 for short- to mid-term deployment. Appl Energy 2019;236:662–80.
14 doi:10.1016/j.apenergy.2018.11.096.
15 [32] Artz J, Müller TE, Thenert K, Kleinekorte J, Meys R, Sternberg A, et al. Sustainable Conversion
16 of Carbon Dioxide: An Integrated Review of Catalysis and Life Cycle Assessment. Chem Rev
17 2018;118:434–504. doi:10.1021/acs.chemrev.7b00435.
18 [33] Lewandowska-Bernat A, Desideri U. Opportunities of power-to-gas technology in different
19 energy systems architectures. Appl Energy 2018;228:57–67.
20 doi:10.1016/j.apenergy.2018.06.001.
21 [34] European Power to Gas Platform. Online n.d. http://europeanpowertogas.com/ (accessed March
22 4, 2019).
23 [35] Bailera M, Lisbona P, Romeo LM, Espatolero S. Power to Gas projects review: Lab, pilot and
24 demo plants for storing renewable energy and CO2. Renew Sustain Energy Rev 2017;69:292–
25 312. doi:10.1016/j.rser.2016.11.130.
26 [36] Wulf C, Linßen J, Zapp P. Review of power-to-gas projects in Europe. Energy Procedia
27 2018;155:367–78. doi:10.1016/j.egypro.2018.11.041.
28 [37] Su X, Xu J, Liang B, Duan H, Hou B, Huang Y. Catalytic carbon dioxide hydrogenation to
29 methane: A review of recent studies. J Energy Chem 2016;25:553–65.
30 doi:10.1016/j.jechem.2016.03.009.
31 [38] Miguel CV, Soria MA, Mendes A, Madeira LM. A sorptive reactor for CO 2 capture and
32 conversion to renewable methane. Chem Eng J 2017;322:590–602.
33 doi:10.1016/j.cej.2017.04.024.
34 [39] Eveloy V, Gebreegziabher T. A Review of Projected Power-to-Gas Deployment Scenarios.
35 Energies 2018;11. doi:10.3390/en11071824.
36 [40] Thema M, Bauer F, Sterner M. Power-to-Gas: Electrolysis and methanation status review.
37 Renew Sustain Energy Rev 2019;112:775–87. doi:10.1016/j.rser.2019.06.030.
38 [41] Becker WL, Penev M, Braun RJ. Production of Synthetic Natural Gas From Carbon Dioxide and
39 Renewably Generated Hydrogen: A Techno-Economic Analysis of a Power-to-Gas Strategy. J
40 Energy Resour Technol 2018;141:021901. doi:10.1115/1.4041381.
41 [42] Morosanu EA, Saldivia A, Antonini M, Bensaid S. Process Modeling of an Innovative Power to
42 LNG Demonstration Plant. Energy and Fuels 2018;32:8868–79.
43 doi:10.1021/acs.energyfuels.8b01078.
44 [43] Meunier N, Chauvy R, Mouhoubi S, Thomas D, De Weireld G. Alternative production of
45 methanol from industrial CO2. Renew Energy 2020;146:1192–203.
46 doi:10.1016/j.renene.2019.07.010.
47 [44] Song Y, Chen C-C. Symmetric Electrolyte Nonrandom Two-Liquid Activity Coefficient Model.
48 Ind Eng Chem Res 2009;48:7788–97. doi:10.1021/ie9004578.
49 [45] Peng D-Y, Robinson DB. A New Two-Constant Equation of State. Ind Eng Chem Fundam
50 1976;15:59–64. doi:10.1021/i160057a011.
51 [46] Laribi S, Dubois L, De Weireld G, Thomas D. Study of the post-combustion CO2 capture process
52 by absorption-regeneration using amine solvents applied to cement plant flue gases with high
53 CO2 contents. Int J Greenh Gas Control 2019;90:102799. doi:10.1016/j.ijggc.2019.102799.
54 [47] Muradov N. Liberating Energy from Carbon: Introduction to Decarbonization. vol. 22. New
55 York, NY: Springer New York; 2014. doi:10.1007/978-1-4939-0545-4.

23
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 [48] IRENA. Hydrogen from renewable power - Technology outlook for the energy transition. Abu
2 Dhabi: 2018. doi:978-92-9260-077-8.
3 [49] Schmidt O, Gambhir A, Staffell I, Hawkes A, Nelson J, Few S. Future cost and performance of
4 water electrolysis: An expert elicitation study. Int J Hydrogen Energy 2017;42:30470–92.
5 doi:10.1016/j.ijhydene.2017.10.045.
6 [50] Dubois L, Thomas D. Optimization of the post-combustion CO2 capture process applied to
7 cement plant flue gases: parametric study with different solvents and configurations combined
8 with intercooling. Int. Conf. Greenh. Gas Control Technol., Melbourne: 2018.
9 [51] Knudsen JN, Jensen JN, Vilhelmsen PJ, Biede O. Experience with CO2 capture from coal flue
10 gas in pilot-scale: Testing of different amine solvents. Energy Procedia 2009;1:783–90.
11 doi:10.1016/j.egypro.2009.01.104.
12 [52] IEAGHG. Gaseous Emissions from Amine Based Post-Combustion CO2 Capture Processes and
13 Various Methods for their Deep Removal. 2012.
14 [53] Dubois L, Thomas D. Simulations of various Configurations of the Post-combustion CO2
15 Capture Process Applied to a Cement Plant Flue Gas: Parametric Study with Different Solvents.
16 Energy Procedia 2017;114:1409–23. doi:10.1016/j.egypro.2017.03.1265.
17 [54] Bremer J, Rätze KHG, Sundmacher K. CO2 methanation: Optimal start-up control of a fixed-bed
18 reactor for power-to-gas applications. AIChE J 2017;63:23–31. doi:10.1002/aic.15496.
19 [55] Gao J, Wang Y, Ping Y, Hu D, Xu G, Gu F, et al. A thermodynamic analysis of methanation
20 reactions of carbon oxides for the production of synthetic natural gas. RSC Adv 2012;2:2358.
21 doi:10.1039/c2ra00632d.
22 [56] Schildhauer TJ, Biollaz SMA. Synthetic Natural Gas from Coal, Dry Biomass, and Power-to-
23 Gas Applications. Hoboken, NJ, USA: John Wiley & Sons, Inc.; 2016.
24 doi:10.1002/9781119191339.
25 [57] Xu J, Froment GF. Methane steam reforming, methanation and water-gas shift: I. Intrinsic
26 kinetics. AIChE J 1989;35:88–96. doi:10.1002/aic.690350109.
27 [58] Kao YL, Lee PH, Tseng YT, Chien IL, Ward JD. Design, control and comparison of fixed-bed
28 methanation reactor systems for the production of substitute natural gas. J Taiwan Inst Chem
29 Eng 2014;45:2346–57. doi:10.1016/j.jtice.2014.06.024.
30 [59] Schaaf T, Grünig J, Schuster MR, Rothenfluh T, Orth A. Methanation of CO2 - storage of
31 renewable energy in a gas distribution system. Energy Sustain Soc 2014;4:1–14.
32 doi:10.1186/s13705-014-0029-1.
33 [60] Götz M, Lefebvre J, Mörs F, McDaniel Koch A, Graf F, Bajohr S, et al. Renewable Power-to-
34 Gas: A technological and economic review. Renew Energy 2016;85:1371–90.
35 doi:10.1016/j.renene.2015.07.066.
36 [61] Ducamp J, Bengaouer A, Baurens P, Fechete I, Turek P, Garin F. Statu quo sur la méthanation
37 du dioxyde de carbone : une revue de la littérature. Comptes Rendus Chim 2018;21:427–69.
38 doi:10.1016/j.crci.2017.07.005.
39 [62] Lin H, Thompson SM, Serbanescu-Martin A, Wijmans JG, Amo KD, Lokhandwala KA, et al.
40 Dehydration of natural gas using membranes. Part I: Composite membranes. J Memb Sci
41 2012;413–414:70–81. doi:10.1016/j.memsci.2012.04.009.
42 [63] Sijbesma H, Nymeijer K, van Marwijk R, Heijboer R, Potreck J, Wessling M. Flue gas
43 dehydration using polymer membranes. J Memb Sci 2008;313:263–76.
44 doi:10.1016/j.memsci.2008.01.024.
45 [64] Potreck J, Nijmeijer K, Kosinski T, Wessling M. Mixed water vapor/gas transport through the
46 rubbery polymer PEBAX® 1074. J Memb Sci 2009;338:11–6.
47 doi:10.1016/j.memsci.2009.03.051.
48 [65] Lin H, Thompson SM, Serbanescu-Martin A, Wijmans JG, Amo KD, Lokhandwala KA, et al.
49 Dehydration of natural gas using membranes. Part II: Sweep/countercurrent design and field test.
50 J Memb Sci 2013;432:106–14. doi:10.1016/j.memsci.2012.12.049.
51 [66] Baker RW. Gas Separation. Membr. Technol. Appl. Second Edi, Chichester, UK: John Wiley &
52 Sons, Ltd; 2004, p. 301–53. doi:10.1002/0470020393.ch8.
53 [67] Sanders DF, Smith ZP, Guo R, Robeson LM, McGrath JE, Paul DR, et al. Energy-efficient
54 polymeric gas separation membranes for a sustainable future: A review. Polymer (Guildf)
55 2013;54:4729–61. doi:10.1016/j.polymer.2013.05.075.

24
To cite : DOI: 10.1016/j.apenergy.2019.114249

1 [68] Scholz M, Frank B, Stockmeier F, Falß S, Wessling M. Techno-economic analysis of hybrid


2 processes for biogas upgrading. Ind Eng Chem Res 2013;52:16929–38. doi:10.1021/ie402660s.
3 [69] Fendt S, Buttler A, Gaderer M, Spliethoff H. Comparison of synthetic natural gas production
4 pathways for the storage of renewable energy. Wiley Interdiscip Rev Energy Environ
5 2016;5:327–50. doi:10.1002/wene.189.
6 [70] Molino A, Migliori M, Ding Y, Bikson B, Giordano G, Braccio G. Biogas upgrading via
7 membrane process: Modelling of pilot plant scale and the end uses for the grid injection. Fuel
8 2013;107:585–92. doi:10.1016/j.fuel.2012.10.058.
9 [71] Peters MS, Timmerhaus KD. Plant Design & Economics for Chemical Engineers. McGraw-Hill
10 Publishing Company; 1991.
11 [72] Zhang C, Gao R, Jun K-W, Kim SK, Hwang S-M, Park H-G, et al. Direct conversion of carbon
12 dioxide to liquid fuels and synthetic natural gas using renewable power: Techno-economic
13 analysis. J CO2 Util 2019;34:293–302. doi:10.1016/j.jcou.2019.07.005.
14 [73] de Medeiros JL, de Oliveira Arinelli L, Teixeira AM, Araújo O de QF. Offshore Processing of
15 CO2-Rich Natural Gas with Supersonic Separator. Springer International Publishing; 2019.
16 doi:10.1007/978-3-030-04006-2.
17 [74] Michailos S, McCord S, Sick V, Stokes G, Styring P. Dimethyl ether synthesis via captured CO2
18 hydrogenation within the power to liquids concept: A techno-economic assessment. Energy
19 Convers Manag 2019;184:262–76. doi:10.1016/j.enconman.2019.01.046.
20 [75] Li K, Leigh W, Feron P, Yu H, Tade M. Systematic study of aqueous monoethanolamine (MEA)-
21 based CO2 capture process: Techno-economic assessment of the MEA process and its
22 improvements. Appl Energy 2016;165:648–59. doi:10.1016/j.apenergy.2015.12.109.
23 [76] Glenk G, Reichelstein S. Economics of converting renewable power to hydrogen. Nat Energy
24 2019. doi:10.1038/s41560-019-0326-1.
25 [77] Bejan A, Tsatsaronis G, Moran M. Thermal Design and Optimization. 1st Editio. Wiley-
26 Interscience; 1996.
27 [78] Gutiérrez-Martín F, Rodríguez-Antón LM. Power-to-SNG technology for energy storage at large
28 scales. Int J Hydrogen Energy 2016;41:19290–303. doi:10.1016/j.ijhydene.2016.07.097.
29 [79] Gahleitner G. Hydrogen from renewable electricity: An international review of power-to-gas
30 pilot plants for stationary applications. Int J Hydrogen Energy 2013;38:2039–61.
31 doi:10.1016/j.ijhydene.2012.12.010.
32

25

View publication stats

You might also like