You are on page 1of 13

Journal of CO2 Utilization 60 (2022) 101991

Contents lists available at ScienceDirect

Journal of CO2 Utilization


journal homepage: www.elsevier.com/locate/jcou

Environmental and economic performance of carbon capture with


sodium hydroxide
Enrique Medina-Martos a, b, José-Luis Gálvez-Martos a, *, Jorge Almarza c, Carlos Lirio a,
Diego Iribarren a, Antonio Valente a, Javier Dufour a, b
a
Systems Analysis Unit, IMDEA Energy, Avda. Ramón de la Sagra 3, 28935 Móstoles, Madrid, Spain
b
Chemical and Environmental Engineering Group, Rey Juan Carlos University, Calle Tulipán s/n, 28933 Móstoles, Madrid, Spain
c
Ecological World for Life España (EWL), Calle Castelló, 23, 2º Izda., 28001 Madrid, Spain

A R T I C L E I N F O A B S T R A C T

Keywords: Carbon capture and utilization processes can lead to the manufacture of carbonates with low, neutral or even
Carbon capture and utilization negative carbon footprint. In this sense, we consider the utilization of solid sodium hydroxide, NaOH, as an agent
CO2 mineralisation for CO2 capture in different conceptual approaches to produce usable carbonates. The experimental background
Sodium hydroxide
on this application has proven its technical feasibility, although the high cost and carbon footprint of the elec­
Sodium carbonate
trolytic production of NaOH, mainly due to high electricity consumption, are the main barriers for a widespread
Carbon footprint
Techno-economic analysis adoption. However, the use of NaOH has not yet been compared and benchmarked in the current and future
scenarios of carbon capture, where the share of renewable sources is progressively increasing in many national
electricity mixes and further applications of carbonates are potentially available. To account for these scenarios,
we considered different sources of CO2 and NaOH, and the potential utilization of the generated products.
Altogether, we evaluated 18 scenarios regarding the impact on life cycle global warming potential and fossil
resources depletion, along with the cost of capture. Results show that net emissions savings and potential cost
reductions are attainable for both flue-gas and air-based capture processes, provided that the resulting sodium
carbonate is used to substitute the current, Solvay-based, production of sodium carbonate.

1. Introduction emitted by industrial activities. Furthermore, the continued price esca­


lation of fossil fuels might be a synergic driver for CCU [14,15]. Yet, the
The mitigation of the long-term impacts of global warming requires a contribution of CCU to climate change mitigation is still of scarce sig­
net reduction of CO2 emissions [1,2]. Among other options, Carbon nificance [8], since the demand for CO2-based products is far under its
Capture and Storage (CCS) is recognised as a relevant strategy to achieve production rate [7,8,14]. In spite of this, CCU is still conceptualized as
industrial decarbonisation [3–5]. However, despite some large CCS an initial incentive to skip the economic gap for CCS, as well as a vector
projects are currently operating [3,6], a broader deployment is pre­ for local circular economy solutions [13,15].
vented by the uncertainty related to long-term security and high energy Carbon capture technologies, which are common to both CCS and
penalty and costs [7]. CCU, are classified into three categories: (i) pre-combustion capture, (ii)
In this context, Carbon Capture and Utilization (CCU) is referred to post-combustion capture and (iii) oxy-fuel combustion capture [14].
those processes aimed to utilize captured CO2 as a raw material for Post-combustion technologies, which involve capture from waste gas
products or services with a potential market value [8–12]. Globally, streams, are the most developed and widespread option [6]. This in­
~230 Mt CO2 are used every year, being urea manufacturing and cludes absorption in solvents, adsorption by solid sorbents, membrane
Enhanced Oil Recovery (EOR) the largest commercial applications [13]. separation, cryogenic separation, and pressure or vacuum swing
Novel usage pathways involve the conversion of CO2 into fuels, chem­ adsorption [14]. Within capture on solid sorbents, a subcategory is that
icals and building materials. When compared to fossil carbon sources, of CO2 carbonation (also referred to as mineralisation), which consists in
captured CO2 provides a more secure supply to produce such com­ the reaction of CO2 with a metal oxide (typically Ca or Mg minerals) to
modities, due to its higher availability, as long as it continues to be form insoluble solid carbonates [3,10,14,16]. As the energy state of

* Corresponding author.
E-mail address: joseluis.galvez@imdea.org (J.-L. Gálvez-Martos).

https://doi.org/10.1016/j.jcou.2022.101991
Received 4 August 2021; Received in revised form 11 March 2022; Accepted 18 March 2022
Available online 4 April 2022
2212-9820/© 2023 The Authors. 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/).
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

carbonates (60–180 kJ mol− 1) is lower than that of CO2 (400 kJ mol− 1), Table 1
they are considered a secure, long-term carbon sink [3,4]. Carbonation Overview of analyzed scenarios.
can be performed in-situ, when CO2 is injected into geological forma­ Scenario Carbon Source Source of Final Processing /
tions of minerals, such as serpentine or olivine, or ex-situ, if it occurs # capture of CO2 alkali/solvent product
above ground over already mined rocks or other sorbents [3]. Moreover, technology
carbonation might be direct or indirect, depending on the process taking 1 NaOH - solid Air Commercial Na2CO3
place in one or multiple steps [3,4,14]. NaOH
In this study, we consider the utilization of dry, solid NaOH as a 2 NaOH - solid Air Commercial Regeneration of
NaOH NaOH, CO2
sorbent in a fixed bed carbonation reactor. An experimental background compression and
on this application can be found in several references [16–21]. Its use in storage
CCS processes is often discarded due to the high energy required for the 3 NaOH - solid Air Commercial Construction
thermal regeneration of NaOH from carbonates [18]. This fact, in NaOH material
4 NaOH - solid Air Re-NaOH Na2CO3
addition to the high costs and the relatively high carbon footprint of
5 NaOH - solid Air Re-NaOH Regeneration of
electrolytic NaOH production, mainly due to high electricity consump­ NaOH, CO2
tion, are the primary barriers to employ it in carbon capture applica­ compression and
tions. However, this use has not yet been compared and benchmarked in storage
the current and future scenarios of carbon capture, where the electricity 6 NaOH - solid Air Re-NaOH Construction
material
for NaOH production is becoming increasingly renewable, further uti­ 7 NaOH - solid Flue Gas Commercial Na2CO3
lization options of carbonates are potentially available, and novel NaOH
technologies are being developed [22]. 8 NaOH - solid Flue Gas Commercial Regeneration of
In this context, the main novelty of this work is the thorough analysis NaOH NaOH, CO2
compression and
of such potential and prospective role of sodium hydroxide as a vehicle
storage
for the capture of CO2 within a life cycle perspective. The goal is to assess 9 NaOH - solid Flue Gas Commercial Construction
the environmental and economic feasibility of a carbon capture concept NaOH material
based on the direct carbonation of solid NaOH. A life cycle assessment 10 NaOH - solid Flue Gas Re-NaOH Na2CO3
(LCA) was conducted by comparing the target technology with others of 11 NaOH - solid Flue Gas Re-NaOH Regeneration of
NaOH, CO2
reference, namely amine scrubbing (currently the benchmark option for
compression and
post-combustion capture) and indirect calcium looping, which is an storage
alternative carbonation option often proposed for atmospheric CO2. This 12 NaOH - solid Flue Gas Re-NaOH Construction
study considers different sources of CO2 and NaOH, and the potential material
13 Amine - Air MDEA CO2 compression and
utilization of the generated products (carbonates or pure CO2). Alto­
MDEA storage
gether, 18 scenarios were evaluated regarding the life cycle greenhouse 14 Amine - Flue Gas MDEA CO2 compression and
gases (GHG) emissions and fossil resources depletion, along with the MDEA storage
evaluation of the cost of capture. 15 Ca-loop Air Commercial CO2 compression and
(NaOH) NaOH storage
16 Ca-loop Air Re-NaOH CO2 compression and
2. Methodology (NaOH) storage
17 Ca-loop Flue Gas Commercial CO2 compression and
2.1. Overview (NaOH) NaOH storage
18 Ca-loop Flue Gas Re-NaOH CO2 compression and
(NaOH) storage
This work follows a comparative approach to evaluate and bench­
mark the environmental and economic performance of several CCU
scenarios and CCS reference alternatives. Spanish conditions were abovementioned factors are described.
considered as far as possible (regarding e.g. electricity mix) due to the
presence of technology actors interested in the development and
commercialization of the core solutions assessed in this work. The 2.2. System description
formulation of scenarios for carbon capture depends on the following
factors: 2.2.1. Source of CO2 and reference flows
Source of CO2. Two sources of CO2 were considered: atmospheric air When air was the inlet gas, it was assumed to be composed by 0.04%
and flue gas from fossil fuel combustion. CO2, 79.01% N2 and 20.95% O2, on a volume basis, dry-free, since the
Source of NaOH (or the capture agent in the reference cases). For the water content does not affect calculations. The environmental condi­
analyzed system, two sources of NaOH were considered: commercial tions were 25.0 ºC and 1.00 bar in the calculations. The reference flow in
NaOH, produced from brine electrolysis assuming the 2017 Spanish this case was set to 1 t of CO2 per day at the capture point (i.e. the
electricity mix (31% renewable), and renewable sodium hydroxide (Re- amount of CO2 retained onsite by the capture equipment). This is the
NaOH), which assumes 100% electricity from wind turbines. For the amount of CO2 foreseen for the application of capture to the exhaust air
reference cases, methyl diethanolamine (MDEA) was considered in the stream from large commercial buildings, or from large scale street urban
amine-based processes. air cleaners, at an airflow of 50 – 60,000 m3/h. Larger scales are possible
Utilization of products. In solid NaOH systems, sodium carbonate was for the application of CO2 capture from air, but those would fall out of
assumed to be directly commercialized due to its high purity, employed the initial scope of the application of small size NaOH filters in urban air
for the fabrication of construction materials, or used to regenerate cleaners [23]. In the base case, this option is penalized with the energy
NaOH, while producing compressed CO2 for storage. consumption required to pump air and overcome the pressure drop
Capture technology. Three technologies were chosen: reaction with along the bed. Solid beds pressure drop were assumed at 5KPa, which
solid NaOH in a fixed bed reactor, solvent absorption with amines, and can be reduced by further design, but it is a first approximation to a
reaction with dissolved NaOH in a closed cycle using indirect calcium complex process where the bed material would be changing shape, size
looping. and chemistry. For liquid-solid absorbers 2 kPa were assumed as pres­
A summary of all analyzed scenarios is presented in Table 1. In the sure drop, as per measured ranges [24].
following sections, the assumptions and specificities of the In the case of flue gas, a combustion process (800 ºC, 16.0 bar, 15%

2
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

air excess) of pure CH4 was assumed to produce the inlet stream to the
CO2 + 2H2 O⇋HCO3 − + H3 O+ (4)
capture process. The composition (w/w) of the flue gas after expansion
(40.0 ºC, 2.50 bar) and NOx removal was 14.90% CO2, 80.50% N2,
HCO3 − + H2 O⇋CO3 2− + H3 O+ (5)
2.80% O2, and 1.80% H2O. It must be noted that, in this study, these
combustion and conditioning stages are out of the system boundaries. The capture of CO2 from air through indirect calcium looping and
Here, the reference flow was set to 1 Mt of CO2 per year at the capture causticisation is based on the work by Keith et al. [29]. The process
point. For the base case, the flue gas stream was considered to overcome consists of two connected chemical loops. In the first stage (air contactor
the pressure drop in the absorption beds for all scenarios and no energy or absorber), CO2 is absorbed from the inlet gas by using an aqueous
consumption for compression or pumping is required at the inlet. solution of alkali to form carbonate. Then, in the second stage (pellet
reactor), the carbonate reacts with Ca(OH)2 to form again the alkali and
2.2.2. Capture technology solid CaCO3, which precipitates. While the alkali is recycled to the first
The targeted capture technology in this work is based on a gas-solid stage, closing the first loop, the CaCO3 is calcined (thermally decom­
chemical reaction in which solid NaOH reacts with CO2 in a gas stream. posed) at 900 ºC using natural gas as fuel with oxygen in a third stage,
Previous experiments have shown a high yield of CO2 uptake [22], producing CaO and CO2, which is compressed to 150 bar as per frame­
observing undetectable concentrations of CO2 after the reaction of pure work assumptions. Finally, in stage 4 (slaker), CaO is hydrated to pro­
CO2 in a bed of sodium hydroxide. The obtained solid is mostly duce Ca(OH)2, which is recycled to stage 2, closing the second loop.
composed by sodium carbonate, with a net consumption of 2 moles of Although the original publication assumes KOH for the cycle, NaOH was
NaOH per mole of CO2, although the reaction can continue till the for­ analyzed in this work (Fig. S3 in ESI). Since the process utilizes an
mation of sodium bicarbonate (1 mol NaOH / mol CO2) under wet aqueous solution of NaOH and further includes its regeneration through
conditions, as shown in the reaction scheme below: a thermal route using CaCO3 produced in a causticisation reactor, it can
be compared to other options for NaOH regeneration. The reaction
2NaOH + CO2 ⇋Na2 CO3 + H2 O (1 A)
scheme is shown below:
Na2 CO3 + CO2 + H2 O⇋2NaHCO3 (1B) Absorption:

From a thermodynamic standpoint, in a theoretically adiabatic CO2 (g) + 2NaOH(aq)⇋H2 O(l) + Na2 CO3 (aq) (6)
process, a 100% conversion reaction of NaOH is plausible and was NaOH regeneration in the pellet reactor (causticisation):
assumed. This means the capture process would only rely on the utilized
amount of NaOH, being independent of the influx CO2 concentration. Na2 CO3 (aq) + Ca(OH)2 (s)⇋CaCO3 (s) + 2NaOH(aq) (7)
Considering this, the functional unit (i.e. the unit describing the function Calcination:
provided by the system) was selected as 1 kg of captured CO2 (i.e. 1 kg
CO2 at the point of capture). It was assumed that all NaOH converts to CaCO3 (s)⇋CaO(s) + CO2 (g) (8)
sodium carbonate, as the heat released by reaction 1 A, which is highly Hydration:
exothermic, would reverse reaction 1B and produce bicarbonate
decomposition. Experimentally, the reaction of CO2 with alkali can CaO(s) + H2 O(l)⇋Ca(OH)2 (s) (9)
reach yields higher than 90% in a few minutes in the appropriate re­ In this work, we assess four scenarios for this technology, scenarios
action set-up for flue gases [16], and over 40% in the case of atmo­ 15–18 in Table 1, in which fresh NaOH (either conventional or renew­
spheric concentrations of CO2; these yields were assumed in the able energy-based) is applied to both atmospheric CO2 capture and flue
calculation of outlet gas composition. gas processing.
Two reference systems were studied: absorption of CO2 with amines
and indirect calcium looping using NaOH. Chemical absorption using 2.2.3. Sodium hydroxide production
aqueous amine solutions is currently the only mature technology The chlor-alkali industry is the main producer of sodium hydroxide
regarding CO2 capture, which has been widely deployed on industrial in the world [30,31]. In this process, NaOH is produced along with Cl2
applications from the late 1970s [10,25]. In a typical amine system, the and H2 by electrolysis of NaCl. Two technologies account for most of the
gas stream containing CO2 is fed at the bottom of an absorption column installed capacity in Europe: membrane cell (85%) and diaphragm cell
(scrubber), while the aqueous amine solution is injected at the top. CO2 (10%). The remaining 5% corresponds to the mercury cell (currently
is transferred to the amine solution when both phases come into contact, phased-out) and alternative production routes [32]. For the sake of LCA
obtaining a CO2-depleted gas stream. The CO2-rich solution is then sent and cost calculations, in this study we assumed a weighted input-output
to a stripper, where it is heated to release the CO2, whereas clean amine inventory that only considers the membrane and diaphragm technolo­
is recycled to the absorber (Fig. S2 in ESI). gies. Two options were regarded, depending on the electricity source
Traditionally, monoethanolamine (MEA) has been the benchmark used for the electrolytic process:
solvent for this application. However, other compounds such as piper­
azine (PZ), diethanolamine (DEA) and methyldiethanolamine (MDEA) - Commercial NaOH: utilizes the average Spanish electricity mix for
have increasingly been studied and used over last years. Here, we the year 2017, composed of ca. 30% input of renewable energy
considered the use of MDEA following the approach of other authors sources, according to ecoinvent 3.7.1. cutoff modelling approach
[25]. Furthermore, thermal regeneration, which is of high energy de­ [33].
mand and accounts for 80% of the operating cost for MEA-based cap­ - Renewable energy-based NaOH (Re-NaOH): based on the same
ture, can be reduced if MDEA is chosen instead [26,27]. The energy process inventory, but assumes 100% wind turbine electricity.
demand of MEA and MDEA solutions regeneration was reported 4.4 GJ
and 2 GJ per tonne of captured CO2, respectively [27]. In those scenarios where NaOH was not considered, scenarios 13 and
The following electrolyte reaction system describes the absorption of 14, commercial sources were assumed for the supply of the considered
CO2 by MDEA [28]: capture agent, such as fresh MDEA, which impact was assumed equal to
MDEA − H + + H2 O⇋MDEA + H3 O+ (2) that of diethanolamine for amine-based processes. Fresh lime was
considered the main input for indirect Ca-looping scenarios.
2H2 O⇋OH − + H3 O+ (3)

3
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

2.2.4. Final processing and products utilization (e.g. desalination brines) to form MgCO33 H2O (nesquehonite), which
As a general approach for NaOH-based cases, the substitution of the after a dehydration/rehydration sequence forms a coherent solid with
commercial production of Na2CO3 through a modified Solvay process similar characteristics to gypsum plaster, which was assumed to be
[34] was considered, using the dataset from ecoinvent 3.7.1 to model substituted by the final magnesium carbonate. Calcium carbonate is
the avoided impacts. The main inputs and outputs of such processes are obtained as a by-product, which can substitute construction aggregates.
shown in Fig. 1a. The hardening process releases 1/5 moles of the CO2 captured, as a
Sodium carbonate could also be used for the regeneration of NaOH consequence of a chemical change in the carbonates. The mass and
and the production of a pure CO2 stream; in this regard, the thermal energy balance of this option was modelled based on previous experi­
route was discarded, as it was included in the reference processes (sce­ ence by the authors [11], as shown in Fig. 1c.
narios 15–18), so an electrolytic Cl2-free NaOH production process from In all scenarios producing CO2 gaseous streams, e.g. as those with
sodium carbonate was modelled as proposed in the state-of-the-art [35]. NaOH regeneration, amine-based or indirect Ca looping (reference
In this process, originally designed to decouple the production of NaOH technologies), a compression stage (150 bar) was included within the
and Cl2, pure CO2 is produced and NaOH is obtained at a theoretical analyzed system; no further impact from transportation or storage of
maximum concentration of 16% (w/w) in water, achievable due to the CO2 was considered within the system boundaries. All the solid products
solubility limitations of Na2CO3. The applied voltage is that of water are assumed to be transported by lorry 100 km after their production.
electrolysis, slightly less than 2 V per cell. The achieved current effi­
ciency is 85%. With these data, we estimated the mass and energy
2.3. Life cycle assessment
input-output balance, shown in Fig. 1b.
In addition to the generic substitution of commercial sodium car­
2.3.1. Goal and scope
bonate, we also considered the final utilization of sodium carbonate to
The goal of this LCA study was to evaluate and compare the life cycle
produce a plaster-like construction material using Mg-rich brines in a
environmental performance of a number of systems aimed to capture
CCU approach [11,36,37]. Sodium carbonate reacts with Mg-rich brines
CO2 from atmospheric air or from a flue gas stream (coming from a

Fig. 1. Material and energy inventories of considered Na2CO3 utilization alternatives: a) Substitute for the production of commercial Na2CO3 via a modified Solvay
Process; b) Raw material for electrolytic chlorine-free NaOH regeneration; c) Raw material for the production of construction materials.

4
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

nearby production plant). The functional unit of the systems was set as ( )
EUR2020
1 kg of CO2 at the point of capture (after its absorption/reaction) from Capture ​ process ​ cost ​
either the air or the target flue gas stream. When other functions are kg CO2 captured
performed, multifunctionality was addressed through the consideration =
Annual Operating Costs (EUR2020 )
(10)
of avoided burdens, which provide environmental credits. The flue gas Annual mass of CO2 absorbed or adsorbed (kg)
emitter was not credited with environmental benefits or impacts from The Annual Operating Costs (numerator in Eq. 10) are composed of
the processes. Furthermore, it should be noted that the emissions variable (flow-dependent) and fixed (flow-independent) costs. For the
captured from flue gas are avoided and those from the air are removed variable costs we utilized unitary cost data published either by public
from the atmosphere, and both consequently have a different character. organisations or available in the literature. The revenues from secondary
However, the two captured streams were modelled as negative emis­ products or by-products, which can be directly marketed, were sub­
sions in order to benchmark the life cycle performance of each capture tracted (i.e. accounted as negative costs). Fixed costs were mostly
approach. In the case of flue gas, the industrial plant originating this calculated as a certain percentage of the investment cost. To estimate the
stream was not included within the system’s boundaries. A graphical investment cost, we used the capacity ratioing approach (William’s
representation of the LCA scope is provided in Fig. 2. method), based on reference plant costs. No annualization of capital
costs was included in the calculation. All the main assumptions used in
2.3.2. Inventory data the cost calculations are reported in Table 2.
Data acquisition for the inventory data of the foreground system, i.e. Cost data were all updated to December 2020, by using the Producer
the capture process, was based on the mass and energy balances of the Plant Cost Index (PPI) for raw materials and the Chemical Engineering
simulation models presented in the ESI. The ecoinvent database 3.7.1 Plant Cost Index (CEPCI) for investment cots. For consistency, as most
was used to model the inputs and outputs of background processes in the cost references were found on a USD basis, we utilized the American PPI,
supply chain of raw materials and energy. The full list of ecoinvent ac­ reported by the United States Bureau of Labor Statistics. Then, all costs
tivities used to model the process are reported in the ESI. were converted to EUR. All assumed unitary data costs, reference plant
investment costs and a deeper insight on the used indicators are pro­
2.3.3. Life cycle impact assessment vided in the ESI.
Two life cycle indicators were assessed: carbon footprint (i.e., global Special attention was given to the unitary cost assumed for NaOH, as
warming impact potential with 100-year horizon; IPCC method 2013), it is the key feature studied in this work. The reported cost of NaOH from
and the fossil resources depletion (i.e. abiotic resource depletion – fossil the UN Comtrade database for the Spanish imports from France was EUR
fuels, ADP-fossil, from the CML2002 method). Both are recommended 480 per tonne in 2019. To account for the influence of the considered
by the Product Environmental Footprint methodology [38]. electricity source, first, we updated this cost to January 2020. Then a
cost change factor was added, which includes the cost change for the
2.4. Techno-economic assessment consumption of wind electricity.

We evaluated the economic performance of the addressed scenarios


by means of the cost of the CO2 capture process, referred to 1 kg of CO2 2.5. Sensitivity Analysis
retained at the point of capture. It can be defined as:
In order to complete the analysis of the CO2 capture process, a
sensitivity analysis of the carbon footprint and costs of the approach
were implemented for the following factors:

Fig. 2. General system definition for the studied scenarios.

5
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Table 2 - alkali, sorbents and other raw materials, which include the life-cycle
Main techno-economic considerations assumed in the study. impact derived from the supply chain of chemicals;
Feature Value/Calculation/Comment - heating and electricity related impacts for each part of the process in
each scenario where it is relevant: capture (including regeneration of
Base Year 2020(EUR/USD) = 0.83
Currency EUR amines and of CaO in the reference units), electrolytic regeneration
Geographical Scope Europe of NaOH, or the manufacture of construction materials from sodium
Annual Operating Costs Variable Costs + Fixed Costs carbonate;
Variable Costs RM+U+C&A - avoided emissions of CO2 for (i) captured CO2, which is the same for
Raw Materials (RM) See ESI
Utilities (U) See ESI
all scenarios, (ii) avoided through the substitution of commercial
Consumables (C&A) 5% of M sodium carbonate, or (iii) avoided through the substitution of con­
Fixed costs OL&S+TAX+M ventional construction materials;
Labour, Supervision, Laboratory Based on European Salary, 50% of indirect - and those direct emissions from the processes, due to the decompo­
and Overheads (OL&S) costs, high level automation
sition of carbonates in the case of construction materials, and those
Taxes and Insurance (TAX) 1% of TIC
Maintenance (M) 3% of TIC related to transport needs.
Total Investment Cost (TIC) ISBL+OSBL+WC
Fixed Capital Investment ISBL+OSBL First, it is noticeable that under the considered assumptions, the
ISBL C2 = C1 (S2/S1)n a performance of CO2 capture from air is significantly worse than that
OSBL Not considered in this workb
Working Capital (WC) Not considered in this workb
from flue gas.The carbon footprint of all air-based systems is positive.
Thus, there is a net increase of life cycle greenhouse gas emissions to the
a
William’s method. The capital cost (C2) of a plant with a certain capacity (S2) atmosphere, driven mainly by the required electricity to overcome the
can be inferred from that (C1) of a similar plant with a different capacity (S1).
pressure drop. This observation is valid independently of what product
The exponent n accounts for the effect of scale (reduction of capital cost for
is obtained and assumed to offset the environmental impact (sodium
bigger plant capacities). In this work we used a value of 0.6, which is the average
carbonates or construction material), if sodium hydroxide is regenerated
for most chemical engineering plants. Reference plant costs are reported in the
ESI. (thermally or through electrolytic processes), or the capture agent is
b
OSBL and WC, as compared to ISBL, do not have an effect on the calculation changed (sodium hydroxide or amines). For air-based processes, the
of operating costs, which was the main focus of this work. carbon footprint of solid sodium hydroxide-based processes (scenarios
1–6), with higher energy consumption, is remarkably higher than that of
- The stoichiometry of the reaction. The molar ratio for the carbon­ amines capture (scenario 13), and 4–5 times that of indirect calcium
ation reaction assumed in the work is NaOH/CO2 = 2, producing looping (scenarios 15–16), which achieves the lowest footprint, at 0.5 kg
sodium carbonate as final product of capture. However, a different CO2 eq per kg of captured atmospheric CO2. These results were revisited
stoichiometry can occur if the reactor’s efficiency allows for a for 100% renewable electricity input and are shown along with its costs
recarbonation of sodium carbonate towards sodium bicarbonate. in a later stage.
Although the specific input of NaOH per mole of reacted CO2 would In the case of flue gas, several aspects can be highlighted from the
be halved, the avoidance of emissions would be conditioned by the results of Fig. 3:
industrial process used as the reference for the production of car­
bonates. This last option was considered in the sensitivity analysis. In ▪ The use of electrolytic regeneration of NaOH always involves a
this case, the substitution of commercial sodium bicarbonate pro­ net increase of the global warming impact in the assessed sce­
duction was assumed for the calculation of the environmental narios. In such cases, the consumption of electricity and heat in
impact, in parallel to the assumption made for sodium carbonate. the electrolytic regeneration of sodium hydroxide is a main
- Heat recovery. In the case of flue gas, the production of sodium contributor to positive emissions.
carbonate is an exothermic process, which can heat adiabatically up ▪ The avoidance of emissions from sodium carbonate production
to 350 ℃ the flue gas when it contains 10% (w/w) of CO2. A 50% is highly beneficial and leads to the lowest carbon footprint of
recovery of the heat produced was assumed, corresponding to the all the gas-solid adsorption approaches with sodium hydroxide.
production of low pressure steam to be used elsewhere, yielding 1.33 ▪ The use of renewable energy in the production of sodium hy­
MJ of heat coproduced per kg of CO2. This heat was considered to droxide is a major factor for a carbon-neutral or carbon-
avoid heat produced from natural gas in a low NOx burner. negative process, as observed for scenarios 10 and 12.
- Alkali and energy inputs. The footprint and costs of the alkali and ▪ A carbon-negative or carbon-neutral production system of a
energy (electricity and heat) inputs to the process were varied construction material is achievable in the case of renewable
± 20% in order to observe the process economic and environmental energy use in the production of sodium hydroxide. The off­
performance sensitivity against these factors. setting of gypsum plasterboard production generates a signifi­
cant environmental credit, although less important than the
3. Results and discussion avoidance of sodium carbonate production.
▪ Flue gas-based scenarios on the capture of CO2 with NaOH to
3.1. Inventory data substitute Solvay sodium carbonate (scenarios 7 and 10) have a
negative balance of GHG emissions and outperform amine-
Table 3 shows the inventory data for the foreground processes of all based processes. In these cases, the potential carbon footprint
the scenarios defined in Table 1. is lower than − 1 kg CO2 eq, i.e. these systems are able to avoid
emissions beyond the amount of CO2 contained in the flue gas.
3.2. LCA results
3.2.2. Energy footprint
3.2.1. Carbon footprint The results observed for the net balance of fossil fuel resources of the
The balance of the greenhouse gases for all the evaluated scenarios assessed scenarios are parallel to those for the GHG emissions, as
are shown in Fig. 3. The assessed categories of contribution are: observed in Fig. 4. However, the net values, in this case, are always
positive except for scenario 10, when CO2 is captured using solid sodium
hydroxide produced with renewable energy.
The other options involve an increase in the consumption of fossil

6
E. Medina-Martos et al.
Table 3
Main inventory components and data per analyzed carbon capture scenario.
Scenario Unit 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

CO2 capture
Input
Commercial NaOH kg 1.82 1.82 1.82 1.82 1.82 1.82 0.15 0.109
Re-based NaOH kg 1.82 1.82 1.82 1.82 1.82 1.82 0.15 0.1087
Electricity, grid MJ 9.00 9.00 9.00 9.00 9.00 9.00 2.7 0.53 2.53 2.53 0.57 0.57
Heat, from NG MJ 71 5.1 0.04 0.04 0.0017 0.0017
NG kg 0.149 0.149 0.143 0.143
O2 (from ASU) kg 0.596 0.596 0.572 0.572
Water L 0.032 0.033 9.15 9.15 1.35 1.35
MDEA, consumed kg 0.0011 4E-6
Hydraulic lime kg 0.054 0.054 0.034 0.034
Output
MDEA emitted to atm. kg 0.0011 4E-6
Wastewater to treatment L 0.032 0.033 9.15 9.15 1.35 1.35
Avoided CO2 to atm. kg -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1
Avoided Na2CO3 kg -2.41 -2.41 -2.41 -2.41
Distribution and logistics
7

Transport tkm 0.24 0.267 0.24 0.267 0.24 0.267 0.24 0.267
NaOH electrolytic regeneration
Input
Water L 19.4 19.4 19.4 19.4
Electricity, grid MJ 5.1 5.1 5.1 5.1
Heat, NG MJ 20.1 20.1 20.1 20.1
Output
Comm. NaOH, regenerated kg -1.54 -1.54
Re-based NaOH, regenerated kg -1.54 -1.54
Wastewater to treatment. L 18.9 18.9 18.9 18.9
Plaster-like material production
Input
Water L 26.4 26.4 26.4 26.4
Electricity MJ 0.66 0.66 0.66 0.66
Heat, NG MJ 2 2 2 2
Output

Journal of CO2 Utilization 60 (2022) 101991


CO2 emissions to atm. kg 0.138 0.138 0.138 0.138
Avoided plaster-like product kg -2.16 -2.16 -2.16 -2.16
Avoided limestone-type aggregate kg -0.51 -0.51 -0.51 -0.51
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Fig. 3. Life cycle greenhouse gas emissions for the assessed scenarios.

Fig. 4. Fossil fuel resource depletion for the assessed scenarios.

fuel resources: air-based processes (scenarios 1–6) produce a large of CO2.


depletion of fossil resources, while for flue gas, amine-based capture of
CO2 produces the lowest impact (scenario 14). Indirect calcium looping 3.3. Economic results
(scenarios 17–18) increases the demand of fossil resources with respect
to amines (scenario 14); as a matter of fact, all of the alkali regeneration The cost of CO2 capture for each scenario is represented in Fig. 5,
approaches assessed in this studio (electrolytic or indirect Ca-looping) where the main contributions are also shown. As stated in the meth­
have highly significant energy footprints, independently of the source odology section, revenues correspond to negative costs as a consequence

8
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Fig. 5. Costs of capture for the assessed scenarios.

of the commercialization of the produced carbonates (sodium carbonate materials; ii) the regeneration of sodium hydroxide at low cost, which is
for scenarios 1, 4, 8 and 10, and a plasterboard substitute for scenarios 3, not achievable within the studied possibilities; and iii) the minimization
6, 9 and 12). of energy and raw materials costs.
All the options involving CO2 capture from air involve prohibitive Another aspect derived from the analysis of costs is the parallelism
costs. Even in the case of revenues from carbonates, the cost of capture between the carbon footprint and the cost of capture. Indeed, both can
still remains above EUR 2 per kg of CO2. Most of this cost is caused by be interpreted as indicators of process performance and resource effi­
fixed costs derived from the maintenance and labour for the in­ ciency in the supply chain and produce a very similar prioritization of
stallations. This is a main effect from the chosen size of the installation; scenarios. The chart in Fig. 6a represents the cost of capture against the
it is possible to further reduce the cost if a higher scale, e.g. 10000 carbon footprint and shows a clear division between air- and flue gas-
tonnes per year were targeted; costs would be in the range EUR 0.4 – 0.8 based processes; air-based approaches, under our assumptions, gener­
per kg of captured CO2, in line with those estimated for direct air capture ally result in higher carbon footprints and costs.
[39]. It is also noticeable that the amount, and therefore costs, of heat However, all these results can easily be challenged by the assump­
required in indirect Ca-looping is independent from the concentration of tions made in the study, so a further evolution of the results can be
CO2, since the generated calcium carbonate to be calcined is the same evaluated by increasing the scale of the installation for the atmospheric
per kg of captured CO2; this makes this option an interesting alternative capture of CO2 (Fig. 6b), from the initial 365 tonnes per year to 10,000
in comparison to amines for air capture, which require a much higher tonnes per year, and an scenario assuming a fully renewable electricity
amount of heat in the regeneration step for air-based products. scenario (Fig. 6c). The scale of the axis was kept for comparison pur­
For flue gas, costs can be compensated if sodium carbonate is sold at poses. For further info, the full set of data for Fig. 6 is provided in the ESI.
market prices, although the variability of the market could introduce As observed, the range of costs would align between flue gas and air-
significant uncertainty in this balance. In any case, at the assumed pri­ based processes when a higher scale of production is assumed, therefore,
ces, the capture of CO2 with commercial sodium hydroxide would yield minimizing the contribution of fixed costs to the operating costs. The
around EUR 11 per tonne of CO2, which could be increased up to 88 if range of costs for air-based process would be around EUR 0.2 – 0.8 per
renewable hydroxide is used instead. These cases are represented by the kg of captured CO2 from air. It is also noticeable that when 100%
total negative values of scenarios 7 and 10 in Fig. 5. On the other hand, renewable electricity is assumed, most of the scenarios would fall into
the rest of the options with flue gas show costs of capture over EUR 0.3 negative footprint, i.e. they would effectively remove CO2 from the
per kg of CO2, a cost still too high if compared with the cost of amine- emissions balance; three scenarios would remain positive: the two sce­
based capture processes, at EUR 0.055 per kg of CO2. narios using commercial NaOH and air to produce carbonate-based
Fig. 5 shows the large contribution of alkalis to the cost of capture, at construction materials (scenarios 3 and 9), and the conventional
around EUR 0.8 – 0.9 per kg of CO2 captured. This situation is derived amines process applied to capture of CO2 from air (scenario 13).
from the high current market price (estimated at EUR 0.48 per kg in
2020) and the stoichiometry of the reaction towards sodium carbonate, 3.4. Sensitivity
which results in a consumption of 1.82 kg of NaOH per kg of captured
CO2. Therefore, the viability of a capture approach based on sodium A sensitivity analysis was performed over scenarios 10, 11 and 12,
hydroxide is highly dependent on several aspects: i) an efficient which are based on renewable sodium hydroxide to capture CO2 from
compensation mechanism of costs through the commercialization of by- flue gas. The results are summarized in Fig. 7.
products, such as sodium carbonate and carbonate-based construction As observed, the production of sodium bicarbonate instead of

9
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Fig. 6. Costs of capture vs GHG emissions (number of scenarios in brackets as datapoint): (a) original assumptions of the study, (b) using a 10000 tonnes per year as
calculation basis for atmospheric CO2 capture, (c) as per option b plus 100% of electricity from renewable sources (wind).

10
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Fig. 7. Sensitivity analysis of GHG gases and costs for scenarios 10, 11 and 12.

carbonate has almost no effect on the carbon footprint. This was only captured CO2 (range is 0.26 kg CO2eq), while operating costs vary from
assessed for scenario 10, since bicarbonate is not applicable to electro­ EUR 0.14–0.46 per kg of captured CO2 (range is EUR 0.32).
lytic regeneration or the production of construction materials. The effect
of the higher capture capacity towards sodium bicarbonate is compen­ 4. Further discussion
sated, in terms of GHG savings, by the lower emissions avoidance ca­
pacity of the final solid. However, the achievable costs savings produced This study focuses on the question whether NaOH is able to capture
in the capture of CO2 would increase by EUR 44 per tonne of captured CO2 in an environmentally sound and economically feasible manner.
CO2. Results from scenarios 7 and 10 indicate an affirmative answer to such
Although energy recovery can slightly improve the environmental question, provided that the resulting sodium carbonate is used to sub­
performance and reduce costs, its influence is small in comparison to stitute the actual, Solvay-based, production of sodium carbonate. We
other factors. Alkali costs and footprint, however, are highly significant showed in this study net GHG emissions savings to the atmosphere and
in scenario 10, which at the same time is not sensitive to energy costs or potential cost reductions (negative capture costs) are attainable.
carbon footprint. On the other hand, the electrolytic regeneration of the Using the current performance of industrial processes, net reduction
alkali is highly sensitive to the energy source footprint and costs, in emissions can sum up to − 2 kg of CO2 eq. per kg of captured CO2, if
contrast to a minor effect from the alkali supply chain. sodium hydroxide is produced with renewable electricity, or − 1 kg CO2
The case of the production of a construction material based on car­ if commercial sodium hydroxide is used. If sodium carbonate is not
bonates shows a particular behaviour. Its sensitivity is significant to commercialized or it does not substitute commercial carbonates, these
alkali footprint and shows some minor dependence to the energy sources values are − 0.32 kg of CO2 eq (for renewable energy-based NaOH) and
footprint. Such influence is more relevant when costs are assessed. In + 0.54 kg CO2 eq (for commercial NaOH) per kg of captured CO2. These
fact, the GHG balance varies from − 0.66 to − 0.40 kg CO2 eq per kg of figures can be compared with benchmarks from studies using similar

11
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

system boundaries. For example, Müller et al. [40], calculated an in­ as modelled by Keith et al. [29]. Our estimation assumes a NaOH solu­
terval of avoided emissions of − 0.59 to − 0.95 kg of CO2 eq. per kg of tion as main absorber, a higher pressure drop in the absorption stage,
captured CO2 for all the identified sources [40]. In the case of miner­ and losses of alkali are calculated from the required washing steps, while
alization to carbonates, the use of alkaline wastes avoids also a signifi­ Keith et al. assumed no losses or consumption of KOH in their approach.
cant amount of CO2 emissions, ranging from clinker kiln dust at In any case, a similar set of assumptions for our model of indirect cal­
− 0.86 kg of CO2 eq to fly ashes at − 0.55 kg CO2 eq per kg of captured cium looping would yield a carbon footprint of − 0.497 kg of CO2 eq per
CO2 [41]. kg of CO2 captured from air.
Therefore, the merit of climate change mitigation of sodium hy­ The direct capture of CO2 into solid NaOH is a technically simple
droxide is based on (i) the possibility of substituting industrially pro­ process with relatively low investment in comparison to other complex
duced sodium carbonate, (ii) the performance of sodium hydroxide, able approaches. However, its deployment is rather limited by the market of
to capture a large amount of CO2, and (iii) the use of renewable energy by-products, the variability derived from the chlor-alkali industry, and
in the production of sodium hydroxide. This last option constitutes the the price volatility in a complex relationship with the market of chem­
most obvious low-hanging fruit: the current and future implementation icals. In fact, the recent trends in Europe regarding PVC manufacturing,
of renewable energy sources in the electricity market, along with the one of the key drivers of chlorine demand, and the COVID-19 pandemic
increasing need to reduce the emissions of CO2 in the industry facilitates are making the market and price of NaOH highly variable; in 2020 a
the production of sodium hydroxide of lower carbon footprint. The share maximum at around EUR 415 per tonne in the Iberian market was
of renewable sources in the assumed Spanish electricity mix is 31.4% reached, which further decreased to EUR 249 by the end of 2020 [45].
(2017 mix) and it is planned to evolve towards 74% by 2030 and to be Hence, the eco-efficiency of NaOH-based approaches for CCU is heavily
“climate neutral” by 2050 [42]. In line with this, the prospective anal­ affected by market factors.
ysis by Müller et al., showed that, in a low carbon economy, most of the It is important to remark that CO2 capture technologies can involve
carbon capture processes should be driven by renewable energy [40]. co-benefits linked to the reduction of other emissions such as SO2, NOx
In any case, the adoption of sodium hydroxide as a carbon capture and PM contained in flue gases or air, which concentration in the outlet
reactant can be further deployed if a method for its regeneration from stream can drop down to by 95%, 2.5% and 50%, respectively [46]. This
carbonates is feasible. In this study, two options were studied: electro­ would lead to the improvement of the environmental performance in
lytic regeneration and indirect calcium looping. None of these regen­ other environmental impact categories, which would probably help in
eration options was able to actually reduce the GHG emissions to the the further development of urban air cleaners. Since the present envi­
atmosphere, as the regeneration approach consumes a significant ronmental assessment is limited to the climate change impact, other
amount of energy. But, the carbon intensity derived from energy con­ environmental pressures, such as those associated with resource con­
sumption could be reduced by (i) increasing the energy efficiency of the sumption and emissions to soil and water as well as non-GHGs emissions
process, e.g. by using another carbonate cycling to reduce the calcina­ to air are not reflected in the results of this analysis. In this regard,
tion temperature such as titanates [43], and by (ii) reducing the fossil further studies could focus on extending the scope of the present study to
energy footprint of the process using cleaner energy sources as presented other environmental impact categories and social implications which
in Fig. 6c. are highly demanded aspects in decision- and policy-making processes.
For all the carbon capture and utilization approaches analyzed in this
work, the net GHG savings are attributed to the captured CO2 and no 5. Conclusions
emissions are accounted (avoided or added) to the emitter, e.g. if the
flue gas comes from power generation, the carbon footprint of electricity Is sodium hydroxide able to capture CO2 in an environmentally
is not assumed to be reduced. As per our assumptions, the emitter is not sound and economically feasible manner? The experimental evidence
affected by the capture process, thus allowing a fair prioritization of from recent literature has proven the technical feasibility of its use in
carbon capture options. There is, however, a possibility of unfair carbon gas-solid absorption reactors, although the high costs and the relatively
footprint assessments, since, in many cases, the emitter is also respon­ high carbon footprint of NaOH are the main barriers for carbon capture
sible (totally or partially, at least) for the investment on the CCU/S applications. In this study, several scenarios were formulated to calcu­
application [44]. For example, if 50% of emissions savings are allocated late and compare their carbon and fossil energy footprints, and the cost
to the combustion of natural gas in scenario 10, the footprint of the of capture. Net greenhouse gas emissions savings at low or even negative
capture approach would yield approximately − 1 kg of CO2 eq per kg of costs were proven for flue gas and air under certain assumptions. The
captured CO2, while power generation with natural gas would be allo­ merit of the climate change mitigation of sodium hydroxide was found
cated − 1 kg CO2 eq, so the balance would yield a carbon-neutral elec­ to be based on the possibility of substituting industrially produced car­
tricity production. bonates or construction materials and on the use of renewable energy in
Carbon neutrality or negativity from atmospheric CO2 was not the electrolytic production of sodium hydroxide. None of the considered
observed in the base cases used in this work, but cannot be discarded, NaOH regeneration cases was able to prove an actual reduction of
since the use of 100% renewable electricity yields negative footprints for emissions, and only approaches regarding commercial substitution of
most of the air-based scenarios. The high energy consumption assumed final products seem economically feasible. In addition, carbon neutrality
in several parts of the process, such as the initial air pumping, is or negativity from atmospheric CO2 seems possible in the case of certain
responsible for the larger impacts. In our assessment, we assumed a process improvements, such as fully renewable energy use in the pro­
pressure drop of 5 kPa in the NaOH bed, and 2 kPa in the other reference cess. The economic performance is in any case highly influenced by the
options. These pressure drops are responsible for a significant energy cost of sodium hydroxide production, which is in turn heavily affected
penalty of processes. An optimization towards the reduction of elec­ by the complex market of chlorine-based chemicals.
tricity consumption in the feed of air would efficiently reduce the carbon
footprint of air-based capture processes. For example, scenario 4, on the CRediT authorship contribution statement
use of renewable energy-based sodium hydroxide to capture CO2 from
air, yields a carbon footprint of + 1.43 kg of CO2 eq per kg of CO2. This Enrique Medina-Martos: Conceptualization, Methodology, Formal
value can be reduced to − 1.1 kg of CO2 eq if the pressure drop decreases analysis, Investigation, Writing – original draft, Writing – review &
to 2 kPa. The case of indirect calcium looping shows a positive footprint editing. José-Luis Gálvez-Martos: Conceptualization, Methodology,
for air capture (+0.50 to +0.57 kg CO2 per kg of captured CO2). This is Formal analysis, Investigation, Writing – original draft, Writing – review
different from the carbon footprint reported by Müller et al. [40] for the & editing. Jorge Almarza. Conceptualization, Investigation, Resources,
same technology, using KOH (− 0.59 kg CO2eq per kg of captured CO2), Writing – review & editing. Carlos Lirio. Methodology, Formal analysis,

12
E. Medina-Martos et al. Journal of CO2 Utilization 60 (2022) 101991

Investigation. Diego Iribarren. Supervision, Writing – review & editing, [21] M. Saeidi, A. Ghaemi, K. Tahvildari, P. Derakhshi, Exploiting response surface
methodology (RSM) as a novel approach for the optimization of carbon dioxide
Formal analysis. Antonio Valente. Conceptualization, Investigation,
adsorption by dry sodium hydroxide, J. Chin. Chem. Soc. 65 (12) (2018)
Writing – review & editing. Javier Dufour. Supervision, Investigation, 1465–1475.
Writing – review & editing. [22] C. Ruiz, L. Rincón, R.R. Contreras, C. Sidney, J. Almarza, Sustainable and negative
carbon footprint solid-based NaOH technology for CO2capture, ACS Sustain.
Chem. Eng. 8 (51) (2020) 19003–19012.
Declaration of Competing Interest [23] C. Ruiz, L. Rincón, R.R. Contreras, C. Sidney, J. Almarza, Sustainable and negative
carbon footprint solid-based NaOH technology for CO2capture, ACS Sustain.
Chem. Eng. 8 (51) (2020) 19003–19012.
The authors declare that they have no known competing financial [24] A. Zakeri, A. Einbu, P.O. Wiig, L.E. Øi, H.F. Svendsen, Experimental investigation
interests or personal relationships that could have appeared to influence of pressure drop, liquid hold-up and mass transfer parameters in a 0.5 m diameter
absorber column, Energy Procedia 4 (2011) 606–613.
the work reported in this paper. [25] S. Santos, J. Gomes, J. Bordado, Scale-up effects of CO2 capture by
methyldiethanolamine (MDEA) solutions in terms of loading capacity,
Technologies 4 (3) (2016 28) 19.
Appendix A. Supporting information
[26] B. Dutcher, M. Fan, A.G. Russell, Amine-based CO2 capture technology
development from the beginning of 2013—a review, ACS Appl. Mater. Interfaces 7
Supplementary data associated with this article can be found in the (4) (2015 4) 2137–2148.
online version at doi:10.1016/j.jcou.2022.101991. [27] X. Li, S. Wang, C. Chen, Experimental study of energy requirement of CO2
desorption from rich solvent, Energy Procedia 37 (2013) 1836–1843.
[28] T.A. Adams, Y.K. Salkuyeh, J. Nease, Processes and simulations for solvent-based
References CO2 capture and syngas cleanup. In: Reactor and Process Design in Sustainable
Energy Technology [Internet]. Elsevier; 2014 [cited 2018 Dec 19]. p. 163–231.
Available from: https://linkinghub.elsevier.com/retrieve/pii/
[1] IPCC. Global Warming of 1.5◦ C. An IPCC Special Report on the impacts of global
B9780444595669000065.
warming of 1.5◦ C above pre-industrial levels and related global greenhouse gas
[29] D.W. Keith, G. Holmes St, D. Angelo, K. Heidel, A process for capturing CO2 from
emission pathways, in the context of strengthening the global response to the
the atmosphere, Joule 2 (8) (2018) 1573–1594.
threat of climate change, sustainable development, and efforts to eradicate
[30] T. Brinkmann G.G. Santonja F. Schorcht S. Roudier L.D. Sancho Best. Available
poverty. 2018.
Tech. (BAT) Ref. Doc. Prod. Chlor-Alkali 1156 2014 344.Report No.: JRC9.
[2] F. Creutzig, C. Breyer, J. Hilaire, J. Minx, G.P. Peters, R. Socolow, The mutual
[31] F. Du, D.M. Warsinger, T.I. Urmi, G.P. Thiel, A. Kumar, V.J.H. Lienhard, Sodium
dependence of negative emission technologies and energy systems, Energy
hydroxide production from seawater desalination brine: process design and energy
Environ. Sci. 12 (2019) 1805–1817.
efficiency, Environ. Sci. Technol. 52 (10) (2018 15) 5949–5958.
[3] A. Azdarpour, A review on carbon dioxide mineral carbonation through pH-swing
[32] Eurochlor. Chlor-alkali Industry Review 2017/2018. 2018 p. 25.
process, Chem. Eng. J. (2015) 16.
[33] G. Wernet, C. Bauer, B. Steubing, J. Reinhard, E. Moreno-Ruiz, B. Weidema, The
[4] A. Sanna, M. Uibu, G. Caramanna, R. Kuusik, M.M. Maroto-Valer, A review of
ecoinvent database version 3 (part I): overview and methodology. Int J Life Cycle
mineral carbonation technologies to sequester CO2, Chem. Soc. Rev. 43 (23) (2014)
Assess. 2016 Sep 1;21(9):1218–30.
8049–8080.
[34] European Commission. Reference Document on Best Available Techniques for the
[5] J.L. Galvez-Martos, J. Morrison, G. Jauffret, E. Elsarrag, Y. AlHorr, M.S. Imbabi, et
Manufacture of Large Volume Inorganic Chemicals - Solids and Others industry.
al., Environmental assessment of aqueous alkaline absorption of carbon dioxide
[Internet]. 2007 [cited 2022 Jan 19]. Available from: https://eippcb.jrc.ec.europa.
and its use to produce a construction material, Resour. Conserv. Recycl. 107 (2018)
eu/sites/default/files/2019–11/lvic-s_bref_0907.pdf.
129–141 (Available from), 〈http://www.sciencedirect.com/science/article/pii/
[35] D. Pletcher, J.D. Genders, N.L. Weinberg, E.F. Spiegel, Methods for producing
S0921344915301567〉.
caustic soda without chlorine [Internet]. WO1993016216A1, 1993 [cited 2018 Dec
[6] M. Bui, C.S. Adjiman, A. Bardow, E.J. Anthony, A. Boston, S. Brown, et al., Carbon
10]. Available from: 〈https://patents.google.com/patent/WO1993016216A1/en〉.
capture and storage (CCS): the way forward, Energy Environ. Sci. 11 (5) (2018)
[36] J.L. Galvez-Martos, J. Morrison, G. Jauffret, E. Elsarrag, Y. AlHorr, M.S. Imbabi, et
1062–1176.
al., Environmental assessment of aqueous alkaline absorption of carbon dioxide
[7] J. Morrison, G. Jauffret, J.L. Galvez-Martos, F.P. Glasser, Magnesium-based
and its use to produce a construction material, Resour. Conserv. Recycl. 107 (2016)
cements for CO2 capture and utilisation, Cem. Concr. Res. 85 (2016) 183–191.
129–141.
[8] N. Mac Dowell, P.S. Fennell, N. Shah, G.C. Maitland, The role of CO2 capture and
[37] Y.M. Alhorr, A. Elhoweris, J. Morrison, J.-L. Galvez-Martos, The conversion of
utilization in mitigating climate change. Nature, Clim. Change 7 (4) (2017)
magnesium carbonates into plaster-like products: a preliminary study of the
243–249.
hardening mechanism, Adv. Civ. Eng. Technol. 2 (1) (2018) 6.
[9] R.S. Norhasyima, T.M.I. Mahlia, Advances in CO₂ utilization technology: a patent
[38] V. Castellani, E. Diaconu, S. Fazio, S. Sala, E. Schau, M. Secchi, et al. Supporting
landscape review, J. CO2 Util. 26 (2018) 323–335.
information to the characterisation factors of recommended EF Life Cycle Impact
[10] A. Rafiee, K. Rajab Khalilpour, D. Milani, M. Panahi, Trends in CO2 conversion and
Assessment methods: new methods and differences with ILCD. [Internet]. 2018
utilization: a review from process systems perspective, J. Environ. Chem. Eng. 6 (5)
[cited 2020 Aug 24]. http://publications.europa.eu/publication/manifestation_
(2018) 5771–5794.
identifier/PUB_KJNA28888ENN.
[11] J.-L. Galvez-Martos, A. Elhoweris, J. Morrison, Y. Al-horr, Conceptual design of a
[39] Facts Behind the Debat: Direct Air Capture [Internet]. 2019 [cited 2022 Jan 18].
CO2 capture and utilisation process based on calcium and magnesium rich brines,
https://ec.europa.eu/jrc/sites/default/files/factsheet_direct_air_capture_04.pdf.
J. CO2 Util. 27 (2018) 161–169.
[40] L.J. Müller, A. Kätelhön, S. Bringezu, S. McCoy, S. Suh, R. Edwards, et al., The
[12] A. Zimmerman, H. Naims, L. Müller, K. Armstrong, S. Michailos, A. Marxen, et al.
carbon footprint of the carbon feedstock CO2, Energy Environ. Sci. 13 (9) (2020)
Techno-Economic Assessment & Life-Cycle Assessment Guidelines for CO2
2979–2992. Sep 16.
Utilization [Internet]. The Global CO2 initiative at the University of Michigan;
[41] A. Kirchofer, A. Brandt, S. Krevor, V. Prigiobbe, J. Wilcox, Impact of alkalinity
2018 [cited 2019 Jul 18]. Available from: http://hdl.handle.net/2027.42/145436.
sources on the life-cycle energy efficiency of mineral carbonation technologies,
[13] I.E.A. Putting CO2 to Use: Creating value from emissions. 2019 Sep.
Energy Environ. Sci. 5 (9) (2012) 8631.
[14] R.M. Cuéllar-Franca, A. Azapagic, Carbon capture, storage and utilisation
[42] MITECO. NATIONAL INTEGRATED ENERGY AND CLIMATE PLAN (PNIEC) 2021
technologies: a critical analysis and comparison of their life cycle environmental
2030 [Internet]. Ministerio para la Transición Ecológica y el Reto Demográfico;
impacts, J. CO2 Util. 9 (2015) 82–102 (Mar).
2020 [cited 2020 Apr 1]. Available from: https://www.miteco.gob.es/images/es/
[15] H. Naims, Economics of carbon dioxide capture and utilization—a supply and
pniec_2021–2030_borradoractualizado_tcm30–506491.pdf.
demand perspective, Environ. Sci. Pollut. Res 23 (22) (2016) 22226–22241 (Nov).
[43] M. Mahmoudkhani, K.R. Heidel, J.C. Ferreira, D.W. Keith, R.S. Cherry, Low energy
[16] M. Kianpour, M.A. Sobati, S. Shahhosseini, Experimental and modeling of CO2
packed tower and caustic recovery for direct capture of CO2 from air, Energy
capture by dry sodium hydroxide carbonation, Chem. Eng. Res. Des. 90 (11) (2012)
Procedia (2009) 1535–1542.
2041–2050.
[44] N. von der Assen, P. Voll, M. Peters, A. Bardow, Life cycle assessment of CO 2
[17] R.V. Siriwardane, C. Robinson, M. Shen, T. Simonyi, Novel regenerable sodium-
capture and utilization: a tutorial review, Chem. Soc. Rev. 43 (23) (2014 20)
based sorbents for CO 2 capture at warm gas temperatures, Energy Fuels 21 (4)
7982–7994.
(2007) 2088–2097.
[45] C. Barker , INSIGHT: European chloralkali markets remain divided as high PVC
[18] V. Nikulshina, N. Ayesa, M.E. Gálvez, A. Steinfeld, Feasibility of Na-based
crushes caustic soda [Internet]. ICIS Explore. 2021 [cited 2021 Feb 24]. Available
thermochemical cycles for the capture of CO2 from air—thermodynamic and
from: https://www.icis.com/explore/resources/news/2021/02/19/10608293/
thermogravimetric analyses, Chem. Eng. J. 140 (1–3) (2008) 62–70.
insight-european-chloralkali-markets-remain-divided-as-high-pvc-crushes-caustic-
[19] S. Naeem, S. Shahhosseini, A. Ghaemi, Simulation of CO2 capture using sodium
soda.
hydroxide solid sorbent in a fluidized bed reactor by a multi-layer perceptron
[46] J. Koornneef, A. Ramirez, T. van Harmelen, A. van Horssen, W. Turkenburg,
neural network, J. Nat. Gas Sci. Eng. 31 (2016) 305–312.
A. Faaij, The impact of CO2 capture in the power and heat sector on the emission of
[20] S. Naeem, A. Ghaemi, S. Shahhosseini, Experimental investigation of CO2 capture
SO2, NOx, particulate matter, volatile organic compounds and NH3 in the
using sodium hydroxide particles in a fluidized bed, Korean J. Chem. Eng. 33 (4)
European Union, Atmos. Environ. 44 (11) (2010 1) 1369–1385.
(2016) 1278–1285.

13

You might also like