Atmosphere 13 01958 v3

Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

atmosphere

Review
Review of Carbon Capture and Methane Production from
Carbon Dioxide
Stephen Okiemute Akpasi 1, * and Yusuf Makarfi Isa 2

1 Chemical Engineering Department, Durban University of Technology, Durban 4000, South Africa
2 School of Chemical and Metallurgical Engineering, University of Witwatersrand,
Johannesburg 2050, South Africa
* Correspondence: [email protected] or [email protected]; Tel.: +27-717-770-571

Abstract: In the last few decades, excessive greenhouse gas emissions into the atmosphere have
led to significant climate change. Many approaches to reducing carbon dioxide (CO2 ) emissions
into the atmosphere have been developed, with carbon capture and sequestration (CCS) techniques
being identified as promising. Flue gas emissions that produce CO2 are currently being captured,
sequestered, and used on a global scale. These techniques offer a viable way to encourage sustain-
ability for the benefit of future generations. Finding ways to utilize flue gas emissions has received
less attention from researchers in the past than CO2 capture and storage. Several problems also need
to be resolved in the field of carbon capture and sequestration (CCS) technology, including those
relating to cost, storage capacity, and reservoir durability. Also covered in this research is the current
carbon capture and sequestration technology. This study proposes a sustainable approach combining
CCS and methane production with CO2 as a feedstock, making CCS technology more practicable. By
generating renewable energy, this approach provides several benefits, including the reduction of CO2
emissions and increased energy security. The conversion of CO2 into methane is a recommended
practice because of the many benefits of methane, which make it potentially useful for reducing
pollution and promoting sustainability.
Citation: Akpasi, S.O.; Isa, Y.M.
Review of Carbon Capture and Keywords: adsorption; absorption; chemical looping; CO2 capture; CO2 conversion; methane; membrane
Methane Production from Carbon
Dioxide. Atmosphere 2022, 13, 1958.
https://doi.org/10.3390/
atmos13121958 1. Introduction
Academic Editors: Hung-Lung Chiang, Energy plants, industries, as well as other sources of carbon dioxide (CO2 ) result
Vanessa Bach and Federica Raganati in global warming and affect the planet [1]. The temperature of the Earth’s surface has
Received: 28 August 2022
risen by 0.8 ◦ C in tandem with an increase in CO2 concentration of 280 to 400 ppm [2]. In
Accepted: 8 November 2022
the last century, the CO2 level has increased slightly to 408.8 ppm [3], with projections of
Published: 24 November 2022
600 to 700 ppm, raising average surface temperatures by 4.5–5 ◦ C [4]. Unstable financial,
technological, and sociological advances, along with natural and human developments,
Publisher’s Note: MDPI stays neutral
are all important factors affecting the CO2 concentration [5]. In addition to CO2 , other
with regard to jurisdictional claims in
greenhouse gas (GHG) emissions have significantly increased in recent years, such as sulfur
published maps and institutional affil-
hexafluoride (SF6), perfluorocarbons, hydrofluorocarbons (HFC), nitrous oxide (N2 O),
iations.
and methane (CH4 ) [1,6]. According to the International Panel on Climate Change, GHG
emissions should be lessened by 50–80% to prevent our planet from disintegrating [2].
Almost 190 countries met in Paris in December 2015 (Conference of Parties—COP 21) to
Copyright: © 2022 by the authors.
agree on limiting global temperatures to 2 ◦ C by the end of the 21st century. COP 21 pro-
Licensee MDPI, Basel, Switzerland. posed several techniques, including increasing energy efficiency, utilizing renewable fuels,
This article is an open access article and implementing geoengineering techniques such as afforestation and, more specifically,
distributed under the terms and developing carbon capture and storage (CCS) technologies [3,4]. Therefore, carbon capture
conditions of the Creative Commons and storage (CCS) is generally acknowledged as a potential, if not suitable, method to
Attribution (CC BY) license (https:// minimize CO2 emissions globally [7,8].
creativecommons.org/licenses/by/ An estimated 60% of CO2 emissions are attributed to many stationary sources, includ-
4.0/). ing industrial plants, fossil fuel power plants and thermoelectric cement plants, iron and

Atmosphere 2022, 13, 1958. https://doi.org/10.3390/atmos13121958 https://www.mdpi.com/journal/atmosphere


Atmosphere 2022, 13, 1958 2 of 42

steel mills, gas processing industries, refineries, and power plants. These industries serve a
variety of sectors such as the industrial sector, transportation, and electricity generation [9].
According to the above analysis, stationary emission sources will continue to be the greatest
drivers of greenhouse gases (GHGs) resulting from the burning of fossil fuels for several
decades to come. The production of electricity is responsible for 25% of all emissions, while
forestry and agriculture account for 24% [10]. Thermal power plants typically use coal
as fuel, because it is cheap and readily available, yet it is considered a major emission
source, with emissions approaching 2249 lbs/MWh [11]. As a result, new and improved
technologies such as CCS for swift CO2 removal from the atmosphere to tackle global
warming have received a lot of attention lately. Figure 1 illustrates the various energy
sectors contributing the most to CO2 emissions.
The type of combustion process utilized, as well as the volume of CO2, absorbed,
separated, and transported for reuse or storage, are the primary determinants of CCS [4].
CCS can reduce CO2 emissions from power plants to limit global warming to 2 ◦ C [4].
Post-combustion is the most advanced of the three major CCS technologies, followed by
pre-combustion and oxy-fuel combustion. In terms of CO2 capture, coal gasification is
the best option; however, it cannot be used in gas-fired power plants. It makes use of
most of the existing power plant infrastructure as a retrofit technology. There has been a
demonstration of the ability to recover 300 t/day of CO2 on a small scale [4].
During oxy-fuel combustion, pure oxygen replaces air, thereby eliminating the require-
ment for flue gas de-NOx and the incorporation of 80–98 ◦ C CO2 in the exhaust, resulting
in a more effective and reliable CO2 removal method. Although the process requires a lot of
energy, corrosion is aided by SO2 concentration due to the increased energy consumption [12].
Gasification under low O2 pressure in pre-combustion CCS produces syngas (CO + H2 ) from
the fuel pretreatment [4]. In the presence of high CO2 concentrations in fuel gas, separation is
less difficult than with other methods, resulting in the formation of H2 as a transport medium
for energy. This helps to offset some of the technology’s operational costs [12].
The purpose of this paper is to provide an overview of the state of carbon capture
systems, both conventional and the emerging, and present CO2 as a feedstock for producing
methane (CH4 ). It further concentrates on improving the concept and fostering research
on CO2 capture and utilization to circumvent the storage issues with the current CCS
technology, and make sustainability a priority for future generations.
There are currently no studies that address all the gaps; however, this paper covers
CO2 capture, sequestration, utilization, and CO2 methanation.

2. Methodology
The focus of this research on the literature was intended to identify the benefits
and drawbacks of current carbon capture systems and porous carbon-based adsorbents
for CO2 adsorption.
In addition, the challenges and prospects for CO2 conversion into methane were
discussed. The manuscript’s section on methodology describes how the literature was
found, gathered, and then arranged by identifying the knowledge gaps that exist in earlier
studies. Using scientific databases and a variety of search engines, such as NCBI, Google
Scholar, Scopus, Science Direct, Web of Science, and major publishers, an in-depth analysis
of the articles and literature from peer-reviewed journals was carried out. The search for
relevant literature was constrained to studies from the years 2000 to 2022 to draft this
manuscript. The limited literature search for this manuscript’s preparation was mostly
focused on studies between the years 2000 and 2022. For this paper’s main theme, several
pertinent keywords were employed in the literature search. These keywords include zeolite,
activated carbon, graphene, silica, and metal–organic framework, as well as CO2 capture
and conversion. Research articles with the most pertinent material were selected after the
articles were evaluated and the information from the abstracts was reviewed. To prevent
translation challenges, articles published in languages other than English were omitted.
Atmosphere 2022, 13, 1958 3 of 42

3. CO2 Capture Technologies


CCS technology relies on CO2 capture. Three methods for trapping and sequestering
CO2 are envisaged, based on the settings of a fossil fuel-based power plant, gas steam
pressure, and CO2 partial pressure. Among the carbon capture technologies are pre-
combustion, oxy-fuel combustion, and post-combustion [13].

3.1. Pre-Combustion Capture Method


This method involves capturing CO2 before combustion, rather than after combustion,
thus making this method more feasible. Gasifiers and catalytic reactors are involved. In a
gasifier with low oxygen pressure, the following equations (Equations (1)–(3)) demonstrate
how pure syngas (CO, H2 ) is produced. Moreso syngas is enriched by passing it through a
shift reactor with a steam, as shown in Figure 1.

2C + 3H2 O → 3H2 + CO + CO2 (1)

C + H2 O → H2 + CO (2)
CO + H2 O → CO2 + H2 (3)
Natural gas could also be applied for steam reforming via this method; between 700
and 850 ◦ C, an endothermic reaction occurs, culminating in the conversion of CH4 and
H2 O and the generation of syngas (CO, H2 ,). An exothermic reaction occurs after partial
oxidation, in which CH4 burns in oxygen to create CO2 (Equations (4) and (5)) [14].

CH4 + H2 O → CO + 3H2 (4)

2CH4 + O2 → CO2 + 4H2 (5)


A key element of pre-combustion is the hydration of gas separated from air, which is
the most promising technology today [15].
The CO2 captured in this technique produces hydrogen fuel, which is used to generate
electricity in several industries while emitting little CO2 [16]. Fertilizers and hydrogen
are produced through pre-combustion capture [17]. CO2 is generated from this process in
concentrations ranging from 5 to 60%, making it easy to capture; however, the water gas
shift reactions and gasification are costly and difficult to operate [18].

3.2. Oxy Fuel Combustion Capture Method


A system for air separation during oxy-fuel combustion capture is used to obtain pure
oxygen, as flue gas and coal are fed into the oxy-fuel boiler [19–21]. Coal is burned at a
fixed temperature and pressure. Figure 1 illustrates how the flue gases, lacking oxygen
and nitrogen, are only made up of CO2 and water vapors. Additionally, this reduces the
quantity of flue gas required to remove NOx and sulfur [22,23]. Condensing water vapor
maintains the boiler’s temperature. Dehydrated CO2 is compressed and delivered to a
storage facility or an industry where it is utilized in beverage carbonation, as a solvent, and
as a fire extinguisher [24]. Flue gas is recycled in its natural state to maintain the boiler’s
internal temperature [20]. When compared to other carbon capture technologies, it is the
most efficient and adaptable [25]. There are, however, some obstacles to overcome. The air
separation unit (ASU) uses cryogenic distillation, which consumes a great deal of energy to
produce essentially pure oxygen. The integrated gasification combined cycle (IGCC) can be
replaced by this method, which converts CO to CO2 .
A major drawback of this method is that it demands more energy to separate oxygen
from the air [4,26].

3.3. Post Combustion Capture Method


After burning carbonaceous materials (such as biomass) or fossil fuels, CO2 is removed
from the flue gas using a viable adsorbent such as amine [4,27]. Currently, it is the most
effective method of capturing CO2 from natural gas plants. As separating agents, it
Atmosphere 2022, 13, 1958 4 of 42

employs amine solutions including potassium carbonate (K2 CO3 ) mono-ethanolamine


(MEA), methyl-di-ethanolamine (MDEA), and di-ethanolamine (DEA) [14]. The novel
sorbent piperazine collects contaminants efficiently because of its high volatility [28]. A
chemical reaction deposits CO2 in the solution, which is then separated by flowing high-
temperature (100–200 ◦ C) steam over it.
An equation representing the reversible absorption–regeneration reaction is
shown below [14].

C2 H4 OHNH2 + H2 O + CO ↔ C2 H4 OHNH3+ + HCO3− (6)

However, this method has some drawbacks, such as a high parasitic load, high flue
gas temperatures, and the inability to operate at low CO2 levels. Figure 1 illustrates a direct
post-combustion mechanism. Gas-fired power plants use about 4% CO2 , while coal-fired
power plants utilize 7–14%, increasing electricity costs by 32% and 65%, respectively, in
gas-fired and coal power plants [4]. As a result, cutting costs while improving system
performance is still a subject of great interest [29]. Membrane technology, including ceramic
or polymeric membranes, marine algae, and cryogenic distillation, can all be used for
post-combustion CO2 capture. This has many benefits, including the ability to retrofit and
eliminate the need for a cryogenic separator and shift reactor. In post-combustion, CO2 is
usually absorbed or desorbed using a solution containing alkanol amine [30–32].

Figure 1. CO2 capture technologies [33].

4. Various Combustion Technologies for CO2 Capture


A comparison of the three CO2 capture technologies is illustrated in Table 1. Coal
gasification uses pre-combustion, while oxy-fuel combustion and post-combustion are
applied to coal- and gas-fired plants. CO2 capture using post-combustion technology
is presently the most advanced [28,34]. According to Gibbins and Chalmers [35], three
technologies were compared in terms of the costs of both gas-fired and coal-fired power
plants (Table 2). CO2 mitigation for coal-fired power plants was the cheapest using pre-
combustion technology, while oxy-fuel and post-combustion technologies were comparable.
For post-combustion capture, the costs per tonne of avoided CO2 were 50% lower for gas-
fired plants than for the two other technologies. Moreover, the most inefficient method
of CO2 capture is post-combustion CO2 capture, having an energy cost of roughly 8% for
coal-fired plants and 6% for gas-fired plants [36].
Atmosphere 2022, 13, 1958 5 of 42

Table 1. Comparison of various CO2 capture technologies [37].

Parameter Post-Combustion Pre-Combustion Oxy-Fuel Combustion


Integrated gasification Synergy between novel cycle
Can be utilized with current combined cycle and turbines, and integrated gasification
Application area
coal combustion plants. which can effectively use combined cycle
H2 -rich syngas. has been employed.
Lower energy penalty than
Small concentrations of CO2 can
post combustion methods. The CO2 capture efficiency is
be captured.
Pros The pressure and temperature 100%. Absence of
Possible to retrofit to
of the regeneration process hazardous NOx.
existing plants.
can be altered.
Syngas treatment and drying
High costs for operation and High operational and capital
before CO2 capture. Capital
Cons regeneration. Excessive solvent costs. Annexing to existing
costs and investment
losses. plants is difficult.
costs are high.
CO2 concentration (vol.%) 4–14 15–40 75–80
CO2 capture efficiency (%) 85–90 85–90 90–100
CO2 purity (%) 99.6–99.8 95–99 87.0–94.8
Cryogenic temperatures are
used to separate oxygen. The
Flue gas must be cooled, and High pressure and low
flue gas is recycled to lower
Temperature and pressure pressure is dependent on the temperature (depending on
the temperature because
CO2 capture process. the technique utilized).
oxy-fuel combustion produces
high temperatures.
Large size equipment required
Equipment size Medium equipment. The equipment is of low size.
with high investment.
Steam/air is required for High-purity oxygen
Combustion medium Air is used.
gasification to generate CO2 . for combustion.
Amine scrubbing plants
Integrated gasification
(reaction with
combined cycle and ammonia
Start of the art monoethanolamine) are in Efficient CO2 separation.
production plants
practice. Currently, power
are running currently.
plants use this technology.
NOx is not present; however,
Contains H2 S, COS, Sulfur compounds
Acid gases gas desulfurization
NOx, and SOx. need to be removed.
is necessary.

Table 2. Comparison of the costs of various capture methods [35,37]. Costs exclude storage and
shipping costs, but include CO2 compression to 110 bars.

Capture Technology
Oxy-Fuel
Fuel Type Parameter Pre-Combustion Post-Combustion No Capture
Combustion
Capital cost (USD/kW) 1180 870 1530 500
Cost of CO2 avoided (USD/t CO2 112 58 102 -
Gas-fired
Cost of electricity (c/kWh) 9.7 8.0 10.0 6.2
Thermal efficiency (% LHV) 41.5 47.4 44.7 55.6
Capital cost (USD/kW) 1820 1980 2210 1410
Cost of CO2 avoided (USD/t CO2 23 34 36 -
Coal-fired
Cost of electricity (c/kWh) 6.9 7.5 7.8 5.4
Thermal efficiency (% LHV) 31.5 34.8 35.4 44.0

5. CO2 Separation Techniques


CO2 is separated from flue gas during combustion using a variety of advanced separa-
tion techniques [4]. There are several techniques involved, such as absorption, adsorption,
chemical looping, membrane separation, and cryogenics [18,38,39]. Figure 2 displays a
flowchart outlining the methods and techniques for CO2 capture/separation technologies.
Atmosphere 2022, 13, 1958 6 of 42

Figure 2. Separation techniques for CO2 capture [40].

5.1. Absorption
The ability of absorption to capture huge quantities of emissions from chemical fac-
tories and power plants has gained considerable attention in recent years. Chemical
absorption is a reliable technique for CO2 separation in coal-fired power plants because it
is well-suited for existing plants with high operating costs and limited infrastructure [41].
The chemical absorption of CO2 is a commercially viable technology due to its many ad-
vantages, including technical efficiency, handling capacity, and sophistication [42]. The
potential absorbents and processes of absorption CO2 capture are highlighted in Table 3.
CO2 is separated from flue gas by absorption using a liquid sorbent [3,43]. It is
possible to regenerate the sorbent via a regenerative process or stripping by depressurizing
and/or heating. This is the latest and most advanced method for separating CO2 [44].
Potassium carbonate (K2 CO3 ), monoethanolamine (MEA), and diethanolamine (DEA) are
examples of common sorbents [45]. MEA is very reactive and absorbs more quickly, and it
is quite inexpensive [44]. However, their main drawback is the substantial parasitic energy
load in relation to solvent regeneration, which adversely affects the total effectiveness of
systems combined with aqueous amine-based absorption processes [46]. DEA and other
alkanolamines have also been employed for absorption, although they have comparable
defects. Methyldiethanolamine (MDEA), a mixture of MEA and DEA, has been used with
moderate success. It has higher CO2 loading capacity, and degradation and corrosion
resistance, as well as cheaper regeneration costs, but lower rates of absorption [47–52].
Veawab et al. [53] reported that MEA is the most efficient aqueous alkanolamine
for CO2 absorption, with a performance rate greater than 90%. Additionally, Aaron and
Tsouris [54] reviewed various CO2 capture technologies and determined that MEA absorp-
tion is the most viable method for CO2 capture in CCS. Applying a solvent containing
30% MEA, a 1 t CO2 /h absorption pilot plant was designed and experimentally validated
in conjunction with a coal-fired power plant’s post-combustion capture technology [55]. In
recent times, other adsorbents, including anion-functionalized ionic liquid and piperazine,
have attracted a lot of attention [56]. Even though piperazine rapidly reacts compared
to MEA, its use in CO2 absorption is more costly. Due to its higher volatility, it is still in
the experimental phase [28]. The risk of amine degradation, which could lead to equip-
ment corrosion, solvent loss, and the formation of volatile degradation compounds, is a
significant barrier to the widespread adoption of this technology for the CCS [57,58], while
environmental degradation has gone unnoticed.
Atmosphere 2022, 13, 1958 7 of 42

Furthermore, amine emissions can deteriorate into nitramines and nitrosamines, which
are highly toxic to human health and the environment. The chilled ammonia process
captures CO2 using aqueous ammonium salts (including ammonium carbonate) and can
regenerate the CO2 at elevated temperatures and pressures using waste heat, thereby
minimizing the downstream compression [59]. There are fewer problems with this process
than those caused by amine degradation.
Water’s use as a co-solvent, which has higher thermal characteristics than other co-
solvents, is one of the key precursors for the high solvent regeneration energy of MEA [20].
In the context of CO2 absorption, the predicted regeneration energy for 30 wt. % aqueous
MEA showed that more than 50% of the total energy was used to heat and vaporize the wa-
ter co-solvent. The remaining energy was used to reverse the chemical interaction between
CO2 and MEA at the same time [60,61]. Considering this, it was thought that either totally
or partially substituting other organic diluents for water as co-solvents could potentially
reduce solvent regeneration energy, since they effectively create water-free/water-lean
hybrid solvents with poorer thermal properties than water [62–68]. Instead of vaporizing
and heating the co-solvent, comparable to aqueous amines, the regeneration energy will be
used more effectively to reverse acid gas chemisorption.
Additionally, hybrid water-free/water-lean solvents have been thoroughly studied
in recent years, primarily for their CO2 capture applications [68–70]. They provide a wide
range of potentially alluring substitutes to conventional aqueous amines [46,71]. The main
objective of water-lean solvents is to preserve the chemical selectivity of water-based sol-
vents, while enabling step gains in efficiency due to the lower specific heats of organics
than water [46]. However, two problematic regions refute the claim of their attractiveness.
The stated performance of these solvents when scaled up from lab-scale to industrial-scale
settings has not been adequately examined due to a lack of availability of a few essential
properties. This is predicted given the labor-intensive nature of experimental work, which
makes it impossible to expand experimental testing to the broad range of transport and
thermophysical parameters needed for precise and representative performance evaluation
on an industrial scale. The second issue is that, when carried out on a lab scale, the poten-
tiality of a particular solvent is typically demonstrated using a limited set of parameters,
most notably the low enthalpy of absorption and high absorption capacity [72–75]. These
two characteristics are indeed of great concern for chemical absorption procedures, but
they are still unsuitable for accurately gauging the potential of the tested solvents for their
intended use. However, they ignore significant trade-offs between competing environmen-
tal, economic, and operational factors. The results of a straightforward assessment can help
direct the development of novel generating solvents [76,77].
However, these difficulties can be overcome if the proper tools or novel process
configurations are available. Due to recent developments in computational power and
thermodynamic modeling tools, the first issue can be resolved by scaling the data from
lab to industrial operating conditions. The most appealing models for this application are
molecular equations of state (EoSs) centered on the Statistical Associating Fluid Theory
(SAFT) [78,79], due to their strong theoretical background, demonstrated correctness for a
range of complex systems, and predictive abilities.
The solution to the second problem, which is to demonstrate the viability of a chosen
solvent typically acknowledged using a limited number of requirements, may appear
relatively apparent: add more evaluation criteria to the already-existing standard key per-
formance indicators (KPIs). Moreover, making such a preference is more difficult because
early design phases may not have access to information on a particular criterion [80]. There
must be a justification for why certain criteria should be included or excluded when narrow-
ing the search space among the numerous properties that are available [81–83]. The effect
of solvent characteristics on economic metrics such as total capital expenditures (CAPEX)
and operating expenditures (OPEX) typically serves as the foundation for justification.
Through the careful process modeling of hypothetical solvents, Mota-Martinez et al. [84]
ranked various solvent characteristics according to how they affected the process’ overall
Atmosphere 2022, 13, 1958 8 of 42

economics. Leclaire and Heldebrant [43] recently recommended the use of ideas from
green engineering and chemistry to address problems with the advancement of CCUS
technologies. They asserted that by applying the 12 + 12 principles of engineering and green
chemistry [85], they could indirectly encourage the improvement of chemical processes’
economic attractiveness and efficiency, which goes beyond their environmental motivation.
Similarly to this, it may be beneficial to consider sustainability, health, and safety issues
while assessing the possibility of promising solvents for the removal of acid gas [86,87].
Figure 3 displays the schematic diagram for the absorption carbon capture process.

Table 3. Summary of absorption-based carbon capture.

Reactive Operating Conditions CO2 Capture (%),


Type Absorbent Kinetics/Mass Transfer Ref.
Separator P, C, T, G AC (kg/kg)
C:8 −16; T:10–40; C2 H4 OHNH2(1) + CO2(g) + H2 O
MEA Flow (SC) 94, 0.4 [88–90]
G:2–10 ↔ C2 H4 OHNH3+(aq) + HCO3−(aq)
Fixed-bed
K2 CO3 (Con-O, T:60 G:40 mL/min 99.4, NA NA [91,92]
bench scale)
2NH3(g) + CO2(g) ⇐⇒ NH2 COONH4(s)
Single solvent Ammonia Sieve plate (CC) C:10–14; T:25–55 ◦ C 95–99, 1.2 NH2 COONH4(s) + H2 O(g) ⇐⇒ [88,89]
(NH4 )2 CO(s)CO3(s)
Stirred cell
Piperazine P:0.032 T:42 and 0.042 100, 0.32 1st order partial reaction occurs [93]
(SC, BS)
Double stirred T:25–50; P:0.1;
Ionic liquids 99.11 at 60 ◦ C, NA [94–97]
cell (BS) A:0.5–1.2
Promoter selection is very critical. It
Split flow
DEA-K2 CO3 T:115 L:63.66 m3 /h 99, NA is a reversible exothermic reaction [98]
(CC, bench scale)
CO2 + K2 CO3 + H2 O ↔ 2KHCO3
PEI-SiO2 Alco- Packed CP ∆msolution
L:33.66 m3 /h NA, NA qsensible = ∆mCO2
[99,100]
Mixed Solvents hol/amine/water (bench scale)
46 (HCL), 48
T:40 (absorption) T:90
(HCC), 11(HCE) Carbamate and bicarbamate
BDA-DEEA Packed (CC, BS) (desorption) [101]
than MEA formations
G: 24.78 m3 /h
with 5 M
L/G:2.9; packing
AMP-PZ Packed (pilot) 90, NA - [102–109]
height=10 m

Figure 3. Schematic of an absorption carbon capture process [110].

5.2. Adsorption
The process of adsorption [43] involves molecules in liquids and gases adhering to
solid surfaces by weak van der Waals interactions. Unlike liquid absorbent processes, solid
adsorbents bind CO2 to their surfaces during adsorption. Selection criteria for this sorbent
include a large surface area, high regeneration capability, and high selectivity. Common
Atmosphere 2022, 13, 1958 9 of 42

adsorbents include activated carbon, molecular sieves, zeolites, lithium zirconate, and
hydrotalcite [27]. Table 4 highlights the potential adsorbents and adsorption parameters
for CO2 capture.
It is possible to achieve CO2 adsorption by changing the pressure or temperature of a
saturated sorbent. Pressure swing adsorption (PSA) is a commercially applied technology
that recovers more than 85% of CO2 from power plants [111,112]. A solid adsorbent
selectively adsorbs CO2 at high pressures, then the solid desorbs, releasing CO2 for low-
pressure transport (usually atmospheric pressure). The temperature swing adsorption
(TSA) releases the CO2 in the system by increasing its temperature through steam injection
or hot air distribution [113]. A CO2 purity of over 95% and recovery of over 80% are
possible when using CO2 regeneration, although regeneration is more time-consuming
than PSA [114]. It was estimated that the operating costs of a particular TSA process ranged
between USD 80 and 150 per tonne of CO2 captured [115]. Significant attention has been
paid to developing CO2 capture sorbents from agricultural and industrial wastes to lower
the overall cost of CO2 capture. An adsorption carbon capture process is shown in Figure 4.

Table 4. Summary of adsorption-based carbon capture.

Reactive Operating Conditions CO2 Capture (%),


Adsorbent Kinetics/Mass Transfer Ref.
Separator P, T, C, G Ad-C (gCO2 /gads)
4–4.9 wt. %, 8.31
TEPA-Mg- Regeneration temp is N2 adsorption–desorption
PBR (LS) mmol CO2 /g [109]
MOF-74 250–300 ◦ C isotherm
absorbent, NA
PBR (3-bed,
85–95, NA, 73–82% Langmuir adsorption
ZX-APG, 8-step, T:35; P: 0.007–0.008 [116]
CO2 purity isotherm is adopted
VPSA, LS)
Water vapour (H2O): 4.6
mol%, Vf: 44; TDes:100T:
Activated PBR (1 bed, 3 60, ICC:11.2, Bd:0.493, Dual-site Langmuir
69.5, NA [117]
carbon step, VSA, LS) Lg:50, P:0.113, PVP = 3, equation has been adopted
Trpt:3; SA:921.7,
PV:0.37, Tads:35
Langmuir adsorption
NPC10 PBR (TSA, LS) T: 25, P: 0.1, SA: 639 NA, 0.041 [118]
isotherm
CO2 + 2RR0 NH ↔
Fly ash + RR0 NCOO− + RR0 NH2+ CO2
PBR (LS, TSA) St: 24 h, P: 0.11, T: 70 4.5 at 85 ◦ C [119]
PEI + PEG +2H2 O + RR0 NCOO− ↔
2HCO3− + RR0 NH2+
Bed dimensions (m):
FRR: 0.5; CT: 650; AT: 950;
SA: 1873.9; 2b: 0.03, Nm:
MBA (LS, 36, W: 1.5; L:1.5; Xpth: Extended Langmuir
ZX 80, NA, 97% purity [120]
PSA) 0.012 isotherm was used
Bd: 0.65, Cs: 1.07, Dp: 3420,
ε: 0.31
ks: 0.275
Langmuir adsorption
Rayon–HCM PBR (TSA) 97, 0.2 [121,122]
isotherm adopted
Atmosphere 2022, 13, 1958 10 of 42

Figure 4. Schematic of an adsorption carbon capture process [110].

5.3. Chemical Looping Combustion


In contrast to oxy-fuel combustion, which uses pure oxygen for combustion, metal
oxides are used as oxygen carriers in combustion. Metal oxides are reduced to metal during
the process, while fuels are oxidized to create CO2 and water. In a subsequent stage, the
metal is oxidized and recycled. The removal of water by condensation from the process
2
byproducts is easy, but the separation of pure CO2 requires no energy. Numerous low-cost
metal oxides, including Mn2 O3 , CuO, NiO, and Fe2 O3 , are suitable for this process. The
potential sorbents and processes of chemical looping combustion are highlighted in Table 5.
Several researchers [123–127] have examined the performance efficiency of various
metal oxides in this process. According to Adánez, de Diego [126], a metal oxide can be
optimized by using support inert materials, but the selection of an inert material will vary
depending on the characteristics of the metal oxide. Chemical looping combustion (CLC)
was studied by Lyngfelt, Leckner [128] in a boiler consisting of two fluidized beds. Lyngfelt,
Leckner [128] recently reviewed this technology. This process has been demonstrated
to be a very promising CO2 capture technology by both Lyngfelt, Leckner [128] and
Adánez, de Diego [126]. The IGCC’s CO2 separation is based on pre-combustion, but
Erlach, Schmidt [113] found chemical looping combustion to have a 2.8% higher net plant
efficiency than the former method. Figure 5 illustrates the basic CLC system.

Table 5. An overview of chemical looping combustion-based carbon capture.

Operating
Reactive CC (%), Kinetics/Mass
Fuel Type Conditions Challenges Ref.
Separator Purity (%) Transfer
P, T, C, G
Combustion of iso-octane
T: 200–1200; The ∆Hr is dependent on
Coal, C2 H5 OH, NA, 93 (with (−5101.58 kj/mole) with Na2 SO4 and
molar ratios of the fuel but not the
Isooctane, C3 H8 TGA CaSO4 at CaSO4 produces without SO2 [129,130]
carbon/CaSO4 = 0.5 amount of OC utilized.
and CH4 . 850–975 ◦ C) formation between 200 ◦ C
and carbon/steam = 1 The yield depends on OC.
and 344.3 ◦ C.
Atmosphere 2022, 13, 1958 11 of 42

Table 5. Cont.

Operating
Reactive CC (%), Kinetics/Mass
Fuel Type Conditions Challenges Ref.
Separator Purity (%) Transfer
P, T, C, G
The packed bed of OC reduces the
need for highly efficient cyclone to
reduce costs; boron nitride (BN) used
XOC: 80–95, HR: 2-stage PBR- PP of O2 in reactors; high
Syngas, H2 100, NA as the dense support material due to [131–134]
90–99, T: 370–1030 CLC solid inventories.
high thermal conductivity, low
thermal expansion and
high thermal stability.
83–99.3% at Scale-up, fuel conversion,
Coal, kerosene, Bd: 4.750; Dp: 128 ∆PRC increases linearly with solid
IFBR 800–950 ◦ C, agglomeration [135–141]
biomass Umf: 0.0129, Φ: 0.64 flow rate.
NA and attrition.
The reduction kinetics and activation
Iron oxide: 950 ◦ C, CMBS or Reaction heat exceeds the energy parameters are critical to find
CH4 , coal FF: 1.18, CO2 EF: 10, RPBR >99, >95 convective heat-transfer fuel conversion efficiency, [142,143]
DT: 5.25 (1 MWth) rate to the gas flow. temperature distribution and carbon
separation efficiency.
T: 700–975; >99% CH4 and The formation of FeO and At 900 ◦ C, the reduction of Fe2 O3 to
SITC:20–30; SFRR: 100% syngas FeAl2 O4 indicates further Fe with CO generates 37.7 KJ/mol
CH4 , syngas 8–10 for CO CC-MBR conversion. utilization of oxygen in Fe2 O3 of heat but its reduction with [143,144]
SFRR:4–12 for H2 >99.99% H2 iron-based OC’s can be H2 gas needs 61.8 KJ/mol
Fsolids:1.7–2.5 purity. achieved. –φ > 1.14. Fe2 O3 of heat.

Figure 5. Schematic of chemical-looping combustion (adapted from Yang [145]).

5.4. Membrane Separation


Membrane separation uses a semipermeable membrane or barrier to physically sepa-
rate CO2 from other flue gases [146]. Membrane separation uses less energy than traditional
solvent absorption methods, making it less expensive [147]. Membrane separation has
successfully been used for selective gas separation in a variety of fields for the past two
decades, including natural gas sweetening, air separation, hydrogen production, and bio-
gas upgrading. Researchers are working on developing membrane-based materials to
separate CO2 released by various industries. Furthermore, this technology has produced
increased efficiency in terms of both the economy and the environment [148]. Scientists
have developed a variety of different membranes for CO2 separation, including inorganic
membranes, polymers, carbon molecular sieve membranes (CMSMs), microporous organic
polymers (MOPs), and mixed matrix membranes (MMMs) [149]. The potential sorbents
and processes of membrane separation are highlighted in Table 6.
In addition, membrane separation technology can also separate gases in CCS processes
such as pre- and post-combustion capture. It is generally considered that polymeric
membranes are more flexible, durable, and efficient at capturing CO2 from industrial
processes. An upper bound relationship analysis describes how selectivity and permeability
are related to CO2 capture by polymeric membranes [150]. To improve results, glassy and
rubbery materials with varying separation principles based on their size and diffusion
Atmosphere 2022, 13, 1958 12 of 42

ability can be used to synthesize polymeric membranes. The condensability and differences
in kinetic properties of gas molecules are responsible for gas separation by glassy and
rubbery polymers [151]. Considering how difficult it is to examine operating conditions for
rapid performance, membranes applied in gas separation systems are typically modeled to
determine their working capacity [152]. For optimal results in industrial settings, membrane
performance must not be interfered with by flue gas impurities [149]. Researchers were able
to separate CO2 from other gases with an efficiency of 82–88% [153,154]. In fact, despite
membrane materials having poor permeability and selectivity [155], it is also problematic
to use this extraction method in flue gas with low pressure and CO2 concentration in flue
gas conditions [156]. A membrane carbon capture process is displayed in Figure 6.

Table 6. Summary of membrane-based carbon capture.

Reactive Kinetics/Mass
Membrane Operating Parameters Challenges Ref.
Separator Transfer
Solution–diffusion; among
S-P, T, P, La , pressure ratio of
Hollow fiber and Lower selectivity at the mechanisms are
Dense membranes the permeate side to the feed [157]
flat-sheet higher permeability Knudsen diffusion and the
side, pore size and porosity
molecular sieve effect
P, T, pore size and ε of the
Reaction kinetics depend
membrane–membrane Wetting of the membrane [157]
on solvent
wettability
Micro-porous Hollow fiber and
Even at high pressures, Ko
Membranes flat-sheet There are other compounds
Gas flow area is controlled by the [158]
present in the gas stream
resistance of the liquid film
Solvent volatility and limited Pore diffusion depends on
Liquid flow area [159]
long-term stability membrane support
Solvent “wash-out” causes
Liquid in the The overall mass
Flat-sheet only Ga, La, VVIS, P, T the membrane’s stability [160]
membrane pores transfer coefficient
to decrease

Figure 6. Schematic of membrane carbon capture process (adapted from Wang [161]).

5.5. Cryogenic Distillation


This process separates CO2 from gas mixtures by focusing on their boiling points
at temperatures ranging from 100 to 135 ◦ C [6,43]. In the presence of high pressures
(100–200 atm), solidified CO2 provides two significant benefits: a lack of solvents and
liquefied CO2 for more convenient transport and injection [37]. It does, however, have
some drawbacks that need to be investigated further, as do other processes. When cold and
pressurized nitrogen is used as a refrigerant, ice formation compromises equipment safety,
causing pressure fluctuations and pipe blockages, as well as increasing the consumption
of energy [37,162]. This enhanced CO2 separation can nullify the need for refrigerant
preparation and storage [162,163]. However, CO2 is separated using cryogenic distillation
coupled with biogas upgrading. A comparison of different separation methods for CO2
capture is shown in Table 7.
Atmosphere 2022, 13, 1958 13 of 42

Table 7. Current status of different separation technologies for CO2 capture [16,37,54,111–114,161,164–187].

Chemical Physical Chemical-Looping Membrane


Parameter Adsorption Cryogenic
Absorption Absorption Combustion Separation
Rectisol, Selexol, etc. Mostly Pressure swing adsorption
Amine, chilled ammonia, Polymeric, inorganic and
Separation technique integrated gasification and pressure–temperature FeO, CuO, MnO, and NiO Cryogenic distillation.
and amino acid salt solvent. mixed membranes.
combined cycle. swing adsorption.
Very high CO2
Highly recommended for concentration.
High reactivity, low cost of
separating CO2 during Recycling is possible since it Low-cost oxygen carrier High capture efficiency (up
the solvent, and low No regeneration processes.
pre-combustion processes is a reversible process. materials. to 99.9%).
molecular weight result in a Less solid waste produced.
that operate at elevated CO2 It is possible to achieve high Truly and directly reduces Mature technology.
Pros high mass-based absorption Less chemical consumption.
partial pressures. adsorption efficiency the atmospheric CO2 For many years, CO2 has
capacity, and moderate High efficiency (>95% for
Captures CO2 selectively (485%). concentration. been recovered in the
thermal stability and single metal).
from a gas stream without a Low waste generation. Viable alternative for CO2 industry by this method.
thermal degradation rate.
chemical reaction capture from mobile and
decentralized sources.
High energy requirement
due to refrigeration.
High capital expenditure.
High energy is required to
Requires adsorbent capable Fouling and low fluxes are Need for removal of water,
compress feed gas to a high
of operating at elevated Currently, the process is examples of operational NOx, SOx, and other trace
pressure.
Relatively high temperatures. under development, and issues. components to avoid the
Cons Low CO2 solubility.
maintenance cost. The significant amount of large-scale operations have High running costs. freezing and eventual
Less efficient absorption
energy needed for CO2 not yet been carried out. Removal (%) decreases with blockage of process
process.
desorption is high. the presence of other metals. equipment.
Large equipment sizing.
The procedure consumes a
significant amount of
energy.
CO2 concentration (vol.%) <30.4 >59.3 28–34 3–8 11.8 <90
CO2 capture efficiency (%) 95 >90 <85 52–60 90 99.9
CO2 capture cost
26.2 25.1 6.94 16–26 3–10 32.7
(USD/tonne CO2 )
CO2 purity (%) 99 <99 99.98 >96 95 99.95
SaskPower, Saskatchewan, Schwarze Pumpe power
Summit Power Group, LLC,
Canada (Boundary Dam station, Spremberg,
Seattle, USA (Texas Clean
Carbon Capture Project) Germany (Oxy-fuel
Energy Project)
TransAlta Corporation, technology)
Don valley, Yorkshire, UK
Status of research and Alberta Canada (Project Less large-scale CS Energy: Callide Power Air Products and Chemicals,
(Don Valley Power Project) Under developmental stage.
development Pioneer Keephills 3 Power demonstration plants. Plant A, Queensland, Inc., Pennsylvania USA
Nuon Power, Buggenum,
Plant) Australia (Callide Oxy-fuel
The Netherlands
American Electric Power, Project)
(Integrated gasification
OH, USA (Mountaineer OxyCoal, UK (Oxy-fuel
combined cycle plant)
Power Plant) technology)
Atmosphere 2022, 13, 1958 14 of 42

6. CO2 Capture Using Dry Solid Sorbents


The selective separation of CO2 based on interactions between gases and solids is
required for CO2 capture using a dry adsorbent [188]. In packed columns, universal dry
adsorbents are typically utilized, including activated carbon and molecular sieves [189].
The surface tension and pore size of the adsorbent, as well as the process temperature
and partial pressure, are critical factors in a dry adsorption process [190]. Adsorption and
desorption cycles are repeated throughout the process (regeneration).
The following are the several adsorption types: (i) pressure swing adsorption (PSA) [4–6];
(ii) temperature swing adsorption (TSA) [191,192], which combines two processes—low-
temperature adsorption followed by desorption or regeneration by raising the pressure; (iii)
electric swing adsorption (ESA) [193], which entails adsorption and desorption by altering the
electricity supply while a low-voltage current flows through the adsorbent; and (iv) vacuum
swing adsorption (VSA) [194]. Additionally, due to their low energy needs and relative
simplicity, adsorption-based technologies such as pressure/vacuum swing adsorption (PVSA)
have been extensively researched [195–197].
To effectively treat high volumes of combustion emissions from numerous sources, it
is imperative to increase a dry adsorbent’s CO2 capture selectivity and adsorptive capacity.
The CO2 adsorption can be enhanced and stabilized by adding functional groups to the
surface of the adsorbent material that has a high affinity for CO2 to react with it. CO2
can then be selectively adsorbed using the adsorbent’s sizeable specific surface area and
pore structure [198–203].
The addition of different amine groups to solid materials used as CO2 capture sor-
bents is anticipated to increase polarization and CO2 capture to achieve high selectivity
and capture performance. These adsorbents have several benefits, including the potential
eradication of corrosion issues and reduced costs of energy for regeneration. Under anhy-
drous conditions, the reactions of CO2 with amine functional groups result in ammonium
carbamates, as follows [204]:
CO2 + 2RNH2 → RNHCOO− + RNH3+ (7)

CO2 + 2R2 NH → R2 NCOO− + R2 NH2+ (8)


The reversible nature of the adsorption process and the possibility to increase ad-
sorption efficiency by altering the composition of adsorbent materials make this process
important. Therefore, by choosing a suitable adsorbent material, the CO2 adsorption ef-
ficiency can be improved. Activated carbons [205,206], zeolites [207], hollow fibers, and
alumina are currently the major commercially available adsorbents. Each material has a
unique surface area, pore structure, and surface functional groups, and their application
areas are very specific. The following section discusses and provides descriptions of some
typical carbonaceous and non-carbonaceous dry sorbents for CO2 capture.
Non-carbonaceous adsorbents
Non-carbonaceous adsorbents include zeolites, metalorganic frameworks, silica, etc.
The CO2 uptake capacity of these materials is high; however, they are expensive and highly
sensitive to moisture. As a result, zeolites can capture CO2 molecules of 0.33 nm due to
their suitable channel diameters (0.3–1.0 nm) [208].

6.1. Adsorbents Based on Zeolite


Adsorbents with a natural structure, such as zeolites, are made up of an interlocking
structure of AlO4 and SiO4 tetrahedrons that share atoms of oxygen [209]. Small pore size
and high porosity are responsible for zeolite’s excellent CO2 adsorption at temperatures as
low as 30 ◦ C.; however, their CO2 capture capacity rapidly declines above this temperature
and is almost non-existent above 200 ◦ C [209]. Zeolite’s high hydrophilicity significantly
reduces its CO2 adsorption capacity when the gas contains water. In this case, high
temperatures are required for the regeneration [210], resulting in significant energy losses
as a result of the CO2 regeneration [200].
Atmosphere 2022, 13, 1958 15 of 42

The CO2 selectivity of zeolites remains low despite their ability to separate gases [211–213].
The moisture content of the gas is a constraint on zeolites’ capacity to adsorb CO2 from flue gas.
Water competes for adsorption with CO2 , reducing the amount of CO2 adsorbed by Y-type
zeolites (CaY and NiY) [214]. By contacting an ion and a dipole at low desorption temperatures,
physisorption occurs, resulting in a linear orientation of the CO2 molecule [209,215].
Additionally, zeolites exhibit excellent chemical reactivity, recyclability, high stability,
excellent recyclability, and structural diversity, which make them a promising candidate
for the CO2 adsorption [216]. This structure consists of a network of streams into which
gas molecules are adsorbable, ranging from 0.5 to 1.2 nm in size [200,209,217]. Flue gas
mixtures (CO2 , H2 O, CO) are not suitable for zeolites because they are not CO2 selective,
along with their limited capacity to adsorb at high temperatures [209,218,219].
Zeolite contains exchangeable cations including K+ , Mg2+ , and Na+ that act as a
balance to the negative charge that results from a SiO4 tetrahedron replacing an AlO4
tetrahedron. Specifically, the charge density, cation distribution, and size of the zeolites
influence their adsorption and separation properties. Zeolite particles have a large pore
size that allows CO2 to diffuse into them. The adsorption of CO2 molecules by zeolites is
increased by the presence of cations in their structure via molecules and the adsorbent’s
interaction electrostatically [220]. CO2 adsorption has been studied mostly for the zeolites
13X and 5A. Using a type 13X zeolite, Moura, Bezerra [221] investigated the effect of cation
exchange on CO2 adsorption, concluding that zeolites rich in alkali-cations, such as Li+
and Na+ , could effectively adsorb CO2 . The order of decreasing adsorption capacity was
Li+ > Na+ > NH4+ > Ba2+ > Fe3+ . According to Walton [222], it is believed that the ionic
charge of zeolites, the ionic radius, the shielding effects, as well as the nuclear charge are
the main causes of the observed decline in CO2 adsorption capacity. According to Calleja,
Jimenez [223], zeolites 13X and 5A have CO2 adsorption capacities of 3–25 wt. % and
2–12 wt. %, respectively, at a CO2 partial pressure of 15% and a CO2 pressure of 100%.

6.2. Adsorbents Based on Metal-Organic Frameworks


The molecular organic framework (MOF) is a crystalline material containing metal
ions or clusters of metal ions combined with organic molecules (ligands) [9,43]. As the MOF
has been adapted to a variety of chemical processes and applications, over 20,000 MOFs
have now been created, each with distinct geometries, pore sizes, and functions [224–228].
Having large mesopores, high surface areas (approximately 10,000 m2 /g), and an adapt-
ability to different geometries, MOFs are ideal chemisorbents for CO2 capture [228,229]. Al-
though MOFs have some advantages when it comes to CO2 capture at high temperatures, they
also have some drawbacks. For example, moisture is absorbed while CO2 is being captured,
the manufacturing process is difficult, and the MOFs are not very durable [200].
In a study by Szcz˛eśniak and Choma [230], copper-based MOFs (Cu-BTC), with benzene-
1,3,5-tricarboxylate, had an uptake capacity for CO2 of 9.59 mmol/g adsorbent and a surface
area of 1760 m2 /g at 1 atm at 273 K. The CO2 capacity of most MOFs decreases noticeably
as the temperature increases during the CO2 capture phase. At 298 K, 313 K, and 328 K, the
temperatures for CO2 capture, Aarti et al. [231] reported 4 mmol/g adsorbent for Cu-BTC-
PEI-2.5, 2.61 mmol/g adsorbent for Cu-BTC-PEI-2.5, and 1.66 mmol/g adsorbent for Cu-BTC-
PEI-2.5, respectively, due to MOFs’ structural stability’s degradation with temperature.
There has been a breakthrough in developing microporous coordination polymers
(MCPs) to decrease the production cost of MOFs. As the name implies, MCPs are com-
posed of repeating channels of metal ions linked by ligands whose function is to link
ions. MCPs have porous structures because of their inherited organic functionality [232].
The magnesium-based MOF Mg-MOF-74 was found to have better CO2 adsorption with
8.61 mmol/g adsorbents at 1 bar partial pressure and 298 K [233]. Magnesium dioxy-
benzenedicarboxylate (Mg/DOBDC) is an MCP composed of Mg2+ ions connected to
2,5-dioxide-1,4-benzene-dicarboxylate [232]. Researchers have claimed that the high per-
formance can be attributed to the improved ionic behavior of the magnesium oxide Mg-O
bond when CO2 partial pressures are as low as 0.1 bar. As opposed to MgO, exothermic car-
Atmosphere 2022, 13, 1958 16 of 42

bonation does not form magnesium carbonate (MgCO3 ) from Mg/DOBDC. MCP networks
are inflexible, and prevent insertion into the MgO bonds likely due to their inflexibility.
Due to the highly ionic nature of the Mg-O bonds in Mg/DOBDC, the material can capture
CO2 reversibly. Low CO2 partial pressures can be improved by these techniques [232].
The ability of many early MOF adsorbents to adsorb large amounts of CO2 was
demonstrated. To be used as CO2 adsorbents, MOFs must be stable in an aqueous
medium. Mahdipoor, Halladj [234] reported an MOF CO2 adsorbent containing iron-
based on poly terephthalate (BDC). Observations revealed that the amino-functionalized
MOF MIL 101(FE) is water- and ethanol-stable, and capable of capturing 13 mmol/g CO2
per adsorbent. Additionally to the physical CO2 adsorption by the amino MIL 101(FE)
MOF’s structure, the chemisorption of CO2 by the MIL 101(FE) MOF structure affects the
total CO2 adsorption by the MOF [234].

6.3. Mesoporous Silica Materials


Mesoporous silica has also been proposed as a viable candidate material for capturing
CO2 because of its large surface area and high capacity to be synthesized with a variety
of pore hole sizes [235]. The modification of mesoporous silica for CO2 capture has been
proven to be an effective method for developing adsorbents, despite its low adsorption
capacity [235]. The possibility of chemically treating their surface OH groups to promote
their CO2 selectivity and adsorption capacity may be further investigated to process flue
gases at low pressures using CO2 [235].
Sánchez-Zambrano, Lima Duarte [236] investigated the chemical modifications of
mesoporous silica treated with 3-aminopropyl triethoxysilane (APTES), and then impreg-
nated it with polyethyleneimine (PEI) to find out how the modifications affected kinetic
mechanisms, site energy distributions, and CO2 adsorption under post-combustion condi-
tions [10,43]. As determined by microcalorimetry, the functionalization process resulted
in new adsorption sites. When amine groups were added to the support, physisorption
accounted for the vast majority of adsorption, and CO2 capture capacity and selectivity in-
creased as the temperature increased (50 and 75 ◦ C). Due to sites that could not be restored
due to strong chemical bonds formed by adsorption products, using a turbomolecular
vacuum pump was the only way to regenerate after adsorption at 25 ◦ C with CO2 . In the
first stage of CO2 capture, these sites became available for regeneration, but only at higher
temperatures [236]. Although mesoporous silicas are highly effective adsorbents for CO2
capture due to their large pores with a tunable size, high surface area, and mechanical and
thermal stability, they do not possess sufficient CO2 adsorption abilities to be of any use,
especially at a 1-atmosphere pressure [235].

6.4. Alkali Metal-Based Materials


Alkali metal carbonates, including those made of Na, K, and Al, have been reported
to be efficient dry adsorbents for CO2 capture from flue gas operating below 473 K in
relatively moderate conditions [237,238]. Various inorganic supports, including alumina,
silica, ceramics, zirconia, and carbon materials, are added with alkali metal carbonates
during this procedure; Equation (9) explains how moisture and CO2 react, facilitating CO2
adsorption, and how decarbonization (Equation (10)) regenerates the absorbent (10) [38]:

M2 CO3 + H2 O + CO2 ↔ 2MHCO3 (M = K, Na) (9)

141 kJ/mol and ∆H = –135 kJ/mol for M = K and Na, respectively

2MHCO2 ↔ M2 CO3 + H2 O + CO2 (10)

As shown in Equation (8), alkali–metal bicarbonates are typically formed when CO2
and H2 O react with carbonate sorbents between 333 and 383 K, which then regenerate
alkali–metal carbonates at 373–473 K, releasing CO2 . Theoretically, Na2 CO3 and K2 CO3
have CO2 adsorption capacities of 41.5 and 31.8 wt. %, respectively.
Atmosphere 2022, 13, 1958 17 of 42

The use of lithium-based materials such as lithium-based silicate (Li4 SiO4 ) and lithium-
based zirconate (Li2 ZrO3 ) for direct CO2 capture from flue gas at temperatures of 700–900 K
is another promising method for capturing CO2 [239]. Li4 SiO4 , specifically, is a promising
CO2 captor because of its low volume change during CO2 -adsorption–desorption and its
high CO2 sorption capacity of 36.7% [240].
In a study conducted by Kato et al., Li4 SiO4 and Li2 ZrO3 were examined at low
CO2 concentrations, i.e., 50 ppv. They found that Li4SiO4 was 30 times more capable of
absorbing CO2 than Li2 ZrO3 . Additionally, zirconia materials are more expensive than
silica materials [241]. Recently, Seggiani et al. [242] reported that Li4 SiO4 , with the addition
of 30% Na2 CO3 or K2 CO3 , demonstrated a CO2 sorption capacity of 23 wt. % at an ideal
sorption temperature of 853 K and low CO2 partial pressure of 0.04 bar, equating to a
Li4SiO4 conversion of almost 80%. Li8 SiO6 was proposed by Durán-Muoz et al., as a
substitute dry adsorbent for CO2 capture. It demonstrated a high sorption capacity of
roughly 51.9 wt. % over a wide temperature range, with an efficiency of 71.1% [243].
It is technically and economically desirable to use alkali–metal-based materials to
capture CO2 post-combustion at low temperatures and in low concentrations, since they
do not require additional cooling processes; although, the long-term stability and sustained
performance of such adsorbents under real flue gas conditions must be addressed.

6.5. Alkaline Metal-Based Ceramics


In addition to alkaline ceramics, binary metal oxides are also referred to as alkaline
oxides because they contain more than one alkaline element. Metal-based alkaline ceramics
include Li5 AlO4 , Li2 CuO2 , Li2 ZrO3 , and Li4 SiO4 . Notably, these alkaline metal ceramics can
be regenerated at elevated temperatures. Consequently, they are ideal for adsorption and
desorption cycles, such as those associated with post-combustion capture. The robustness
of alkaline metal ceramics allows them to be made without refractory supports, resulting
in a simpler synthesis method [244]. Alkaline ceramics are synthesized in a variety of ways,
depending on the desired type. When synthesizing zirconates, co-precipitation, sol-gel, or
soft chemistry methods are preferred for achieving the desired morphology of an adsorbent.
A sorbent needs to have a morphology with a large surface area and a small particle size
to capture CO2 to its maximum capacity [244]. The sol-gel method can be used to create a
mesoporous structure for silicates that is advantageous [244].
These materials, including zirconate Na2 ZrO3 , are important for process intensifi-
cation [245], owing to their catalytic activity [244]. Following CO2 chemisorption, the
carbonated derivate, Na2 CO3 , is formed. Equation (11) depicts the reaction pathway for
Na2 ZrO3 [244,245].

Na2 ZrO3(S) + CO2(g) → Na2 CO3(S) + ZrO2(g) ∆H298K = −149 KJ/mol (11)

Sutton, Kelleher [246] described alkaline metal-based ceramics as catalysts of the water
gas shift reaction (WGSR).

CO(g) + H2 O(g) → CO2(g) + H2 O(g) ∆H298K = −42.2KJ/mol (12)

Compared to other ceramics made of alkaline metal, lithium cuprate (Li2 CuO2 ) has
demonstrated the ability to adsorb CO2 over a wide temperature range (120 to 690 ◦ C), and
it retains its efficacy as a CO2 adsorbent at low CO2 partial pressures [247,248]. At CO2
concentrations less than 15%, the alkaline metal zirconate (Li2 ZrO3 ) exhibits excellent CO2
selectivity and CO2 acceptor performance in flue gas streams [244]. In summary, Li2 CuO2
and Li2 ZrO3 are two CO2 adsorbents that have been subjected to extensive research and
have been demonstrated to be promising dependable CO2 adsorbents [244].
Much of the research on alkaline ceramics in recent years has focused on CO2 capture
at high temperatures. The CO2 adsorption capacities of two alkaline yttrium oxides (NaYO2
and LiYO2 ) at high temperatures were recently evaluated. Since lithium ions are not as
close together as sodium ions are in an octahedron, LiYO2 has a greater capacity for CO2
Atmosphere 2022, 13, 1958 18 of 42

capture than NaYO2 . However, both ceramics were capable of adsorption/desorption


for at least ten cycles [249]. Table 8 lists the basic CO2 adsorption capacities of the main
adsorbent types that are commonly addressed in the literature.

Table 8. The CO2 adsorption capacity of the most common adsorbents [250–252].

Adsorbent Examples Temp. (K) Press. (atm) SBET(m2 g−1 ) Capacity (mmol g−1 )
Zeolite-based NaY, NaX, 13X, ZIF-70, ZIF-69 273–384 1 15–1730 ≤5.4
Silica monolith/TEPA,
Amine-based 298–384 1 16–367 ≤5.9
MCM-41/PEI
MOF-based MOF-53, MOF-177 198–304 1–96 270–4500 ≤48.7
Calcium-based Ca(OH)2 , CaO 195–348 1 - ≤11.6
Alkali ceramic-based Na2 ZrO3 , Li2 ZrO3 500–600 1 - ≤6.5
Activated. Carbon BPL,
Carbon-based 195–348 1 1150–3250 ≤8
MAXSORP, Activated carbon

Carbonaceous Adsorbents
Even though carbon is the only component of carbonaceous materials, they offer
several advantages, such as high thermal/chemical stability, heat and electrical conductivity,
bio-affinities, elasticities, and strengths [253–255]. Being lightweight, having a very high
specific surface area, and having a large pore volume make them especially suitable for
applications involving the storage of gases or adsorption [256,257]. Additionally, they offer
benefits for CO2 capture: (i) they are not moisture-sensitive; (ii) they are reasonably priced;
(iii) their desorption/adsorption temperatures are below 373 K.; (iv) they may be employed
at atmospheric pressure; and (v) their energy usage is minimal. Each of these factors has
had an impact on recent research in this field.

6.6. Activated Carbons


The textural characteristics and surface groups of carbon-based adsorbents are highly
correlated with CO2 adsorption capability [258,259]. Due to their materials’ wide range of
pore sizes, from micropores to macropores, activated carbons are not suitable for selective
gas adsorption. Adsorption temperatures of less than 25 kJ/mol are typically seen in
pristine carbon-based adsorbents, which have weak CO2 affinities [260]. At 298 K and
0.1 pressure, the standard CO2 adsorption capacity of activated carbon is ~5 wt. % [261,262].
By modifying the activation and preparation conditions, the pore structures of acti-
vated carbons can be easily regulated [263,264]. Additionally, the activated carbon’s surface
functional groups can be easily modified utilizing a variety of treatment methods [265,266].
Incorporating different basic groups into activated carbon has been extensively researched
for improving the CO2 affinity by enhancing the CO2 adsorption capacity [267,268].
NiO-loaded activated carbons (NiO-ACs) were synthesized by Jang et al., utilizing a
post-oxidation technique that involved nickel electroless plating at 573 K in an air stream.
The NiO-AC samples’ ability to adsorb CO2 increased as oxidation duration increased. The
maximal CO2 adsorption capacity was 49.9 cm3 /g, above the 41.2 cm3 /g capacity of unal-
tered activated carbon at 298 K and 1 bar. They reported that the acid-base characteristics
of the NiO caused it to serve as an electron donor on the carbon surface, increasing the CO2
adsorption, which acts as an electron acceptor [269].

6.7. Graphene
A graphene derivative known as graphene oxide (GO) can be created using different
functional groups on the edges and basal planes [270]. Researchers have extensively studied
modifying the surfaces of GO with functional groups for applications including gas storage,
separation, conversion of energy, and synthesizing newly developed GO-like derivatives
with lightweight frameworks [271–273].
Thermally exfoliated graphene nanoplates were reported by Meng et al. as being
innovative, highly effective sorbents for CO2 capture. At 298 K and 30 bar, the produced
graphene nanoplates demonstrated remarkable capture efficiencies of 248 wt. %. The
Atmosphere 2022, 13, 1958 19 of 42

graphene nanoplates’ wider inter-layer spacing and substantial inner void volume were
attributed to the higher CO2 capture capacity [273].

6.8. Ordered Porous Carbons


Since they are widely used as electrode materials, catalyst supports, and other types
of materials, ordered porous carbon materials have drawn a lot of study interest [274–276].
There are numerous ways to make ordered porous carbons, including (i) direct synthesis
utilizing organic self-assembly, which uses a mixture of carbon precursors and blocks
co-polymers as soft templates, and (ii) nano-casting, employing silica materials as structure-
controlling hard templates [277,278].
Yoo et al., examined the impact of the phenolic resins’ carbonization temperature
on the total pore volumes and the specific surface areas of ordered nanoporous carbons
(ONCs). Due to its greater specific surface area and smaller micropore size distribution,
ONC carbonized at 1173 K had the highest CO2 adsorption capacities at 298 K (15.8 wt. %
at 1 pressure and 68.5 wt. % at 30 bar) [279].

6.9. Activated Carbon Fibers (ACFs)


ACFs are attractive adsorbent materials because of their numerous micrometer porosi-
ties, nanostructures, and other characteristics, including narrow pore size distributions and
large specific surface areas [280,281]. Compared to granular and powdered adsorbents,
ACFs are more flexible due to their fibrous structure [282].
For large-scale CO2 capture testing, Thiruvenkatachari et al. created huge honeycomb-
shaped carbon fiber composite (HMCFC) adsorbents (the adsorbent mass in one column
was 4.486 kg). For a simulated flue gas with 13% CO2 , 5.5% O2 , and the remaining
N2 at 293 K, the average CO2 adsorption capacity was 11.9 wt. %. Additionally, they
demonstrated that the thermal decomposition process and combined vacuum improve the
CO2 collection efficiency of the HMCFC adsorbents [283].

7. Research Progress in Converting CO2 into Valuable Fuels


One of the most stable compounds with carbon in its greatest valence state is CO2 .
Due to its low electron affinity, electrophilic reactions are challenging. Thus, a nucleophilic
assault on the carbon atom is necessary for the conversion of CO2 . As is well known, the
dissociation energy required to rupture the C=O bond within a CO2 molecule is greater
than 750 kJ mol−1 [284]. From a thermodynamic perspective, this is an uphill reaction. To
provide the necessary energy for such a reaction to be completed, high temperatures, high
pressures, or extremely effective catalysts are frequently required.
To date, a range of methods have been employed to reduce CO2 , including the thermal
catalysis [285–289], photocatalysis [290–293], electrocatalysis [294–297], and photoelectro-
chemical (PEC) reactions [298–301], in which heat, light, or electricity are utilized to supply
the reaction’s necessary energy. Eight electrons are required for each CO2 molecule to
convert into a hydrocarbon molecule completely. Due to the large number of compounds
that are created during the reduction process, the purification procedure is complicated,
and the yield of the desired products is decreased. Table 9 presents a clear comparison
of the performance of different CO2 conversion methods employed in recent years. Ex-
tremely concentrated and efficient enzymatic processes were added to the aforementioned
reduction technologies to boost the efficiency and accuracy of the CO2 conversion [302,303].
These reactions were inspired by natural photosynthesis. The schematic of the most modern
CO2 chemical conversion techniques is shown in Figure 7.
Atmosphere 2022, 13, 1958 20 of 42

Figure 7. A schematic of the CO2 chemical conversion methods [304].

Table 9. Performance evaluation of various CO2 reduction systems.

Method of CO2
Materials Efficiency Reference
Conversion
CH3 OH, 95%, 85 mM at −0.2 V vs. standard
CuO + Cu2 O Photo electrocatalysis [305]
hydrogen electrode (SHE) after 1.5 h
C60 polymer film Photocatalysis HCOOH, 239.46 µM after 2 h [306]
Ni-Ru/Al2 O3 Thermal catalysis CO2 conversion: 82.7% CO2 selectivity: 100% [307]
Polyoxometalates (POMs) Electrocatalysis 38.9 mA cm−2 [308]
Co-Pi/Fe2 O3 Photoelectro/enzymatic catalysis HCOOH, 6.4 µM h−1 [309]
Plain graphite rod Photo/enzymatic catalysis HCOOH, 15.49 µM mg Enzyme−1 min−1 [310]

As a result of the use of different CO2 conversion strategies, such as thermal catalysis,
photocatalysis, electrocatalysis, photoenzymatic catalysis, photoelectroenzymatic catalysis,
and photoelectrochemical (PEC), a range of yields and products were obtained. Table 10
the numerous value-added products made using different CO2 utilization strategies and
the catalysts used throughout the conversion process.

1
Atmosphere 2022, 13, 1958 21 of 42

Table 10. List of catalysts utilized in CO2 conversion processes to manufacture valuable products.

Catalysts Products Methods Reference


CuInS2 thin film Methanol Photoelectrochemical [311]
Sulfur modified copper Formate Electrochemical [312,313]
Indium Formate and acetate Electrochemical [314]
W18 O49 Photocatalytic [315]
p-type GaP Methanol Photoelectrochemical [316]
Zn2 GeO4 nanoribbon Methane Photocatalytic [317]
Bi2 WO6 nanoplate Methane Photocatalytic [318]
HNb3 O8 nanobelt Methane Photocatalytic [319]
TiO2 Methanol and methane Photocatalytic [320]
PET supported TiO2 Carbon monoxide Photocatalytic [321]
Nickel (Ni) Methane Thermal (Reforming) [322]

8. Technologies for CO2 Capture at Various Technological Readiness Levels (TRL)


The global concern over the effects of industrial operations, such as chemical processes,
on the environment has increased interest in green technologies. This issue has fueled
significant growth in research and new technologies across many fields since the turn of
the 21st century. Due to their significant contributions to high greenhouse gas emissions,
the power generation, agricultural, and chemical industries stand out numerically in
developing green technologies. It is crucial to track how new technologies are being
developed and how old ones are being improved. This allows for future projections for a
given area of Research, Development and Innovation (RDI).
A tool called technological maturity analysis, created by NASA in 1990 and modi-
fied by The Electric Power Research Institute (EPRI), can rank existing technologies by
highlighting the stage they are at. This system is called the Technology Readiness Level
(TRL) system [323]. TRL has a straightforward nomenclature (TRL1, TRL2, TRL3, etc.) and
enables analysis of the stages of technological development. As illustrated in Table 11,
these stages advance progressively, and each one is necessary for the next.

Table 11. Stages of TRL development as defined by EPRI. Source: adapted from Freeman and Bhown
(2011) [323].

TRL Level Technology Mature Level


6 Integrating pilot testing into an appropriate environment
7 Full-functioning prototype, miniature demonstration
Demonstration
8 Commercial implementation and full-scale deployment
9 Standard trade services
3 Component-level proof-of-concept evaluation
Development 4 A laboratory setting for system validation
5 Validation of a subsystem in an environment
1 Observation of fundamental ideas, initial conception
Research
2 Application-based formulation

TRL analysis assesses current technologies, aids in decision-making, and provides for
forecasts of time, cost, and environmental implications. Politicians and businesspeople use
it as a tool today across the globe, notably in RDI.
Carbon capture, storage and use (CCSU) technologies are typically expensive technologies,
hence it is crucial to conduct a technological maturity analysis to identify potential solutions
that can direct RDI investments by analyzing the TRL levels the technology can achieve [324].
The International Energy Agency (2020) [325] conducted a survey of opinions on CO2
capture systems based on TRL levels. Table 12 presents information from this review’s four cate-
gories of technologies—adsorption, absorption, membrane separation, and chemical capture.
Atmosphere 2022, 13, 1958 22 of 42

Table 12. TRL assessments for adsorption, absorption, membrane separation, and chemical capture
technologies (2020) conducted by the International Energy Agency (2020) [325].

Technology Category TRL Level Considerations


It is one of the most advanced technologies. As a result of the research time,
Absorption 9 this technology has been applied to small and large power plants, fuel
converters, and industrial production facilities.
The technology is used in natural gas and ethanol processing, where CO2 can
Adsorption 9 be captured in large plants. There are many possible applications for this
technology. One of its main advantages is its simplicity of operation.
Among existing separation technologies, it is considered the most effective and
is relatively new. In terms of advances, they depend on existing separation
Membrane separation 6–7
technologies. A small number of its commercial applications have been
developed, while most are in demonstration or development phases.
Due to the time and research intensity involved, the capture involving
Chemical capture 4–6 chemical reactions is provided in the TRL. Its magnitude is explained by the
requirement for extensive pilot-scale testing given that it is still relatively new.

TRL studies have been carried out using information from technology suppliers. Data
from Kearns et al. (2021) [326] were used as a comparative source.
The main drawback of the TRL scale, according to Freeman and Bhown (2011) [323],
is the absence of requirements at each stage of development to move on to the next stage.
Specifically, the amount of effort needed to move from TRL1 to TRL2 and beyond is unknown.

9. The Use of CO2 as a Feedstock for Fuel and Chemical Production


Finding alternatives to fossil fuels is viewed as being of great priority globally due to
the increasing reliance on them and the depletion of resources. In general, it has become
crucial to find a sustainable solution for transforming CO2 , a toxic greenhouse gas that con-
tributes to global warming, into a renewable carbon supply. CO2 can be directly converted
into several useful chemicals through endergonic or exergonic processes [327]. CO2 can
be used to meet the demands of different industries, including those for beverages, food,
and chemicals [328]. Because of the financial and environmental benefits, technologies that
allow CO2 to be converted into value-added products are continuously being researched.
The valence state of CO2 can be altered, in contrast to physical processes [329]. This
procedure can be utilized to create chemical feedstock (carbonates, plastics, polymers [330])
as well as energy forms (syngas, methanol, ethane, methane). There are three types of
chemical conversions that can be distinguished: thermochemical, electrochemical (photo-
electrochemical [331]), and biological processes that involve enzymes [332].
There is a thermodynamic hurdle in CO2 conversion due to its high stability [333].
Hydrogen is a key element in many processes that convert CO2 into value-added chemicals.
It should be produced from renewable energy sources to have an environmentally benign
effect. Syngas (intermediate products) is produced as a byproduct of the reforming process
that converts waste materials into useful fuels and chemicals. It often contains significant
amounts of hydrogen and carbon monoxide, as well as small amounts of water and carbon
dioxide [334]. Through the pyrolysis or gasification of biomass or natural gas conversion,
respectively, reforming can occur in a solid state and with or without a gaseous state into
syngas. Significant amounts of CO2 released from various industrial facilities, such as
fossil fuel-fired power stations, can be used as feedstock in diverse CO2 recycling processes.
The primary obstacle limiting the development of large-scale applications for biofuel is
the availability of source feedstocks (namely, CO2 and H2 ). Methane (CH4 ), methanol
(CH3 OH), and dimethyl ether (CH3 OCH3 ) are just a few of the many biofuels that can
be made from CO2 . This path makes it possible to create a wide range of fuels for both
stationary and mobile applications.
Atmosphere 2022, 13, 1958 23 of 42

9.1. Production of Chemicals


A wide variety of fine compounds can be made from CO2 , in addition to synthetic fuels.
The most significant uses are urea alkylene carbonates (a few kt year−1 ), polycarbonates
(4 Mt year−1 ), acrylic acid and acrylates (10 Mt year−1 ), polyurethane (≈18 Mt year−1 ),
inorganic carbonates (≈60 Mt year−1 ), and urea (≈160 Mt year−1 ) [335]. The greatest
market for the use of carbon dioxide is urea, a key fertilizer [335,336]. It is also frequently
utilized as a feedstock in producing fine chemicals, polymers, medicines, and inorganic
compounds such as urea resins and melamine [337,338].
Other chemicals that can be generated from CO2 capture are also beneficial in various
sectors, including lubricants, polymers, agrochemicals, pharmaceuticals, coatings, and
catalytic processes. These chemical classes include organic carbonates such as diphenyl
carbonate (DPC), dimethyl carbonate (DMC), diallyl carbonate (DAC) and diethyl carbonate
(DEC), as well as cyclic carbonates such as styrene carbonate (SC), cyclohexene carbonate
(CC), propylene carbonate (PC) and ethylene carbonate (EC), and even polycarbonates
such as bisphenol polycarbonate (BPA-PC) and poly(propylene carbonate) [339,340].
This process faces difficulties since it requires a large amount of catalyst inventory and
operates at high temperatures and pressures. Additionally, this procedure faces additional
challenges in separating the catalyst from the products [339,340]. Commercially accessible
Al-based catalysts are frequently utilized, but they are not eco-friendly in manufacturing
polycarbonates from the reaction of CO2 with epoxides. In this context, an alternate
method with a lot of potential for producing polycarbonates from CO2 and olefins is the
oxidative carboxylation approach [341]. Another chemical made from the interaction of
CO2 and cyclic amines, such as azetidines and aziridines, or the N-analogues of epoxides,
is polyurethane [341,342].
Formic acid is another significant chemical that can be created using CO2 . The hydro-
genation of CO2 into formic acid has lately attracted some study interest because of the benign
reaction conditions, the absence of byproduct formation, its ability to store hydrogen in liquid
form, and the ease of formic acid’s decomposition into hydrogen and CO2 [337,343].
The biological use of CO2 provides an additional route for producing biodiesel and
numerous commodity chemicals produced from biomass (used in food, silage, biogas,
and fertilizer) [344]. This method has the benefits of a faster development rate, a shorter
growth cycle, no competition with other plants for land, and the creation of various
valuable by-products. To remove contaminants such as heavy metals, NOx, and SOx that
are harmful to microalgae growth, the captured CO2 should be filtered before being fed
into a photobioreactor [345].
Using CO2 as a technological fluid without converting it into chemicals has found
applications outside of EOR in a variety of industries, such as the air conditioning (as a
coolant), food-preservation, dry-washing, solvent, and beverage industries [335,343,346].
Generally, EOR consumes 50 Mt year−1 CO2 , compared to the 8 Mt year−1 CO2 used by the
beverage and food industries [347].
Overall, the proposed laboratory-scale solutions are still far from being commercialized
for industrial use, even though there is a huge market for converting captured CO2 into
chemicals and fuels. This is partly due to the exorbitant manufacturing costs of the materials
under investigation thus far, which are also not chemically stable, and in part to the
generally low CO2 conversion rates and total yields of the primary products. As a result,
they do not satisfy the criteria for widespread deployment. Furthermore, knowledge of the
mechanisms underlying the chemical reactions involved in the transformations of CO2 is
still in its infancy. Evaluations of the requirements and factors of the procedure have also
been neglected in this discipline.

9.2. Production of Fuels


The most effective method of utilizing CO2 is its conversion to fuels. Alkanes, methane,
methanol, syngas, and other compounds can be produced from the captured CO2 . In addition
to transportation, power plants, and fuel cells, the fuel produced can be used in many different
Atmosphere 2022, 13, 1958 24 of 42

industries [335]. The number of methods for using CO2 to produce fuels is enormous. Since
CO2 is a molecule with a steady thermodynamic state, its use requires a lot of heat and
catalysts to achieve high fuel yields [348]. The hydrogenation and dry reforming of methane
(DRM) are the two main processes for producing fuel from captured CO2 [349].
The possibility of creating fuel, storing H2 , recycling CO2 , and resolving the difficulty
of storing electric energy makes CO2 hydrogenation a very attractive method of using
CO2 [343]. DRM is also regarded as one of the most significant routes for the Fischer–
Tropsch (FT) process’ production of methanol and numerous other liquid fuels [341,350,351].
The source of hydrogen from fossil fuels seems to be a concern in the hydrogenation of
CO2 into methane [352], methanol [353], carbon monoxide [352], and formic acid [354], as
this can result in a rise in the atmospheric emissions of CO2 .
However, renewable energy sources, such as solar, wind, and biomass, can be used as
a substitute for fossil fuels to further reduce CO2 emissions during hydrogenation [355]. Re-
cently, the “e-gas” created by the German Audi Motor Company using CO2 hydrogenation
produced 1000 Mt year−1 of methane [356].
The volumetric gas density of methane is low, which makes it an unsuitable fuel for use
in vehicles. Additionally, its global warming potential (GWP) is 30 [348]. Methane is readily
available, thus producing more will not be profitable for CO2 capture (methane is plentiful
in landfill gas, coal gas, shale gas, and natural gas). It appears that a preferable mechanism
is CO2 hydrogenation to methanol [357]. However, it is extremely difficult to activate C–H
bonds with the currently used (mostly Cu-based) catalysts for the manufacture of methanol,
and the catalysts that have been tried so far are not particularly profitable [358–360].
Methanol has numerous uses in organic solvents, combustion engines, plastics, and
paints [348]; nevertheless, only 0.1% of CO2 emissions are reduced by its synthesis [361].
One of the most significant methods for using CO2 is the reverse water–gas shift (RWGS)
reaction, which converts CO2 into CO because CO is a starting material for the FT reaction,
which produces methanol and hydrocarbon fuels [359]. Despite this, the RWGS reaction’s
endothermic nature and the low conversion at moderate temperatures provide the two
biggest challenges to the implementation of large-scale methanol production from CO2
using the FT process. Another major obstacle is the development of active catalysts that
can speed up the reaction rate and enhance the yield.
DRM has recently gained a lot of study interest in terms of utilizing CO2 to produce
syngas [340,362,363]. According to Ofélia de Queiroz et al. [364], DRM often produces
syngas with a greater purity than partial oxidation and steam reforming. Additionally, the
DRM process only produces 2% of unreacted methane, which is less than steam reforming,
allowing it to be used at remote natural gas locations to produce liquid fuels, which are
more transport-friendly than gaseous fuels [341]. The DRM reaction has been thoroughly
examined for Rh, Ir, Ru, Ni–Co, and Ni supported on lanthanum oxide, alumina, and
silica [341]. Despite major advancements in the design of catalysts with high activity and
ideal stability for DRM, finding a good catalyst for this reaction still poses a significant
difficulty, especially at high operation temperatures when deactivation by coke formation
is inevitable (>700 ◦ C) [365–369].
Another appealing method that can lessen the formation of coke and maintain the
stability of catalysts at high temperatures is the oxidative dehydrogenation of light alkanes
to alkenes (ODA), which uses CO2 as a soft oxidant instead of the usual dehydrogenation
oxidant, O2 [370–374]. Additionally, by eliminating hydrogen via the RWGS process,
CO2 enhances the light alkanes’ oxidative dehydrogenation equilibrium conversion [375].
However, it is important to monitor the temperature since too much heat might lead to
the olefins overoxidizing, which produces carbon oxides and reduces selectivity [376]. The
redox cycle and active oxygen species are both formed by CO2 . The active site type, the
reduction capability of the metal, and the supporting material all affect how CO2 functions
in the ODA and how this reaction works [377]. The catalysts studied thus far have little
stability, while having a high initial activity.
Atmosphere 2022, 13, 1958 25 of 42

The aforementioned discussion makes it clear that creating novel catalysts with high
catalytic activity under a variety of reaction conditions, coke resistance, and long-term
structural and chemical stability is the main obstacle to using captured CO2 as a feedstock
for the creation of synthetic fuels.

9.2.1. Production of Methane (CH4 ) Based on CO2 (Methanation), Challenges and Prospects
Chemical feedstocks such as CO2 can be used to transform renewable energy. One
promising approach to combating CO2 -induced climate change is CO2 methanation for
value-added products, where water electrolysis is a potential energy storage technique for
generating H2 with renewable energy and would contribute to the creation of a carbon-
based cycle that is sustainable. Due to their inherent intermittency, renewable energy
sources are currently limited by the need for scalable storage methods [378]. Therefore, the
most practical and easiest method of storing significant volumes of intermittent energy
generated from renewable sources for extended periods is the generation of synthetic
natural gas or liquid fuels. Power to gas (PtG) is one concept that has gained a lot of
attention over the years (Figure 8) [379]. In this method, CO2 interacts with H2 , which is
created by water electrolysis using renewable solar or wind energy to generate CH4 as a
substitute for natural gas. In Copenhagen, a commercial-scale PtG plant with 1.0 M2 of
capacity was successfully operating in 2016, exploiting the change of the energy system
towards a sustainable system [380]. Five projects, using CO2 methanation at a commercial
or pilot plant size with capacities varying from 25 kW to 6300 kW, were implemented in
Germany between 2009 and 2013 [381]. In fact, natural gas, or methane, is Germany’s
main source of heat and a significant contributor to natural gas supplies. Due to their
robust dynamic properties, natural gas power plants now generate a larger portion of
Germany’s electricity than the country’s present coal-fired power plants do [382]. Because
it has a higher H:C ratio than its conventional counterpart, its use in automobiles instead of
gasoline minimizes CO2 emissions. The following reactions in Table 13 take place in the
methanation reactor [383]:

Figure 8. Schematic illustration of “power to gas” (PtG) technology (adapted from Younas [380]).
Atmosphere 2022, 13, 1958 26 of 42

Table 13. CO2 methanation: main reactions and side reactions [384].

Reaction Equation Response Type ∆H298K


R /kJ.mol−1
CH4
C + 2H2 Methane pyrolysis −75
CO2 + 2H2
C + 2H2 O CO2 reduction −90
CO + H2
C + H2 O CO reduction −131
2CO
C + CO2 Boudouard reaction −172
CO2 + 4H2
CH4 + 2H2 O CO2 methanation −165
2CO + 2H2
CH4 + CO2 reverse dry reforming −247
CO + 3H2
CH4 + H2 O CO methanation −206
CO2 + H2
CO + H2 O reverse water gas shift 41

It is difficult to apply CO2 because of its inertness, which prevents it from being
converted into chemicals with added value. The use of specific catalysts can, however,
help to resolve this problem [385]. According to Park, Kwak [386], utilizing a double-
layered TiO2 /Cu-TiO2 catalyst instead of a typical TiO2 catalyst resulted in a two-fold
improvement in the yield of CH4 production from CO2 (film catalyst). To clean syngas in
ammonia factories, carbon oxides were also hydrogenated to methane. Additionally, this
might result in carbon-neutral fuel (methane) [327]. The conversion of CO2 to methane can
also occur biologically, for example, using methanogens. Methane-producing organisms
are produced via the anoxic enrichment of waste-activated sludge (methanogens). The
effectiveness of methane generation was improved by almost 70-fold when the organism’s
activated cultures were used [387].
Due to the increasing demand for storing renewable energy and mitigating global
warming, CO2 methanation has gained renewed attention due to the French chemist Paul
Sabatier’s discovery in 1902 [388]. Storing energy from renewable resources such as wind
and solar, the efficient conversion of biogas to biomethane, and the conversion of CO2 into
chemical feedstocks and fuels are all made possible by the Sabatier reaction [389,390].
Figure 9 illustrates the exothermic nature of CO2 methanation, which has a high
equilibrium conversion between 25 and 400 ◦ C [391,392]. By using the right catalysts, CO2
methanation can achieve 99% CH4 selectivity, avoid product separation, and circumvent the
challenges of dispersed product distribution. Because of this thermodynamic characteristic,
CO2 methanation is more important in terms of energy effectiveness and economic viability.

Figure 9. A plot of equilibrium CO2 conversion in methanation at various temperatures (based on


literature data) [393,394].

9.2.2. Challenges
In CO2 methanation, deactivating metal catalysts is a major challenge. There are two
different ways that methanation catalysts can be deactivated: (a) chemically and (b) physically.
In contrast to Co/Al2 O3 , which deactivated quickly in the same amount of time, the
Co/ZrO2 catalysts showed a greater CO2 methanation activity and practically consistent
Atmosphere 2022, 13, 1958 27 of 42

performance even after 300 h on stream. Through thermogravimetric analysis and hy-
drothermal (H2 O) treatment verification experiments, the deactivation of the Co/Al2 O3
catalyst was further researched. An excessive amount of CoAl2 O4 was produced because
of the addition of extra H2 O to the reaction system, which hastened the deactivation of the
Co/Al2 O3 catalyst. As a result, the product H2 O encourages the production of the inactive
phase CoAl2 O4 , which causes Co/Al2 O3 catalysts to deactivate quickly. One of the causes
of deactivation is the deposition of carbon, but the primary cause is the creation of the
CoAl2 O4 spinel structure in the inactive phase.
Active metal sintering and carbon deposition are the main causes of physical deactiva-
tion. Ni/YSZ catalysts were synthesized by Kesavana et al. [395] using several techniques.
Thin coatings of carbon are generated on Ni0 particles with a spherical shape, and graphitic
laminations are formed on Ni0 particles that are exposed to surfaces on the Ni/YSZ catalyst
created by the impregnation procedure. When Ni0/Ni2+ oxidation with a high CO2 :H2
ratio was used, the Ni/YSZ(EDTA) catalyst displayed outstanding stability, and operand
XAS demonstrated that it was not deactivated. Increasing the H2 /CO2 ratio or adding
steam can prevent carbon from depositing on the catalyst because hydrogen reacts with the
carbon deposits to keep them from deactivating the catalyst.
Increased metal dispersion via strong metal–support contact, the addition of catalyst
promoters, the development of novel synthesis techniques, and other typical techniques
are used to reduce metal sintering [396]. When the active Co particles are separated by
graphite-like carbon, metal sintering is averted effectively. In a study by Li et al. [397],
NiO–MgO@ SiO2 catalysts were successfully prepared and maintained their activity after
100 h on stream. Therefore, particles with the right size are beneficial to CH4 formation.

9.2.3. Prospects
Increasing atmospheric CO2 concentrations are believed to be one of the primary
drivers of climate change. Although CO2 capture technologies are quite advanced, us-
ing captured CO2 remains a significant difficulty that will require extensive future re-
search. In the short to medium term, the creation of new life cycle assessments (LCA),
techno-economic tools, and benchmark assessments will make it possible to evaluate CO2
conversion pathways consistently and transparently. Current barriers to furthering CO2
conversion include energy and water usage, the need for costly catalysts, and problems
with gas infrastructure. Cost-effective CO2 conversion techniques are especially needed,
and their pilot-scale testing is essential for their commercial application. A thorough intro-
duction to the various CO2 conversion processes aids in comprehension of the mechanisms
and the selection of suitable methods for using captured CO2 . It is preferable to use the
prospective integrated approaches to close the current gaps for widespread industrial
applications. Additionally, it is advantageous to demonstrate research and development
(R&D) initiatives for CO2 conversion and analyze CO2 utilization markets to determine the
likelihood of the full-scale implementation of these technologies.
Economic viability, whether it is attained through technical progress or legislative
changes, is crucial to promoting the development of the industrial feasibility of CO2 con-
version technology. Because of this, future key research objectives should be concentrated
on CO2 conversion and CCS legislation, policy, and assessments, as well as the integration
of CO2 use with other strategies to minimize costs and energy consumption, especially at a
larger scale. Public education and publicity on the capture, sequestration and methanation
of CO2 (CSCM) should be highlighted, and global collaborations need to be further strength-
ened in the meantime, to increase the general public’s awareness of the environmental
repercussions. Additionally, it should be highlighted that CCS is necessary to achieve our
climate goals, and that CO2 is not a substitute for it. To keep the rise in the global mean
temperature below 1.5 ◦ C, governments should strengthen their commitment to CSCM and
play a critical role in supporting its deployment (e.g., through laws, financial options, and
tax incentives). The encouragement of private sector funding for larger-scale demonstration
programs and the commercialization of CO2 technologies is another successful approach.
Atmosphere 2022, 13, 1958 28 of 42

The market for CO2 -based products may soon be impacted by regulations and policies as
CO2 becomes a resource that diverse sectors of the global economy want.
The following are general directions for future CO2 methanation research:
1. To increase the H/C ratio of catalyst surface and facilitate C–C coupling for high-
value-added products;
2. To promote CO2 adsorption and activation by enhancing the oxygen vacancies and
support basicity;
3. To investigate potential novel catalytic materials and enhance the stability of the catalyst;
4. To develop more effective catalysts for CO2 hydrogenation at low temperatures and
with little energy consumption;
5. To analyze the process intensification and optimization of CO2 conversion technolo-
gies, which are crucial to understanding how various operating parameters interact,
improve process effectiveness, and reduce costs.

10. Conclusions
The above sections provide a comprehensive overview of different CO2 separation
and capture technologies. The ideal separation techniques to lower CO2 emissions are also
highlighted, along with their benefits and drawbacks, in this review paper. Additionally,
dry solid sorbents, with a focus on emerging adsorbents, were addressed. A discussion of
the challenges involved in converting CO2 into value-added products was also conducted,
and future perspectives were laid out.
This study leads to the conclusion that the CO2 capture and separation (CCS) problem
cannot be solved by a single method. Multiple technologies covered in this paper would
need to be incorporated to solve this complex problem. To solve the technological problems
in the CCS, it is necessary to scale-up innovative technologies from a laboratory to an
industrial scale. This requires further research on new physical and chemical sorbents and
procedures with improved cost-effectiveness and efficiency.

Author Contributions: Conceptualization, S.O.A.; investigation, S.OA.; methodology, S.O.A.; writing—


original draft preparation, S.OA.; writing—review and editing, S.O.A. and Y.M.I.; supervision, Y.M.I.
All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable for studies not involving humans or animals.
Informed Consent Statement: Not applicable.
Acknowledgments: The authors used data and information from various research papers and
reports as indicated in the references, and gratefully acknowledge the assistance of the facilitators,
namely, Joseph Gbadeyah, Elutunji Buraimoh, and Emmanuel Tetteh, at the Scientific research retreat
organized by the Faculty of Engineering and the Built Environment, Durban University of Technology,
South Africa.
Conflicts of Interest: The authors declare no known competing financial interest or personal rela-
tionships that could influence the information given in this work.

References
1. Al-Mamoori, A.; Lawson, S.; Rownaghi, A.A.; Rezaei, F. Improving adsorptive performance of CaO for high-temperature CO2
capture through Fe and Ga doping. Energy Fuels 2019, 33, 1404–1413. [CrossRef]
2. Rojas-Cuevas, I.-D.; Caballero-Morales, S.-O.; Sánchez-Partida, D.; Martínez-Flores, J.-L. A Proposal to the Reduction of Carbon
Dioxide Emission in Inventory Replenishment: Mitigating the Climate Change. In Humanitarian Logistics from the Disaster Risk
Reduction Perspective; Springer: Cham, Switzerland, 2022; pp. 189–203.
3. Edrisi, S.A.; Tripathi, V.; Dubey, P.K.; Abhilash, P. Carbon sequestration and harnessing biomaterials from terrestrial plantations
for mitigating climate change impacts. In Biomass, Biofuels, Biochemicals; Elsevier: Amsterdam, The Netherlands, 2022; pp. 299–313.
4. Dai, Y.; Lai, F.; Ni, J.; Liang, Y.; Shi, H.; Shi, G. Evaluation of the impact of CO2 geological storage on tight oil reservoir properties.
J. Pet. Sci. Eng. 2022, 212, 110307. [CrossRef]
Atmosphere 2022, 13, 1958 29 of 42

5. Triyanti, A.; Marfai, M.A.; Mei, E.T.W.; Rafliana, I. Review of Socio-Economic Development Pathway Scenarios for Climate
Change Adaptation in Indonesia: Disaster Risk Reduction Perspective. In Climate Change Research, Policy and Actions in Indonesia;
Springer: Cham, Switzerland, 2021; pp. 13–31.
6. Gimžauskaitė, D.; Aikas, M.; Tamošiūnas, A. Recent progress in thermal plasma gasification of liquid and solid wastes. Recent
Adv. Renew. Energy Technol. 2022, 2, 155–196. [CrossRef]
7. Ouyang, T.; Xie, S.; Pan, M.; Qin, P. Peak-shaving scheme for coal-fired power plant integrating flexible carbon capture and
wastewater treatment. Energy Convers. Manag. 2022, 256, 115377. [CrossRef]
8. Holz, F.; Scherwath, T.; del Granado, P.C.; Skar, C.; Olmos, L.; Ploussard, Q.; Ramos, A.; Herbst, A. A 2050 perspective on the role
for carbon capture and storage in the European power system and industry sector. Energy Econ. 2021, 104, 105631. [CrossRef]
9. Zhao, Y.; Yan, Y.; Lee, J.-M. Recent progress on transition metal diselenides from formation and modification to applications.
Nanoscale 2022, 14, 1075–1095. [CrossRef]
10. Li, K.; Zhang, D.; Niu, X.; Guo, H.; Yu, Y.; Tang, Z.; Lin, Z.; Fu, M. Insights into CO2 adsorption on KOH-activated biochars
derived from the mixed sewage sludge and pine sawdust. Sci. Total Environ. 2022, 826, 154133. [CrossRef]
11. Zhang, D.; Zou, R.; Wang, X.; Liu, J.; Mo, F. Carbon Capture, Utilization & Storage: A General Overview. In Climate Mitigation and
Adaptation in China; Springer: Singapore, 2022; pp. 61–107.
12. Alobaid, F.; Ströhle, J. Thermochemical Conversion Processes for Solid Fuels and Renewable Energies; MDPI: Basel, Switzerland, 2021;
Volume 11, p. 1907.
13. Alrebei, O.F.; Amhamed, A.I.; El-Naas, M.H.; Hayajnh, M.; Orabi, Y.A.; Fawaz, W.; Al-tawaha, A.S.; Medina, A.V. State of the
Art in Separation Processes for Alternative Working Fluids in Clean and Efficient Power Generation. Separations 2022, 9, 14.
[CrossRef]
14. Nawkarkar, P.; Ganesan, A.; Kumar, S. Carbon dioxide capture for biofuel production. In Handbook of Biofuels; Elsevier: Amsterdam,
The Netherlands, 2022; pp. 605–619.
15. Xu, C.-G.; Li, X.-S. Research progress of hydrate-based CO2 separation and capture from gas mixtures. Rsc. Adv. 2014, 4,
18301–18316. [CrossRef]
16. Theo, W.L.; Lim, J.S.; Hashim, H.; Mustaffa, A.A.; Ho, W.S. Review of pre-combustion capture and ionic liquid in carbon capture
and storage. Appl. Energy 2016, 183, 1633–1663. [CrossRef]
17. Babu, P.; Kumar, R.; Linga, P. A new porous material to enhance the kinetics of clathrate process: Application to precombustion
carbon dioxide capture. Environ. Sci. Technol. 2013, 47, 13191–13198. [CrossRef]
18. Pires, J.; Martins, F.; Alvim-Ferraz, M.; Simões, M. Recent developments on carbon capture and storage: An overview. Chem. Eng.
Res. Des. 2011, 89, 1446–1460. [CrossRef]
19. Wall, T.; Liu, Y.; Spero, C.; Elliott, L.; Khare, S.; Rathnam, R.; Zeenathal, F.; Moghtaderi, B.; Buhre, B.; Sheng, C. An overview
on oxyfuel coal combustion—State of the art research and technology development. Chem. Eng. Res. Des. 2009, 87, 1003–1016.
[CrossRef]
20. Toftegaard, M.B.; Brix, J.; Jensen, P.A.; Glarborg, P.; Jensen, A.D. Oxy-fuel combustion of solid fuels. Prog. Energy Combust. Sci.
2010, 36, 581–625. [CrossRef]
21. Chen, L.; Yong, S.Z.; Ghoniem, A.F. Oxy-fuel combustion of pulverized coal: Characterization, fundamentals, stabilization and
CFD modeling. Prog. Energy Combust. Sci. 2012, 38, 156–214. [CrossRef]
22. Buhre, B.J.; Elliott, L.K.; Sheng, C.; Gupta, R.P.; Wall, T.F. Oxy-fuel combustion technology for coal-fired power generation. Prog.
Energy Combust. Sci. 2005, 31, 283–307. [CrossRef]
23. Stanger, R.; Wall, T. Sulphur impacts during pulverised coal combustion in oxy-fuel technology for carbon capture and storage.
Prog. Energy Combust. Sci. 2011, 37, 69–88. [CrossRef]
24. Chen, S.S.; Curcic, M. Ocean surface waves in Hurricane Ike (2008) and Superstorm Sandy (2012): Coupled model predictions
and observations. Ocean Model. 2016, 103, 161–176. [CrossRef]
25. Kunze, C.; Spliethoff, H. Assessment of oxy-fuel, pre-and post-combustion-based carbon capture for future IGCC plants. Appl.
Energy 2012, 94, 109–116. [CrossRef]
26. Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J.M.; Balbuena, P.B. CO2 capture and separations using MOFs: Computational and
experimental studies. Chem. Rev. 2017, 117, 9674–9754. [CrossRef]
27. Ben-Mansour, R.; Habib, M.; Bamidele, O.; Basha, M.; Qasem, N.; Peedikakkal, A.; Laoui, T.; Ali, M. Carbon capture by physical
adsorption: Materials, experimental investigations and numerical modeling and simulations—A review. Appl. Energy 2016, 161, 225.
[CrossRef]
28. Bougie, F.; Iliuta, M.C. CO2 absorption in aqueous piperazine solutions: Experimental study and modeling. J. Chem. Eng. Data
2011, 56, 1547–1554. [CrossRef]
29. Kim, H.S.; Gwak, I.C.; Lee, S.H. Numerical analysis of heat transfer area effect on cooling performance in regenerator of free-piston
Stirling cooler. Case Stud. Therm. Eng. 2022, 32, 101875. [CrossRef]
30. Kazemi, A.; Mehrabani-Zeinabad, A. Post combustion carbon capture: Does optimization of the processing system based on
energy and utility requirements warrant the lowest possible costs? Energy 2016, 112, 353–363. [CrossRef]
31. Li, K.; Leigh, W.; Feron, P.; Yu, H.; Tade, M. Systematic study of aqueous monoethanolamine (MEA)-based CO2 capture process:
Techno-economic assessment of the MEA process and its improvements. Appl. Energy 2016, 165, 648–659. [CrossRef]
Atmosphere 2022, 13, 1958 30 of 42

32. Li, K.; Cousins, A.; Yu, H.; Feron, P.; Tade, M.; Luo, W.; Chen, J. Systematic study of aqueous monoethanolamine-based CO2
capture process: Model development and process improvement. Energy Sci. Eng. 2016, 4, 23–39. [CrossRef]
33. Gibbins, J.; Chalmers, H. Carbon capture and storage. Energy Policy 2008, 36, 4317–4322. [CrossRef]
34. Bailey, K.; Kurz, K.; Sushkova, T. Carbon Offset Solutions for International Travel Emissions; University of Richmond: Richmond, VA,
USA, 2014.
35. Gibbins, J.; Chalmers, H. Preparing for global rollout: A ‘developed country first’demonstration programme for rapid CCS
deployment. Energy Policy 2008, 36, 501–507. [CrossRef]
36. Ketzer, J.M.; Iglesias, R.S.; Einloft, S. Reducing greenhouse gas emissions with CO2 capture and geological storage. In Handbook of
Climate Change Mitigation and Adaptation; Springer: New York, NY, USA, 2015; Volume 2, pp. 1–40.
37. Leung, D.Y.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies.
Renew. Sustain. Energy Rev. 2014, 39, 426–443. [CrossRef]
38. Samanta, A.; Zhao, A.; Shimizu, G.K.; Sarkar, P.; Gupta, R. Post-combustion CO2 capture using solid sorbents: A review. Ind. Eng.
Chem. Res. 2012, 51, 1438–1463. [CrossRef]
39. Lim, Y.; Kim, J.; Jung, J.; Lee, C.S.; Han, C. Modeling and simulation of CO2 capture process for coal-based power plant using
amine solvent in South Korea. Energy Procedia 2013, 37, 1855–1862. [CrossRef]
40. Madejski, P.; Chmiel, K.; Subramanian, N.; Kuś, T. Methods and techniques for CO2 capture: Review of potential solutions and
applications in modern energy technologies. Energies 2022, 15, 887. [CrossRef]
41. Liu, Y.; Fan, W.; Wang, K.; Wang, J. Studies of CO2 absorption/regeneration performances of novel aqueous monothanlamine
(MEA)-based solutions. J. Clean. Prod. 2016, 112, 4012–4021. [CrossRef]
42. Liu, Z.; Balasubramanian, R. Upgrading of waste biomass by hydrothermal carbonization (HTC) and low temperature pyrolysis
(LTP): A comparative evaluation. Appl. Energy 2014, 114, 857–864. [CrossRef]
43. Figueroa, J.D.; Fout, T.; Plasynski, S.; McIlvried, H.; Srivastava, R.D. Advances in CO2 capture technology—The US Department
of Energy’s Carbon Sequestration Program. Int. J. Greenh. Gas Control 2008, 2, 9–20. [CrossRef]
44. Khurana, N.; Goswami, N.; Sarmah, R. Carbon Capture: Innovation for a Green Environment. In Advances in Carbon Capture and
Utilization; Springer: Berlin/Heidelberg, Germany, 2021; pp. 11–31.
45. Kashyap, N.; Deka, B.; Choudhari, B.; Singh, B.P.; Pachani, S. Present status of capture of carbon dioxide and its storage
technologies: A review. Pharma Innov. J. 2019, 8, 327–332.
46. Heldebrant, D.J.; Koech, P.K.; Glezakou, V.-A.; Rousseau, R.; Malhotra, D.; Cantu, D.C. Water-lean solvents for post-combustion
CO2 capture: Fundamentals, uncertainties, opportunities, and outlook. Chem. Rev. 2017, 117, 9594–9624. [CrossRef]
47. Bishnoi, S.; Rochelle, G.T. Absorption of carbon dioxide into aqueous piperazine: Reaction kinetics, mass transfer and solubility.
Chem. Eng. Sci. 2000, 55, 5531–5543. [CrossRef]
48. Austgen, D.M.; Rochelle, G.T.; Chen, C.C. Model of vapor-liquid equilibria for aqueous acid gas-alkanolamine systems. 2.
Representation of H2S and CO2 solubility in aqueous MDEA and CO2 solubility in aqueous mixtures of MDEA with MEA or
DEA. Ind. Eng. Chem. Res. 1991, 30, 543–555. [CrossRef]
49. Barzagli, F.; Mani, F.; Peruzzini, M. Continuous cycles of CO2 absorption and amine regeneration with aqueous alkanolamines: A
comparison of the efficiency between pure and blended DEA, MDEA and AMP solutions by 13C NMR spectroscopy. Energy
Environ. Sci. 2010, 3, 772–779. [CrossRef]
50. Bishnoi, S.; Rochelle, G.T. Physical and chemical solubility of carbon dioxide in aqueous methyldiethanolamine. Fluid Phase
Equilibria 2000, 168, 241–258. [CrossRef]
51. Xu, G.-W.; Zhang, C.-F.; Qin, S.-J.; Gao, W.-H.; Liu, H.-B. Gas− liquid equilibrium in a CO2 − MDEA− H2O system and the effect
of piperazine on it. Ind. Eng. Chem. Res. 1998, 37, 1473–1477. [CrossRef]
52. Aroonwilas, A.; Veawab, A. Integration of CO2 capture unit using single-and blended-amines into supercritical coal-fired power
plants: Implications for emission and energy management. Int. J. Greenh. Gas Control 2007, 1, 143–150. [CrossRef]
53. Veawab, A.; Aroonwilas, A.; Tontiwachwuthikul, P. CO2 absorption performance of aqueous alkanolamines in packed columns.
Fuel Chem. Div. Prepr. 2002, 47, 49–50.
54. Aaron, D.; Tsouris, C. Separation of CO2 from flue gas: A review. Sep. Sci. Technol. 2005, 40, 321–348. [CrossRef]
55. Knudsen, J.N.; Andersen, J.; Jensen, J.N.; Biede, O. Evaluation of process upgrades and novel solvents for the post combustion
CO2 capture process in pilot-scale. Energy Procedia 2011, 4, 1558–1565. [CrossRef]
56. Gurkan, B.E.; de la Fuente, J.C.; Mindrup, E.M.; Ficke, L.E.; Goodrich, B.F.; Price, E.A.; Schneider, W.F.; Brennecke, J.F. Equimolar
CO2 absorption by anion-functionalized ionic liquids. J. Am. Chem. Soc. 2010, 132, 2116–2117. [CrossRef]
57. Rochelle, G.T. Thermal degradation of amines for CO2 capture. Curr. Opin. Chem. Eng. 2012, 1, 183–190. [CrossRef]
58. Fredriksen, S.; Jens, K.-J. Oxidative degradation of aqueous amine solutions of MEA, AMP, MDEA, Pz: A review. Energy Procedia
2013, 37, 1770–1777. [CrossRef]
59. Kozak, F.; Petig, A.; Morris, E.; Rhudy, R.; Thimsen, D. Chilled ammonia process for CO2 capture. Energy Procedia 2009,
1, 1419–1426. [CrossRef]
60. Asif, M.; Suleman, M.; Haq, I.; Jamal, S.A. Post-combustion CO2 capture with chemical absorption and hybrid system: Current
status and challenges. Greenh. Gases Sci. Technol. 2018, 8, 998–1031. [CrossRef]
61. Tan, J.; Shao, H.; Xu, J.; Du, L.; Luo, G. Mixture absorption system of monoethanolamine-triethylene glycol for CO2 capture. Ind.
Eng. Chem. Res. 2011, 50, 3966–3976. [CrossRef]
Atmosphere 2022, 13, 1958 31 of 42

62. Rivas, O.; Prausnitz, J. Sweetening of sour natural gases by mixed-solvent absorption: Solubilities of ethane, carbon dioxide, and
hydrogen sulfide in mixtures of physical and chemical solvents. AIChE J. 1979, 25, 975–984. [CrossRef]
63. Barzagli, F.; Di Vaira, M.; Mani, F.; Peruzzini, M. Improved Solvent Formulations for Efficient CO2 Absorption and Low-
Temperature Desorption. ChemSusChem 2012, 5, 1724–1731. [CrossRef]
64. Liu, A.H.; Li, J.J.; Ren, B.H.; Lu, X.B. Development of High-Capacity and Water-Lean CO2 Absorbents by a Concise Molecular
Design Strategy through Viscosity Control. ChemSusChem 2019, 12, 5164–5171. [CrossRef]
65. Li, X.; Liu, J.; Jiang, W.; Gao, G.; Wu, F.; Luo, C.; Zhang, L. Low energy-consuming CO2 capture by phase change absorbents of
amine/alcohol/H2O. Sep. Purif. Technol. 2021, 275, 119181. [CrossRef]
66. Shen, L.; Liu, F.; Shen, Y.; Sun, C.; Zhang, Y.; Wang, Q.; Li, S.; Li, W. Novel biphasic solvent of AEP/1-propanol/H2 O for CO2
capture with efficient regeneration performance and low energy consumption. Sep. Purif. Technol. 2021, 270, 118700. [CrossRef]
67. Karlsson, H.K.; Makhool, H.; Karlsson, M.; Svensson, H. Chemical absorption of carbon dioxide in non-aqueous systems using
the amine 2-amino-2-methyl-1-propanol in dimethyl sulfoxide and N-methyl-2-pyrrolidone. Sep. Purif. Technol. 2021, 256, 117789.
[CrossRef]
68. Alkhatib, I.I.; Khalifa, O.; Bahamon, D.; Abu-Zahra, M.R.; Vega, L.F. Sustainability criteria as a game changer in the search for
hybrid solvents for CO2 and H2S removal. Sep. Purif. Technol. 2021, 277, 119516. [CrossRef]
69. Mathias, P.M.; Afshar, K.; Zheng, F.; Bearden, M.D.; Freeman, C.J.; Andrea, T.; Koech, P.K.; Kutnyakov, I.; Zwoster, A.; Smith,
A.R. Improving the regeneration of CO2 -binding organic liquids with a polarity change. Energy Environ. Sci. 2013, 6, 2233–2242.
[CrossRef]
70. Barzagli, F.; Giorgi, C.; Mani, F.; Peruzzini, M. Screening study of different amine-based solutions as sorbents for direct CO2
capture from air. ACS Sustain. Chem. Eng. 2020, 8, 14013–14021. [CrossRef]
71. Wanderley, R.R.; Pinto, D.D.; Knuutila, H.K. From hybrid solvents to water-lean solvents–A critical and historical review. Sep.
Purif. Technol. 2021, 260, 118193. [CrossRef]
72. Liu, F.; Shen, Y.; Shen, L.; Zhang, Y.; Chen, W.; Wang, Q.; Li, S.; Zhang, S.; Li, W. Sustainable ionic liquid organic solution with
efficient recyclability and low regeneration energy consumption for CO2 capture. Sep. Purif. Technol. 2021, 275, 119123. [CrossRef]
73. Wai, S.K.; Nwaoha, C.; Saiwan, C.; Idem, R.; Supap, T. Absorption heat, solubility, absorption and desorption rates, cyclic capacity, heat
duty, and absorption kinetic modeling of AMP–DETA blend for post–combustion CO2 capture. Sep. Purif. Technol. 2018, 194, 89–95.
[CrossRef]
74. Muchan, P.; Narku-Tetteh, J.; Saiwan, C.; Idem, R.; Supap, T. Effect of number of amine groups in aqueous polyamine solution on
carbon dioxide (CO2 ) capture activities. Sep. Purif. Technol. 2017, 184, 128–134. [CrossRef]
75. Singto, S.; Supap, T.; Idem, R.; Tontiwachwuthikul, P.; Tantayanon, S.; Al-Marri, M.J.; Benamor, A. Synthesis of new amines
for enhanced carbon dioxide (CO2 ) capture performance: The effect of chemical structure on equilibrium solubility, cyclic
capacity, kinetics of absorption and regeneration, and heats of absorption and regeneration. Sep. Purif. Technol. 2016, 167, 97–107.
[CrossRef]
76. Zhou, T.; McBride, K.; Linke, S.; Song, Z.; Sundmacher, K. Computer-aided solvent selection and design for efficient chemical
processes. Curr. Opin. Chem. Eng. 2020, 27, 35–44. [CrossRef]
77. Malhotra, D.; Koech, P.K.; Heldebrant, D.J.; Cantu, D.C.; Zheng, F.; Glezakou, V.A.; Rousseau, R. Reinventing design principles
for developing low-viscosity carbon dioxide-binding organic liquids for flue gas clean up. ChemSusChem 2017, 10, 636–642.
[CrossRef]
78. Chapman, W.G.; Gubbins, K.E.; Jackson, G.; Radosz, M. SAFT: Equation-of-state solution model for associating fluids. Fluid Phase
Equilibria 1989, 52, 31–38. [CrossRef]
79. Chapman, W.G.; Gubbins, K.E.; Jackson, G.; Radosz, M. New reference equation of state for associating liquids. Ind. Eng. Chem.
Res. 1990, 29, 1709–1721. [CrossRef]
80. Papadokonstantakis, S.; Badr, S.; Hungerbühler, K.; Papadopoulos, A.I.; Damartzis, T.; Seferlis, P.; Forte, E.; Chremos, A.; Galindo,
A.; Jackson, G. Toward sustainable solvent-based postcombustion CO2 capture: From molecules to conceptual flowsheet design.
In Computer Aided Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2015; Volume 36, pp. 279–310.
81. Limleamthong, P.; Gonzalez-Miquel, M.; Papadokonstantakis, S.; Papadopoulos, A.I.; Seferlis, P.; Guillén-Gosálbez, G. Multi-
criteria screening of chemicals considering thermodynamic and life cycle assessment metrics via data envelopment analysis:
Application to CO2 capture. Green Chem. 2016, 18, 6468–6481. [CrossRef]
82. Papadopoulos, A.I.; Badr, S.; Chremos, A.; Forte, E.; Zarogiannis, T.; Seferlis, P.; Papadokonstantakis, S.; Adjiman, C.S.; Galindo,
A.; Jackson, G. Efficient screening and selection of post-combustion CO2 capture solvents. Chem. Eng. 2014, 39, 211–216.
83. Papadopoulos, A.I.; Badr, S.; Chremos, A.; Forte, E.; Zarogiannis, T.; Seferlis, P.; Papadokonstantakis, S.; Galindo, A.; Jackson, G.;
Adjiman, C.S. Computer-aided molecular design and selection of CO2 capture solvents based on thermodynamics, reactivity and
sustainability. Mol. Syst. Des. Eng. 2016, 1, 313–334. [CrossRef]
84. Mota-Martinez, M.T.; Hallett, J.P.; Mac Dowell, N. Solvent selection and design for CO2 capture–how we might have been missing
the point. Sustain. Energy Fuels 2017, 1, 2078–2090. [CrossRef]
85. Anastas, P.; Warner, J. Green Chemistry: Theory and Practice; Oxford University Press: Oxford, UK, 1998.
86. Tobiszewski, M.; Tsakovski, S.; Simeonov, V.; Namieśnik, J.; Pena-Pereira, F. A solvent selection guide based on chemometrics and
multicriteria decision analysis. Green Chem. 2015, 17, 4773–4785. [CrossRef]
87. Jessop, P.G. Searching for green solvents. Green Chem. 2011, 13, 1391–1398. [CrossRef]
Atmosphere 2022, 13, 1958 32 of 42

88. Yeh, A.C.; Bai, H. Comparison of ammonia and monoethanolamine solvents to reduce CO2 greenhouse gas emissions. Sci. Total
Environ. 1999, 228, 121–133. [CrossRef]
89. Diao, Y.-F.; Zheng, X.-Y.; He, B.-S.; Chen, C.-H.; Xu, X.-C. Experimental study on capturing CO2 greenhouse gas by ammonia
scrubbing. Energy Convers. Manag. 2004, 45, 2283–2296. [CrossRef]
90. Sreenivasulu, B.; Gayatri, D.; Sreedhar, I.; Raghavan, K. A journey into the process and engineering aspects of carbon capture
technologies. Renew. Sustain. Energy Rev. 2015, 41, 1324–1350. [CrossRef]
91. Russo, M.; Olivieri, G.; Marzocchella, A.; Salatino, P.; Caramuscio, P.; Cavaleiro, C. Post-combustion carbon capture mediated by
carbonic anhydrase. Sep. Purif. Technol. 2013, 107, 331–339. [CrossRef]
92. Kumar, N.; Rao, D.P. Design of a packed column for absorption of carbon dioxide in hot K2 CO3 solution promoted by arsenious
acid. Gas Sep. Purif. 1989, 3, 152–155. [CrossRef]
93. Derks, P.; Kleingeld, T.; van Aken, C.; Hogendoorn, J.; Versteeg, G. Kinetics of absorption of carbon dioxide in aqueous piperazine
solutions. Chem. Eng. Sci. 2006, 61, 6837–6854. [CrossRef]
94. Guo, H.; Zhou, Z.; Jing, G. Kinetics of carbon dioxide absorption into aqueous [Hmim][Gly] solution. Int. J. Greenh. Gas Control
2013, 16, 197–205. [CrossRef]
95. Wappel, D.; Gronald, G.; Kalb, R.; Draxler, J. Ionic liquids for post-combustion CO2 absorption. Int. J. Greenh. Gas Control 2010,
4, 486–494. [CrossRef]
96. Rogers, R.D. Reflections on ionic liquids. Nature 2007, 447, 917–918. [CrossRef]
97. Davis, J.H., Jr. Task-Specific Ionic Liquids for Separations of Petrochemical Relevance: Reactive Capture of CO2 Using Amine Incorporating
Ions; ACS Publications: Washnigton, DC, USA, 2005.
98. Rahimpour, M.; Kashkooli, A. Enhanced carbon dioxide removal by promoted hot potassium carbonate in a split-flow absorber.
Chem. Eng. Process. Process Intensif. 2004, 43, 857–865. [CrossRef]
99. Lin, P.-H.; Wong, D.S.H. Carbon dioxide capture and regeneration with amine/alcohol/water blends. Int. J. Greenh. Gas Control
2014, 26, 69–75. [CrossRef]
100. Hoffman, J.S.; Hammache, S.; Gray, M.L.; Fauth, D.J.; Pennline, H.W. Parametric study for an immobilized amine sorbent in a
regenerative carbon dioxide capture process. Fuel Process. Technol. 2014, 126, 173–187. [CrossRef]
101. Xu, Z.; Wang, S.; Chen, C. CO2 absorption by biphasic solvents: Mixtures of 1, 4-Butanediamine and 2-(Diethylamino)-ethanol.
Int. J. Greenh. Gas Control 2013, 16, 107–115. [CrossRef]
102. Freeman, S.A.; Dugas, R.; van Wagener, D.H.; Nguyen, T.; Rochelle, G.T. Carbon dioxide capture with concentrated, aqueous
piperazine. Int. J. Greenh. Gas Control 2010, 4, 119–124. [CrossRef]
103. Cheng, H.-H.; Tan, C.-S. Reduction of CO2 concentration in a zinc/air battery by absorption in a rotating packed bed. J. Power
Sources 2006, 162, 1431–1436. [CrossRef]
104. Xu, G.; Zhang, C.; Qin, S.; Wang, Y. Kinetics study on absorption of carbon dioxide into solutions of activated methyldiethanolamine.
Ind. Eng. Chem. Res. 1992, 31, 921–927. [CrossRef]
105. Notz, R.; Asprion, N.; Clausen, I.; Hasse, H. Selection and pilot plant tests of new absorbents for post-combustion carbon dioxide
capture. Chem. Eng. Res. Des. 2007, 85, 510–515. [CrossRef]
106. Chen, X.; Closmann, F.; Rochelle, G.T. Accurate screening of amines by the wetted wall column. Energy Procedia 2011, 4, 101–108.
[CrossRef]
107. Rochelle, G.; Chen, E.; Freeman, S.; van Wagener, D.; Xu, Q.; Voice, A. Aqueous piperazine as the new standard for CO2 capture
technology. Chem. Eng. J. 2011, 171, 725–733. [CrossRef]
108. Adeosun, A.; Abbas, Z.; Abu-Zahra, M.R. Screening and characterization of advanced amine based solvent systems for CO2
post-combustion capture. Energy Procedia 2013, 37, 300–305. [CrossRef]
109. Dash, S.K.; Samanta, A.N.; Bandyopadhyay, S.S. Simulation and parametric study of post combustion CO2 capture process using
(AMP+ PZ) blended solvent. Int. J. Greenh. Gas Control 2014, 21, 130–139. [CrossRef]
110. Spigarelli, B.P. A Novel Approach to Carbon Dioxide Capture and Storage. Ph.D. Thesis, Michigan Technological University,
Houghton, MI, USA, 2013.
111. Takamura, Y.; Aoki, J.; Uchida, S.; Narita, S. Application of high-pressure swing adsorption process for improvement of CO2
recovery system from flue gas. Can. J. Chem. Eng. 2001, 79, 812–816. [CrossRef]
112. Wyczalek, F. Energy Independence-A Nation Running on Empty. In Proceedings of the 3rd International Energy Conversion
Engineering Conference, San Francisco, CA, USA, 15–18 August 2022; p. 5545.
113. Erlach, B.; Schmidt, M.; Tsatsaronis, G. Comparison of carbon capture IGCC with pre-combustion decarbonisation and with
chemical-looping combustion. Energy 2011, 36, 3804–3815. [CrossRef]
114. Clausse, M.; Merel, J.; Meunier, F. Numerical parametric study on CO2 capture by indirect thermal swing adsorption. Int. J.
Greenh. Gas Control 2011, 5, 1206–1213. [CrossRef]
115. Kulkarni, A.R.; Sholl, D.S. Analysis of equilibrium-based TSA processes for direct capture of CO2 from air. Ind. Eng. Chem. Res.
2012, 51, 8631–8645. [CrossRef]
116. Wang, L.; Yang, Y.; Shen, W.; Kong, X.; Li, P.; Yu, J.; Rodrigues, A.E. Experimental evaluation of adsorption technology for CO2
capture from flue gas in an existing coal-fired power plant. Chem. Eng. Sci. 2013, 101, 615–619. [CrossRef]
117. Xu, D.; Xiao, P.; Zhang, J.; Li, G.; Xiao, G.; Webley, P.A.; Zhai, Y. Effects of water vapour on CO2 capture with vacuum swing
adsorption using activated carbon. Chem. Eng. J. 2013, 230, 64–72. [CrossRef]
Atmosphere 2022, 13, 1958 33 of 42

118. Yang, H.; Yuan, Y.; Tsang, S.C.E. Nitrogen-enriched carbonaceous materials with hierarchical micro-mesopore structures for
efficient CO2 capture. Chem. Eng. J. 2012, 185, 374–379. [CrossRef]
119. Arenillas, A.; Smith, K.; Drage, T.; Snape, C. CO2 capture using some fly ash-derived carbon materials. Fuel 2005, 84, 2204–2210.
[CrossRef]
120. Kim, K.; Son, Y.; Lee, W.B.; Lee, K.S. Moving bed adsorption process with internal heat integration for carbon dioxide capture. Int.
J. Greenh. Gas Control 2013, 17, 13–24. [CrossRef]
121. Kumar, R. Adsorption column blowdown: Adiabatic equilibrium model for bulk binary gas mixtures. Ind. Eng. Chem. Res. 1989,
28, 1677–1683. [CrossRef]
122. An, H.; Feng, B.; Su, S. Effect of monolithic structure on CO2 adsorption performance of activated carbon fiber–phenolic resin
composite: A simulation study. Fuel 2013, 103, 80–86. [CrossRef]
123. Ishida, M.; Yamamoto, M.; Ohba, T. Experimental results of chemical-looping combustion with NiO/NiAl2O4 particle circulation
at 1200 C. Energy Convers. Manag. 2002, 43, 1469–1478. [CrossRef]
124. Cho, P.; Mattisson, T.; Lyngfelt, A. Reactivity of iron oxide with methane in a laboratory fluidized bed: Application of chemical-
looping combustion. In Proceedings of the 7th International Conference on Circulating Fluidized Bed Technology, Niagara Falls,
ON, Canada, 5–8 May 2002.
125. Guo, Q.; Zhang, J.; Tian, H. Recent advances in CaSO4 oxygen carrier for chemical-looping combustion (CLC) process. Chem. Eng.
Commun. 2012, 199, 1463–1491. [CrossRef]
126. Adánez, J.; de Diego, L.F.; García-Labiano, F.; Gayán, P.; Abad, A.; Palacios, J. Selection of oxygen carriers for chemical-looping
combustion. Energy Fuels 2004, 18, 371–377. [CrossRef]
127. Zafar, Q.; Mattisson, T.; Gevert, B. Integrated hydrogen and power production with CO2 capture using chemical-looping reforming
redox reactivity of particles of CuO, Mn2O3, NiO, and Fe2O3 using SiO2 as a support. Ind. Eng. Chem. Res. 2005, 44, 3485–3496.
[CrossRef]
128. Lyngfelt, A.; Leckner, B.; Mattisson, T. A fluidized-bed combustion process with inherent CO2 separation; application of
chemical-looping combustion. Chem. Eng. Sci. 2001, 56, 3101–3113. [CrossRef]
129. Kale, G. Feasibility study of sulfates as oxygen carriers for chemical looping processes. QScience Connect 2012, 2012, 1. [CrossRef]
130. Zheng, M.; Shen, L.; Xiao, J. Reduction of CaSO4 oxygen carrier with coal in chemical-looping combustion: Effects of temperature
and gasification intermediate. Int. J. Greenh. Gas Control 2010, 4, 716–728. [CrossRef]
131. Spallina, V.; Gallucci, F.; Romano, M.C.; Chiesa, P.; Lozza, G.; van Sint Annaland, M. Investigation of heat management for CLC
of syngas in packed bed reactors. Chem. Eng. J. 2013, 225, 174–191. [CrossRef]
132. Hamers, H.; Gallucci, F.; Cobden, P.; Kimball, E.; van Sint Annaland, M. A novel reactor configuration for packed bed chemical-
looping combustion of syngas. Int. J. Greenh. Gas Control 2013, 16, 1–12. [CrossRef]
133. Harper, R.N.; Boyce, C.M.; Scott, S.A. Oxygen carrier dispersion in inert packed beds to improve performance in chemical looping
combustion. Chem. Eng. J. 2013, 234, 464–474. [CrossRef]
134. Cotton, A.; Patchigolla, K.; Oakey, J. Hydrodynamic characteristics of a pilot-scale cold model of a CO2 capture fluidised bed
reactor. Powder Technol. 2013, 235, 1060–1069. [CrossRef]
135. Yazdanpanah, M.M.; Forret, A.; Gauthier, T.; Delebarre, A. An experimental investigation of loop-seal operation in an intercon-
nected circulating fluidized bed system. Powder Technol. 2013, 237, 266–275. [CrossRef]
136. Moldenhauer, P.; Rydén, M.; Mattisson, T.; Lyngfelt, A. Chemical-looping combustion and chemical-looping with oxygen
uncoupling of kerosene with Mn-and Cu-based oxygen carriers in a circulating fluidized-bed 300 W laboratory reactor. Fuel
Process. Technol. 2012, 104, 378–389. [CrossRef]
137. Ku, Y.; Wu, H.-C.; Chiu, P.-C.; Tseng, Y.-H.; Kuo, Y.-L. Methane combustion by moving bed fuel reactor with Fe2O3/Al2O3
oxygen carriers. Appl. Energy 2014, 113, 1909–1915. [CrossRef]
138. Tong, A.; Sridhar, D.; Sun, Z.; Kim, H.R.; Zeng, L.; Wang, F.; Wang, D.; Kathe, M.V.; Luo, S.; Sun, Y. Continuous high purity
hydrogen generation from a syngas chemical looping 25 kWth sub-pilot unit with 100% carbon capture. Fuel 2013, 103, 495–505.
[CrossRef]
139. Wang, J.; Anthony, E.J. Clean combustion of solid fuels. Appl. Energy 2008, 85, 73–79. [CrossRef]
140. Jin, H.; Ishida, M. A new type of coal gas fueled chemical-looping combustion. Fuel 2004, 83, 2411–2417. [CrossRef]
141. Cao, Y.; Pan, W.-P. Investigation of chemical looping combustion by solid fuels. 1. Process analysis. Energy Fuels 2006,
20, 1836–1844. [CrossRef]
142. Adánez-Rubio, I.; Abad, A.; Gayán, P.; de Diego, L.; García-Labiano, F.; Adánez, J. Biomass combustion with CO2 capture by
chemical looping with oxygen uncoupling (CLOU). Fuel Process. Technol. 2014, 124, 104–114. [CrossRef]
143. Gayán, P.; Abad, A.; de Diego, L.; García-Labiano, F.; Adánez, J. Assessment of technological solutions for improving chemical
looping combustion of solid fuels with CO2 capture. Chem. Eng. J. 2013, 233, 56–69. [CrossRef]
144. Adánez-Rubio, I.; Abad, A.; Gayán, P.; de Diego, L.; García-Labiano, F.; Adánez, J. Performance of CLOU process in the
combustion of different types of coal with CO2 capture. Int. J. Greenh. Gas Control 2013, 12, 430–440. [CrossRef]
145. Yang, H.; Xu, Z.; Fan, M.; Gupta, R.; Slimane, R.B.; Bland, A.E.; Wright, I. Progress in carbon dioxide separation and capture: A
review. J. Environ. Sci. 2008, 20, 14–27. [CrossRef] [PubMed]
146. Khalilpour, R.; Mumford, K.; Zhai, H.; Abbas, A.; Stevens, G.; Rubin, E.S. Membrane-based carbon capture from flue gas: A
review. J. Clean. Prod. 2015, 103, 286–300. [CrossRef]
Atmosphere 2022, 13, 1958 34 of 42

147. Ansaloni, L.; Salas-Gay, J.; Ligi, S.; Baschetti, M.G. Nanocellulose-based membranes for CO2 capture. J. Membr. Sci. 2017, 522,
216–225. [CrossRef]
148. Li, X.; Cheng, Y.; Zhang, H.; Wang, S.; Jiang, Z.; Guo, R.; Wu, H. Efficient CO2 capture by functionalized graphene oxide
nanosheets as fillers to fabricate multi-permselective mixed matrix membranes. ACS Appl. Mater. Interfaces 2015, 7, 5528–5537.
[CrossRef]
149. He, X.; Fu, C.; Hägg, M.-B. Membrane system design and process feasibility analysis for CO2 capture from flue gas with a
fixed-site-carrier membrane. Chem. Eng. J. 2015, 268, 1–9. [CrossRef]
150. Turi, D.M.; Ho, M.; Ferrari, M.-C.; Chiesa, P.; Wiley, D.; Romano, M.C. CO2 capture from natural gas combined cycles by CO2
selective membranes. Int. J. Greenh. Gas Control 2017, 61, 168–183. [CrossRef]
151. Vinoba, M.; Bhagiyalakshmi, M.; Alqaheem, Y.; Alomair, A.A.; Pérez, A.; Rana, M.S. Recent progress of fillers in mixed matrix
membranes for CO2 separation: A review. Sep. Purif. Technol. 2017, 188, 431–450. [CrossRef]
152. Lee, P.S.; Lim, M.S.; Park, A.; Park, H.; Nam, S.-E.; Park, Y.-I. A zeolite membrane module composed of SAPO-34 hollow fibers for
use in fluorinated gas enrichment. J. Membr. Sci. 2017, 542, 123–132. [CrossRef]
153. Brunetti, A.; Scura, F.; Barbieri, G.; Drioli, E. Membrane technologies for CO2 separation. J. Membr. Sci. 2010, 359, 115–125.
[CrossRef]
154. Favre, E. Membrane processes and postcombustion carbon dioxide capture: Challenges and prospects. Chem. Eng. J. 2011,
171, 782–793. [CrossRef]
155. Bernardo, P.; Drioli, E.; Golemme, G. Membrane gas separation: A review/state of the art. Ind. Eng. Chem. Res. 2009,
48, 4638–4663. [CrossRef]
156. Wang, L.; Boutilier, M.S.; Kidambi, P.R.; Jang, D.; Hadjiconstantinou, N.G.; Karnik, R. Fundamental transport mechanisms,
fabrication and potential applications of nanoporous atomically thin membranes. Nat. Nanotechnol. 2017, 12, 509–522. [CrossRef]
[PubMed]
157. Olajire, A.A. CO2 capture and separation technologies for end-of-pipe applications—A review. Energy 2010, 35, 2610–2628.
[CrossRef]
158. Luis, P.; van Gerven, T.; van der Bruggen, B. Recent developments in membrane-based technologies for CO2 capture. Prog. Energy
Combust. Sci. 2012, 38, 419–448. [CrossRef]
159. Paul, S.; Ghoshal, A.K.; Mandal, B. Theoretical studies on separation of CO2 by single and blended aqueous alkanolamine
solvents in flat sheet membrane contactor (FSMC). Chem. Eng. J. 2008, 144, 352–360. [CrossRef]
160. Andersen, A.; Divekar, S.; Dasgupta, S.; Cavka, J.H.; Nanoti, A.; Spjelkavik, A.; Goswami, A.N.; Garg, M.; Blom, R. On the
development of Vacuum Swing adsorption (VSA) technology for post-combustion CO2 capture. Energy Procedia 2013, 37, 33–39.
[CrossRef]
161. Wang, Y.; Zhao, L.; Otto, A.; Robinius, M.; Stolten, D. A review of post-combustion CO2 capture technologies from coal-fired
power plants. Energy Procedia 2017, 114, 650–665. [CrossRef]
162. Knapik, E.; Kosowski, P.; Stopa, J. Cryogenic liquefaction and separation of CO2 using nitrogen removal unit cold energy. Chem.
Eng. Res. Des. 2018, 131, 66–79. [CrossRef]
163. Tan, Y.; Nookuea, W.; Li, H.; Thorin, E.; Yan, J. Cryogenic technology for biogas upgrading combined with carbon capture-a
review of systems and property impacts. Energy Procedia 2017, 142, 3741–3746. [CrossRef]
164. Damen, K.; van Troost, M.; Faaij, A.; Turkenburg, W. A comparison of electricity and hydrogen production systems with CO2
capture and storage. Part A: Review and selection of promising conversion and capture technologies. Prog. Energy Combust. Sci.
2006, 32, 215–246. [CrossRef]
165. Rubin, E.S.; Rao, A.B.; Chen, C. Comparative Assessments of Fossil Fuel Power Plants with CO. In Proceedings of the 7th
International Conference on Greenhouse Gas Control Technologies, Vancouver, BC, Canada, 5–9 September 2004.
166. Porter, R.T.; Fairweather, M.; Pourkashanian, M.; Woolley, R.M. The range and level of impurities in CO2 streams from different
carbon capture sources. Int. J. Greenh. Gas Control 2015, 36, 161–174. [CrossRef]
167. Kanniche, M.; Le Moullec, Y.; Authier, O.; Hagi, H.; Bontemps, D.; Neveux, T.; Louis-Louisy, M. Up-to-date CO2 capture in
thermal power plants. Energy Procedia 2017, 114, 95–103. [CrossRef]
168. Dillon, D.; White, V.; Allam, R.; Wall, R.; Gibbins, J. Oxy Combustion Processes for CO2 Capture from Power Plant. In Technology
& Engineering; Engineering Investigation Report; Mitsui Babcock Energy Limited: Crawley, UK, 2005; Volume 9.
169. Zheng, L. Oxy-Fuel Combustion for Power Generation and Carbon Dioxide (CO2 ) Capture; Elsevier: Amsterdam, The Netherlands, 2011.
170. Bhown, A.S.; Freeman, B.C. Analysis and status of post-combustion carbon dioxide capture technologies. Environ. Sci. Technol.
2011, 45, 8624–8632. [CrossRef] [PubMed]
171. Hendriks, C. Energy Conversion: CO2 Removal from Coal-Fired Power Plant; Kluwer Academic Publishers: Dordrecht, The Nether-
lands, 1995.
172. Ritter, J.A. Radically new adsorption cycles for carbon dioxide sequestration. In Proceedings of the University Coal Research
Contractors Review Meeting, U.S. DOE National Energy Technology Laboratory, Pittsburgh, PA, USA, 10 June 2004; Volume 2, p. 2.
173. Olabi, A.; Obaideen, K.; Elsaid, K.; Wilberforce, T.; Sayed, E.T.; Maghrabie, H.M.; Abdelkareem, M.A. Assessment of the pre-
combustion carbon capture contribution into sustainable development goals SDGs using novel indicators. Renew. Sustain. Energy
Rev. 2022, 153, 111710. [CrossRef]
Atmosphere 2022, 13, 1958 35 of 42

174. Lyngfelt, A.; Linderholm, C. Chemical-looping combustion of solid fuels–technology overview and recent operational results in
100 kW unit. Energy Procedia 2014, 63, 98–112. [CrossRef]
175. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gayan, P.; Luis, F. Progress in chemical-looping combustion and reforming technologies.
Prog. Energy Combust. Sci. 2012, 38, 215–282. [CrossRef]
176. Rackley, S.A. Carbon Capture and Storage; Butterworth-Heinemann: Oxford, UK, 2017.
177. Gielen, D. The energy policy consequences of future CO2 capture and sequestration technologies. In Proceedings of the 2nd
Annual Conference on Carbon Sequestration, Alexandria, VA, USA, 5–8 May 2003; pp. 5–8.
178. Audus, H. Leading options for the capture of CO2 at power stations. In Proceedings of the 5th International Conference on
Greenhouse Gas Control Technologies, Cairns, Australia, 13–16 August 2000; p. 16.
179. Göttlicher, G.; Pruschek, R. Comparison of CO2 removal systems for fossil-fuelled power plant processes. Energy Convers. Manag.
1997, 38, S173–S178. [CrossRef]
180. Tuinier, M.; van Sint Annaland, M.; Kramer, G.J.; Kuipers, J. Cryogenic CO2 capture using dynamically operated packed beds.
Chem. Eng. Sci. 2010, 65, 114–119. [CrossRef]
181. Zhang, X.; Song, Z.; Gani, R.; Zhou, T. Comparative economic analysis of physical, chemical, and hybrid absorption processes for
carbon capture. Ind. Eng. Chem. Res. 2020, 59, 2005–2012. [CrossRef]
182. Li, J.; Zhang, H.; Gao, Z.; Fu, J.; Ao, W.; Dai, J. CO2 capture with chemical looping combustion of gaseous fuels: An overview.
Energy Fuels 2017, 31, 3475–3524. [CrossRef]
183. Castel, C.; Bounaceur, R.; Favre, E. Membrane processes for direct carbon dioxide capture from air: Possibilities and limitations.
Front. Chem. Eng. 2021, 3, 668867. [CrossRef]
184. Zhai, H.; Rubin, E.S. The effects of membrane-based CO2 capture system on pulverized coal power plant performance and cost.
Energy Procedia 2013, 37, 1117–1124. [CrossRef]
185. Cuéllar-Franca, R.M.; Azapagic, A. Carbon capture, storage and utilisation technologies: A critical analysis and comparison of
their life cycle environmental impacts. J. CO2 Util. 2015, 9, 82–102. [CrossRef]
186. Lyngfelt, A.; Leckner, B. A 1000 MWth boiler for chemical-looping combustion of solid fuels–Discussion of design and costs. Appl.
Energy 2015, 157, 475–487. [CrossRef]
187. Moldenhauer, P.; Linderholm, C.; Rydén, M.; Lyngfelt, A. Avoiding CO2 capture effort and cost for negative CO2 emissions using
industrial waste in chemical-looping combustion/gasification of biomass. Mitig. Adapt. Strateg. Glob. Change 2020, 25, 1–24.
[CrossRef]
188. Ma’mum, S.; Svendsen, H.F.; Hoff, K.A.; Juliussen, O. Selection of new absorbents for carbon dioxide capture. In Greenhouse Gas
Control Technologies 7; Elsevier: Amsterdam, The Netherlands, 2005; pp. 45–53.
189. Balsamo, M.; Rodríguez-Reinoso, F.; Montagnaro, F.; Lancia, A.; Erto, A. Highlighting the role of activated carbon particle size on
CO2 capture from model flue gas. Ind. Eng. Chem. Res. 2013, 52, 12183–12191. [CrossRef]
190. Balsamo, M.; Budinova, T.; Erto, A.; Lancia, A.; Petrova, B.; Petrov, N.; Tsyntsarski, B. CO2 adsorption onto synthetic activated
carbon: Kinetic, thermodynamic and regeneration studies. Sep. Purif. Technol. 2013, 116, 214–221. [CrossRef]
191. Sculley, J.P.; Verdegaal, W.M.; Lu, W.; Wriedt, M.; Zhou, H.C. High-Throughput Analytical Model to Evaluate Materials for
Temperature Swing Adsorption Processes. Adv. Mater. 2013, 25, 3957–3961. [CrossRef]
192. Pirngruber, G.D.; Guillou, F.; Gomez, A.; Clausse, M. A theoretical analysis of the energy consumption of post-combustion CO2
capture processes by temperature swing adsorption using solid sorbents. Int. J. Greenh. Gas Control 2013, 14, 74–83. [CrossRef]
193. Datta, S.; Henry, M.P.; Lin, Y.J.; Fracaro, A.T.; Millard, C.S.; Snyder, S.W.; Stiles, R.L.; Shah, J.; Yuan, J.; Wesoloski, L. Electrochemical
CO2 capture using resin-wafer electrodeionization. Ind. Eng. Chem. Res. 2013, 52, 15177–15186. [CrossRef]
194. Huang, Q.; Eić, M. Commercial adsorbents as benchmark materials for separation of carbon dioxide and nitrogen by vacuum
swing adsorption process. Sep. Purif. Technol. 2013, 103, 203–215. [CrossRef]
195. Goeppert, A.; Czaun, M.; May, R.B.; Prakash, G.S.; Olah, G.A.; Narayanan, S. Carbon dioxide capture from the air using a
polyamine based regenerable solid adsorbent. J. Am. Chem. Soc. 2011, 133, 20164–20167. [CrossRef] [PubMed]
196. Kim, Y.H.; Lee, D.G.; Moon, D.K.; Byeon, S.-H.; Ahn, H.W.; Lee, C.H. Effect of bed void volume on pressure vacuum swing
adsorption for air separation. Korean J. Chem. Eng. 2014, 31, 132–141. [CrossRef]
197. Chou, C.-T.; Chen, C.-Y. Carbon dioxide recovery by vacuum swing adsorption. Sep. Purif. Technol. 2004, 39, 51–65. [CrossRef]
198. Prenzel, T.; Wilhelm, M.; Rezwan, K. Tailoring amine functionalized hybrid ceramics to control CO2 adsorption. Chem. Eng. J.
2014, 235, 198–206. [CrossRef]
199. Hicks, J.C.; Drese, J.H.; Fauth, D.J.; Gray, M.L.; Qi, G.; Jones, C.W. Designing adsorbents for CO2 capture from flue gas-
hyperbranched aminosilicas capable of capturing CO2 reversibly. J. Am. Chem. Soc. 2008, 130, 2902–2903. [CrossRef]
200. Lee, S.-Y.; Park, S.-J. A review on solid adsorbents for carbon dioxide capture. J. Ind. Eng. Chem. 2015, 23, 1–11. [CrossRef]
201. Auta, M.; Darbis, N.A.; Din, A.M.; Hameed, B. Fixed-bed column adsorption of carbon dioxide by sodium hydroxide modified
activated alumina. Chem. Eng. J. 2013, 233, 80–87. [CrossRef]
202. Nugent, P.; Belmabkhout, Y.; Burd, S.D.; Cairns, A.J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L. Porous
materials with optimal adsorption thermodynamics and kinetics for CO2 separation. Nature 2013, 495, 80–84. [CrossRef]
203. Wei, H.; Deng, S.; Hu, B.; Chen, Z.; Wang, B.; Huang, J.; Yu, G. Granular bamboo-derived activated carbon for high CO2
adsorption: The dominant role of narrow micropores. ChemSusChem 2012, 5, 2354–2360. [CrossRef]
Atmosphere 2022, 13, 1958 36 of 42

204. Chang, F.-Y.; Chao, K.-J.; Cheng, H.-H.; Tan, C.-S. Adsorption of CO2 onto amine-grafted mesoporous silicas. Sep. Purif. Technol.
2009, 70, 87–95. [CrossRef]
205. Plaza, M.; García, S.; Rubiera, F.; Pis, J.; Pevida, C. Post-combustion CO2 capture with a commercial activated carbon: Comparison
of different regeneration strategies. Chem. Eng. J. 2010, 163, 41–47. [CrossRef]
206. Adelodun, A.A.; Lim, Y.H.; Jo, Y.M. Surface oxidation of activated carbon pellets by hydrogen peroxide for preparation of CO2
adsorbent. J. Ind. Eng. Chem. 2014, 20, 2130–2137. [CrossRef]
207. Ho, M.T.; Allinson, G.W.; Wiley, D.E. Reducing the cost of CO2 capture from flue gases using pressure swing adsorption. Ind. Eng.
Chem. Res. 2008, 47, 4883–4890. [CrossRef]
208. Abd, A.A.; Naji, S.Z.; Hashim, A.S.; Othman, M.R. Carbon dioxide removal through physical adsorption using carbonaceous and
non-carbonaceous adsorbents: A review. J. Environ. Chem. Eng. 2020, 8, 104142. [CrossRef]
209. Wang, M.; Wang, G.; Li, G.; Huang, B.; Pan, J.; Liu, Q.; Han, J.; Xiao, L.; Lu, J.; Zhuang, L. Energy Environ. Sci 2012, 5, 5163–5185.
210. Yu, F.-C.; Phalak, N.; Sun, Z.; Fan, L.-S. Activation strategies for calcium-based sorbents for CO2 capture: A perspective. Ind. Eng.
Chem. Res. 2012, 51, 2133–2142. [CrossRef]
211. Buckingham, J.; Reina, T.R.; Duyar, M.S. Recent advances in carbon dioxide capture for process intensification. Carbon Capture Sci.
Technol. 2022, 298, 100031. [CrossRef]
212. Siriwardane, R.V.; Shen, M.-S.; Fisher, E.P. Adsorption of CO2 , N2 , and O2 on natural zeolites. Energy Fuels 2003, 17, 571–576.
[CrossRef]
213. Xu, X.; Andresen, J.M.; Song, C.; Miller, B.G.; Scaroni, A.W. Preparation of novel CO2 “molecular basket” of polymer modified
MCM-41. ACS Div. Fuel Chem. Prepr. 2002, 47, 67–68.
214. Gallei, E.; Stumpf, G. Infrared spectroscopic studies of the adsorption of carbon dioxide and the coadsorption of carbon dioxide
and water on CaY-and NiY-zeolites. J. Colloid Interface Sci. 1976, 55, 415–420. [CrossRef]
215. Zhao, D.; Cleare, K.; Oliver, C.; Ingram, C.; Cook, D.; Szostak, R.; Kevan, L. Characteristics of the synthetic heulandite-clinoptilolite
family of zeolites. Microporous Mesoporous Mater. 1998, 21, 371–379. [CrossRef]
216. Kumar, S.; Srivastava, R.; Koh, J. Utilization of zeolites as CO2 capturing agents: Advances and future perspectives. J. CO2 Util.
2020, 41, 101251. [CrossRef]
217. Albahar, M.Z. Selective Toluene Disproportionation Over zsm-5 Zeolite; The University of Manchester: Manchester, UK, 2018.
218. Omodolor, I.S.; Otor, H.O.; Andonegui, J.A.; Allen, B.J.; Alba-Rubio, A.C. Dual-function materials for CO2 capture and conversion:
A review. Ind. Eng. Chem. Res. 2020, 59, 17612–17631. [CrossRef]
219. Lee, S.C.; Kwon, Y.M.; Jung, S.Y.; Lee, J.B.; Ryu, C.K.; Kim, J.C. Excellent thermal stability of potassium-based sorbent using ZrO2
for post combustion CO2 capture. Fuel 2014, 115, 97–100. [CrossRef]
220. Harlick, P.J.; Sayari, A. Applications of pore-expanded mesoporous silica. 5. Triamine grafted material with exceptional CO2
dynamic and equilibrium adsorption performance. Ind. Eng. Chem. Res. 2007, 46, 446–458. [CrossRef]
221. Moura, P.; Bezerra, D.; Vilarrasa-Garcia, E.; Bastos-Neto, M.; Azevedo, D. Adsorption equilibria of CO2 and CH4 in cation-
exchanged zeolites 13X. Adsorption 2016, 22, 71–80. [CrossRef]
222. Walton, K.S.; Abney, M.B.; LeVan, M.D. CO2 adsorption in Y and X zeolites modified by alkali metal cation exchange. Microporous
Mesoporous Mater. 2006, 91, 78–84. [CrossRef]
223. Calleja, G.; Jimenez, A.; Pau, J.; Dominguez, L.; Perez, P. Multicomponent adsorption equilibrium of ethylene, propane, propylene
and CO2 on 13X zeolite. Gas Sep. Purif. 1994, 8, 247–256. [CrossRef]
224. I Xamena, F.X.L.; Abad, A.; Corma, A.; Garcia, H. MOFs as catalysts: Activity, reusability and shape-selectivity of a Pd-containing
MOF. J. Catal. 2007, 250, 294–298.
225. Ravon, U.; Domine, M.E.; Gaudillere, C.; Desmartin-Chomel, A.; Farrusseng, D. MOFs as acid catalysts with shape selectivity
properties. New J. Chem. 2008, 32, 937–940. [CrossRef]
226. Britt, D.; Furukawa, H.; Wang, B.; Glover, T.G.; Yaghi, O.M. Highly efficient separation of carbon dioxide by a metal-organic
framework replete with open metal sites. Proc. Natl. Acad. Sci. USA 2009, 106, 20637–20640. [CrossRef]
227. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.-H.; Long, J.R. Carbon dioxide capture in
metal–organic frameworks. Chem. Rev. 2012, 112, 724–781. [CrossRef]
228. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, S.B.; Choi, E.; Yazaydin, A.Ö.; Snurr, R.Q.; O’Keeffe, M.; Kim, J. Ultrahigh
porosity in metal-organic frameworks. Science 2010, 329, 424–428. [CrossRef]
229. Gandara-Loe, J.; Pastor-Perez, L.; Bobadilla, L.; Odriozola, J.; Reina, T. Understanding the opportunities of metal–organic
frameworks (MOFs) for CO2 capture and gas-phase CO2 conversion processes: A comprehensive overview. React. Chem. Eng.
2021, 6, 787–814. [CrossRef]
230. Szcz˛eśniak, B.; Choma, J. Graphene-containing microporous composites for selective CO2 adsorption. Microporous Mesoporous
Mater. 2020, 292, 109761. [CrossRef]
231. Aarti, A.; Bhadauria, S.; Nanoti, A.; Dasgupta, S.; Divekar, S.; Gupta, P.; Chauhan, R. [Cu3 (BTC) 2]-polyethyleneimine: An
efficient MOF composite for effective CO2 . Rsc Adv. 2016, 6, 93003–93009. [CrossRef]
232. Caskey, S.R.; Wong-Foy, A.G.; Matzger, A.J. Dramatic tuning of carbon dioxide uptake via metal substitution in a coordination
polymer with cylindrical pores. J. Am. Chem. Soc. 2008, 130, 10870–10871. [CrossRef]
233. Bao, Z.; Yu, L.; Ren, Q.; Lu, X.; Deng, S. Adsorption of CO2 and CH4 on a magnesium-based metal organic framework. J. Colloid
Interface Sci. 2011, 353, 549–556. [CrossRef]
Atmosphere 2022, 13, 1958 37 of 42

234. Mahdipoor, H.R.; Halladj, R.; Babakhani, E.G.; Amjad-Iranagh, S.; Ahari, J.S. Synthesis, characterization, and CO2 adsorption
properties of metal organic framework Fe-BDC. RSC Adv. 2021, 11, 5192–5203. [CrossRef]
235. Yu, C.; Huang, C.; Tan, C. A review of CO2 capture by absorption and adsorption. Aerosol. Air Qual. Res. 2012, 12, 745–769.
[CrossRef]
236. Sánchez-Zambrano, K.S.; Lima Duarte, L.; Soares Maia, D.A.; Vilarrasa-García, E.; Bastos-Neto, M.; Rodríguez-Castellón, E.;
Silva de Azevedo, D.C. CO2 capture with mesoporous silicas modified with amines by double functionalization: Assessment of
adsorption/desorption cycles. Materials 2018, 11, 887. [CrossRef] [PubMed]
237. Zhao, C.; Chen, X.; Anthony, E.J.; Jiang, X.; Duan, L.; Wu, Y.; Dong, W.; Zhao, C. Capturing CO2 in flue gas from fossil fuel-fired
power plants using dry regenerable alkali metal-based sorbent. Prog. Energy Combust. Sci. 2013, 39, 515–534. [CrossRef]
238. Duan, Y.; Luebke, D.R.; Pennline, H.W.; Li, B.; Janik, M.J.; Halley, J.W. Ab initio thermodynamic study of the CO2 capture
properties of potassium carbonate sesquihydrate, K2CO3 1.5 H2O. J. Phys. Chem. C 2012, 116, 14461–14470. [CrossRef]
239. Essaki, K.; Kato, M.; Uemoto, H. Influence of temperature and CO2 concentration on the CO2 absorption properties of lithium
silicate pellets. J. Mater. Sci. 2005, 40, 5017–5019. [CrossRef]
240. Venegas, M.J.; Fregoso-Israel, E.; Escamilla, R.; Pfeiffer, H. Kinetic and reaction mechanism of CO2 sorption on Li4SiO4: Study of
the particle size effect. Ind. Eng. Chem. Res. 2007, 46, 2407–2412. [CrossRef]
241. Pacciani, R.; Torres, J.; Solsona, P.; Coe, C.; Quinn, R.; Hufton, J.; Golden, T.; Vega, L. Influence of the Concentration of CO2 and
SO2 on the Absorption of CO2 by a Lithium Orthosilicate-Based Absorbent. Environ. Sci. Technol. 2011, 45, 7083–7088. [CrossRef]
242. Seggiani, M.; Puccini, M.; Vitolo, S. Alkali promoted lithium orthosilicate for CO2 capture at high temperature and low
concentration. Int. J. Greenh. Gas Control 2013, 17, 25–31. [CrossRef]
243. Durán-Muñoz, F.; Romero-Ibarra, I.C.; Pfeiffer, H. Analysis of the CO2 chemisorption reaction mechanism in lithium oxosilicate
(Li 8 SiO 6): A new option for high-temperature CO2 capture. J. Mater. Chem. A 2013, 1, 3919–3925. [CrossRef]
244. Memon, M.Z.; Zhao, X.; Sikarwar, V.S.; Vuppaladadiyam, A.K.; Milne, S.J.; Brown, A.P.; Li, J.; Zhao, M. Alkali metal CO2 sorbents
and the resulting metal carbonates: Potential for process intensification of sorption-enhanced steam reforming. Environ. Sci.
Technol. 2017, 51, 12–27. [CrossRef]
245. Alcántar-Vázquez, B.; Vera, E.; Buitron-Cabrera, F.; Lara-García, H.A.; Pfeiffer, H. Evidence of CO oxidation–chemisorption
process on sodium zirconate (Na2 ZrO3 ). Chem. Lett. 2015, 44, 480–482. [CrossRef]
246. Sutton, D.; Kelleher, B.; Ross, J.R. Review of literature on catalysts for biomass gasification. Fuel Process. Technol. 2001, 73, 155–173.
[CrossRef]
247. Palacios-Romero, L.M.; Pfeiffer, H. Lithium cuprate (Li2CuO2): A new possible ceramic material for CO2 chemisorption. Chem.
Lett. 2008, 37, 862–863. [CrossRef]
248. Oh-ishi, K.; Matsukura, Y.; Okumura, T.; Matsunaga, Y.; Kobayashi, R. Fundamental research on gas–solid reaction between CO2
and Li2CuO2 linking application for solid CO2 absorbent. J. Solid State Chem. 2014, 211, 162–169. [CrossRef]
249. Bernabé-Pablo, E.; Duan, Y.; Pfeiffer, H. Developing new alkaline ceramics as possible CO2 chemisorbents at high temperatures:
The lithium and sodium yttriates (LiYO2 and NaYO2) cases. Chem. Eng. J. 2020, 396, 125277. [CrossRef]
250. D’Alessandro, D.M.; Smit, B.; Long, J.R. Carbon dioxide capture: Prospects for new materials. Angew. Chem. Int. Ed. 2010,
49, 6058–6082. [CrossRef]
251. Hao, G.-P.; Li, W.-C.; Lu, A.-H. Novel porous solids for carbon dioxide capture. J. Mater. Chem. 2011, 21, 6447–6451. [CrossRef]
252. Wang, Q.; Luo, J.; Zhong, Z.; Borgna, A. CO2 capture by solid adsorbents and their applications: Current status and new trends.
Energy Environ. Sci. 2011, 4, 42–55. [CrossRef]
253. Park, S.-J.; Kim, K.-D. Adsorption Behaviors of CO2 NH3on Chemically Surface-Treated Activated Carbons. J. Colloid Interface Sci.
1999, 212, 186–189. [CrossRef]
254. Seo, M.K.; Park, S.J. A Kinetic Study on the Thermal Degradation of Multi-Walled Carbon Nanotubes-Reinforced Poly (propylene)
Composites. Macromol. Mater. Eng. 2004, 289, 368–374. [CrossRef]
255. Bilalis, P.; Katsigiannopoulos, D.; Avgeropoulos, A.G. Sakellariou. RSC Adv. 2014, 4, 2911. [CrossRef]
256. Gupta, V.K.; Saleh, T.A. Sorption of pollutants by porous carbon, carbon nanotubes and fullerene-an overview. Environ. Sci.
Pollut. Res. 2013, 20, 2828–2843. [CrossRef] [PubMed]
257. Seema, H.; Kemp, K.C.; Le, N.H.; Park, S.-W.; Chandra, V.; Lee, J.W.; Kim, K.S. Highly selective CO2 capture by S-doped
microporous carbon materials. Carbon 2014, 66, 320–326. [CrossRef]
258. Bai, B.C.; Cho, S.; Yu, H.-R.; Yi, K.B.; Kim, K.-D.; Lee, Y.-S. Effects of aminated carbon molecular sieves on breakthrough curve
behavior in CO2 /CH4 separation. J. Ind. Eng. Chem. 2013, 19, 776–783. [CrossRef]
259. Yu, J.; Guo, M.; Muhammad, F.; Wang, A.; Zhang, F.; Li, Q.; Zhu, G. One-pot synthesis of highly ordered nitrogen-containing
mesoporous carbon with resorcinol–urea–formaldehyde resin for CO2 capture. Carbon 2014, 69, 502–514. [CrossRef]
260. Guo, B.; Chang, L.; Xie, K. Adsorption of carbon dioxide on activated carbon. J. Nat. Gas Chem. 2006, 15, 223–229. [CrossRef]
261. Wang, S.; Xu, Y.; Miao, J.; Liu, M.; Ren, B.; Zhang, L.; Liu, Z. Facile synthesis of microporous carbon xerogels for highly selective
CO2 adsorption. J. Clean. Prod. 2020, 253, 120023. [CrossRef]
262. Sevilla, M.; Fuertes, A.B. CO2 adsorption by activated templated carbons. J. Colloid Interface Sci. 2012, 366, 147–154. [CrossRef]
263. Wang, J.; Kaskel, S. KOH activation of carbon-based materials for energy storage. J. Mater. Chem. 2012, 22, 23710–23725. [CrossRef]
264. Yoo, H.-M.; Lee, S.-Y.; Kim, B.-J.; Park, S.-J. Influence of phosphoric acid treatment on hydrogen adsorption behaviors of activated
carbons. Carbon Lett. 2011, 12, 112–115. [CrossRef]
Atmosphere 2022, 13, 1958 38 of 42

265. Kim, K.-S.; Park, S.-J. Synthesis of nitrogen doped microporous carbons prepared by activation-free method and their high
electrochemical performance. Electrochim. Acta 2011, 56, 10130–10136. [CrossRef]
266. Seo, M.-K.; Park, S.-J. Influence of air-oxidation on electric double layer capacitances of multi-walled carbon nanotube electrodes.
Curr. Appl. Phys. 2010, 10, 241–244. [CrossRef]
267. Pramanik, P.; Ray, P.; Maity, A.; Das, S.; Ramakrishnan, S.; Dixit, P. Nanotechnology for improved carbon management in soil. In
Carbon Management in Tropical and Sub-Tropical Terrestrial Systems; Springer: Singapore, 2020; pp. 403–415.
268. Meng, L.-Y.; Park, S.-J. MgO-templated porous carbons-based CO2 adsorbents produced by KOH activation. Mater. Chem. Phys.
2012, 137, 91–96. [CrossRef]
269. Jang, D.-I.; Park, S.-J. Influence of nickel oxide on carbon dioxide adsorption behaviors of activated carbons. Fuel 2012,
102, 439–444. [CrossRef]
270. Chua, C.K.; Pumera, M. Chemical reduction of graphene oxide: A synthetic chemistry viewpoint. Chem. Soc. Rev. 2014,
43, 291–312. [CrossRef]
271. Liu, W.-W.; Chai, S.-P.; Mohamed, A.R.; Hashim, U. Synthesis and characterization of graphene and carbon nanotubes: A review
on the past and recent developments. J. Ind. Eng. Chem. 2014, 20, 1171–1185. [CrossRef]
272. Mittal, G.; Dhand, V.; Rhee, K.Y.; Park, S.-J.; Lee, W.R. A review on carbon nanotubes and graphene as fillers in reinforced polymer
nanocomposites. J. Ind. Eng. Chem. 2015, 21, 11–25. [CrossRef]
273. Liu, Y.; Park, M.; Shin, H.K.; Pant, B.; Choi, J.; Park, Y.W.; Lee, J.Y.; Park, S.-J.; Kim, H.-Y. Facile preparation and characterization
of poly (vinyl alcohol)/chitosan/graphene oxide biocomposite nanofibers. J. Ind. Eng. Chem. 2014, 20, 4415–4420. [CrossRef]
274. Moradi, S. Microwave assisted preparation of sodium dodecyl sulphate (SDS) modified ordered nanoporous carbon and its
adsorption for MB dye. J. Ind. Eng. Chem. 2014, 20, 208–215. [CrossRef]
275. Dubey, S.P.; Dwivedi, A.D.; Lee, C.; Kwon, Y.-N.; Sillanpaa, M.; Ma, L.Q. Raspberry derived mesoporous carbon-tubules and
fixed-bed adsorption of pharmaceutical drugs. J. Ind. Eng. Chem. 2014, 20, 1126–1132. [CrossRef]
276. How, C.; Khan, M.A.; Hosseini, S.; Chuah, T.; Choong, T.S. Fabrication of mesoporous carbons coated monolith via evaporative
induced self-assembly approach: Effect of solvent and acid concentration on pore architecture. J. Ind. Eng. Chem. 2014, 20, 4286–4292.
[CrossRef]
277. Liu, H.; Ding, W.; Zhou, F.; Yang, G.; Du, Y. An overview and outlook on gas adsorption: For the enrichment of low concentration
coalbed methane. Sep. Sci. Technol. 2020, 55, 1102–1114. [CrossRef]
278. Karthikeyan, S.; Viswanathan, K.; Boopathy, R.; Maharaja, P.; Sekaran, G. Three dimensional electro catalytic oxidation of aniline
by boron doped mesoporous activated carbon. J. Ind. Eng. Chem. 2015, 21, 942–950. [CrossRef]
279. Yoo, H.-M.; Lee, S.-Y.; Park, S.-J. Ordered nanoporous carbon for increasing CO2 capture. J. Solid State Chem. 2013, 197, 361–365.
[CrossRef]
280. Im, J.S.; Park, S.-J.; Lee, Y.-S. Superior prospect of chemically activated electrospun carbon fibers for hydrogen storage. Mater. Res.
Bull. 2009, 44, 1871–1878. [CrossRef]
281. Kim, B.-J.; Park, S.-J. A simple method for the preparation of activated carbon fibers coated with graphite nanofibers. J. Colloid
Interface Sci. 2007, 315, 791–794. [CrossRef]
282. Yoon, S.-H.; Lim, S.; Song, Y.; Ota, Y.; Qiao, W.; Tanaka, A.; Mochida, I. KOH activation of carbon nanofibers. Carbon 2004,
42, 1723–1729. [CrossRef]
283. Thiruvenkatachari, R.; Su, S.; Yu, X.X.; Bae, J.-S. Application of carbon fibre composites to CO2 capture from flue gas. Int. J.
Greenh. Gas Control 2013, 13, 191–200. [CrossRef]
284. Zhang, N.; Long, R.; Gao, C.; Xiong, Y. Recent progress on advanced design for photoelectrochemical reduction of CO2 to fuels.
Sci. China Mater. 2018, 61, 771–805. [CrossRef]
285. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2 O3 :
Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [CrossRef]
286. Rui, N.; Wang, Z.; Sun, K.; Ye, J.; Ge, Q.; Liu, C.-j. CO2 hydrogenation to methanol over Pd/In2 O3 : Effects of Pd and oxygen
vacancy. Appl. Catal. B Environ. 2017, 218, 488–497. [CrossRef]
287. Song, Y.; Zhang, X.; Xie, K.; Wang, G.; Bao, X. High-temperature CO2 electrolysis in solid oxide electrolysis cells: Developments,
challenges, and prospects. Adv. Mater. 2019, 31, 1902033. [CrossRef]
288. Currie, R.; Mottaghi-Tabar, S.; Zhuang, Y.; Simakov, D.S. Design of an Air-Cooled Sabatier Reactor for Thermocatalytic Hy-
drogenation of CO2 : Experimental Proof-of-Concept and Model-Based Feasibility Analysis. Ind. Eng. Chem. Res. 2019, 58,
12964–12980. [CrossRef]
289. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous
catalysis. Chem. Rev. 2020, 120, 7984–8034. [CrossRef]
290. Wang, T.; Meng, X.; Li, P.; Ouyang, S.; Chang, K.; Liu, G.; Mei, Z.; Ye, J. Photoreduction of CO2 over the well-crystallized ordered
mesoporous TiO2 with the confined space effect. Nano Energy 2014, 9, 50–60. [CrossRef]
291. Li, K.; Peng, B.; Peng, T. Recent advances in heterogeneous photocatalytic CO2 conversion to solar fuels. ACS Catal. 2016,
6, 7485–7527. [CrossRef]
292. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. TiO2 –MnO x–Pt hybrid multiheterojunction film photocatalyst with enhanced photocatalytic
CO2 -reduction activity. ACS Appl. Mater. Interfaces 2018, 11, 5581–5589. [CrossRef]
Atmosphere 2022, 13, 1958 39 of 42

293. Ran, J.; Jaroniec, M.; Qiao, S.Z. Cocatalysts in semiconductor-based photocatalytic CO2 reduction: Achievements, challenges, and
opportunities. Adv. Mater. 2018, 30, 1704649. [CrossRef]
294. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal–organic frameworks for
electrocatalytic reduction of carbon dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [CrossRef]
295. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially oxidized atomic cobalt layers for carbon
dioxide electroreduction to liquid fuel. Nature 2016, 529, 68–71. [CrossRef]
296. Huang, J.; Buonsanti, R. Colloidal nanocrystals as heterogeneous catalysts for electrochemical CO2 conversion. Chem. Mater. 2018,
31, 13–25. [CrossRef]
297. Wang, J.; Ji, Y.; Shao, Q.; Yin, R.; Guo, J.; Li, Y.; Huang, X. Phase and structure modulating of bimetallic CuSn nanowires boosts
electrocatalytic conversion of CO2 . Nano Energy 2019, 59, 138–145. [CrossRef]
298. Kim, C.; Cho, K.M.; Al-Saggaf, A.; Gereige, I.; Jung, H.-T. Z-scheme photocatalytic CO2 conversion on three-dimensional
BiVO4/carbon-coated Cu2O nanowire arrays under visible light. ACS Catal. 2018, 8, 4170–4177. [CrossRef]
299. Leung, J.J.; Warnan, J.; Ly, K.H.; Heidary, N.; Nam, D.H.; Kuehnel, M.F.; Reisner, E. Solar-driven reduction of aqueous CO2 with a
cobalt bis (terpyridine)-based photocathode. Nat. Catal. 2019, 2, 354–365. [CrossRef]
300. Chen, J.; Yin, J.; Zheng, X.; Ait Ahsaine, H.; Zhou, Y.; Dong, C.; Mohammed, O.F.; Takanabe, K.; Bakr, O.M. Compositionally
screened eutectic catalytic coatings on halide perovskite photocathodes for photoassisted selective CO2 reduction. ACS Energy
Lett. 2019, 4, 1279–1286. [CrossRef]
301. Kalamaras, E.; Belekoukia, M.; Tan, J.Z.; Xuan, J.; Maroto-Valer, M.M.; Andresen, J.M. A microfluidic photoelectrochemical cell for
solar-driven CO2 conversion into liquid fuels with CuO-based photocathodes. Faraday Discuss. 2019, 215, 329–344. [CrossRef]
[PubMed]
302. Sultana, S.; Sahoo, P.C.; Martha, S.; Parida, K. A review of harvesting clean fuels from enzymatic CO2 reduction. RSC Adv. 2016, 6,
44170–44194. [CrossRef]
303. Rudroff, F.; Mihovilovic, M.D.; Gröger, H.; Snajdrova, R.; Iding, H.; Bornscheuer, U.T. Opportunities and challenges for combining
chemo-and biocatalysis. Nat. Catal. 2018, 1, 12–22. [CrossRef]
304. Xu, L.; Xiu, Y.; Liu, F.; Liang, Y.; Wang, S. Research progress in conversion of CO2 to valuable fuels. Molecules 2020, 25, 3653.
[CrossRef] [PubMed]
305. Rajeshwar, K.; de Tacconi, N.R.; Ghadimkhani, G.; Chanmanee, W.; Janáky, C. Tailoring copper oxide semiconductor nanorod
arrays for photoelectrochemical reduction of carbon dioxide to methanol. ChemPhysChem 2013, 14, 2251–2259. [CrossRef]
[PubMed]
306. Yadav, D.; Yadav, R.K.; Kumar, A.; Park, N.J.; Kim, J.Y.; Baeg, J.O. Fullerene polymer film as a highly efficient photocatalyst for
selective solar fuel production from CO2 . J. Appl. Polym. Sci. 2020, 137, 48536. [CrossRef]
307. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni–Ru/γ-Al2 O3 via modulating
impregnation sequence and controlling surface active species. RSC Adv. 2014, 4, 16472–16479. [CrossRef]
308. Wang, Y.-R.; Huang, Q.; He, C.-T.; Chen, Y.; Liu, J.; Shen, F.-C.; Lan, Y.-Q. Oriented electron transmission in polyoxometalate-
metalloporphyrin organic framework for highly selective electroreduction of CO2 . Nat. Commun. 2018, 9, 1–8. [CrossRef]
309. Nam, D.H.; Kuk, S.K.; Choe, H.; Lee, S.; Ko, J.W.; Son, E.J.; Choi, E.-G.; Kim, Y.H.; Park, C.B. Enzymatic photosynthesis of formate
from carbon dioxide coupled with highly efficient photoelectrochemical regeneration of nicotinamide cofactors. Green Chem.
2016, 18, 5989–5993. [CrossRef]
310. Srikanth, S.; Maesen, M.; Dominguez-Benetton, X.; Vanbroekhoven, K.; Pant, D. Enzymatic electrosynthesis of formate through
CO2 sequestration/reduction in a bioelectrochemical system (BES). Bioresour. Technol. 2014, 165, 350–354. [CrossRef]
311. Yuan, J.; Hao, C. Solar-driven photoelectrochemical reduction of carbon dioxide to methanol at CuInS2 thin film photocathode.
Sol. Energy Mater. Sol. Cells 2013, 108, 170–174. [CrossRef]
312. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [CrossRef]
313. Chen, Y.; Chen, K.; Fu, J.; Yamaguchi, A.; Li, H.; Pan, H.; Hu, J.; Miyauchi, M.; Liu, M. Recent advances in the utilization of copper
sulfide compounds for electrochemical CO2 reduction. Nano Mater. Sci. 2020, 2, 235–247. [CrossRef]
314. Zha, B.; Li, C.; Li, J. Efficient electrochemical reduction of CO2 into formate and acetate in polyoxometalate catholyte with indium
catalyst. J. Catal. 2020, 382, 69–76. [CrossRef]
315. Xi, G.; Ouyang, S.; Li, P.; Ye, J.; Ma, Q.; Su, N.; Bai, H.; Wang, C. Ultrathin W18O49 nanowires with diameters below 1 nm:
Synthesis, near-infrared absorption, photoluminescence, and photochemical reduction of carbon dioxide. Angew. Chem. Int. Ed.
2012, 51, 2395–2399. [CrossRef]
316. Barton, E.E.; Rampulla, D.M.; Bocarsly, A.B. Selective solar-driven reduction of CO2 to methanol using a catalyzed p-GaP based
photoelectrochemical cell. J. Am. Chem. Soc. 2008, 130, 6342–6344. [CrossRef]
317. Liu, Q.; Zhou, Y.; Kou, J.; Chen, X.; Tian, Z.; Gao, J.; Yan, S.; Zou, Z. High-yield synthesis of ultralong and ultrathin Zn2GeO4
nanoribbons toward improved photocatalytic reduction of CO2 into renewable hydrocarbon fuel. J. Am. Chem. Soc. 2010, 132,
14385–14387. [CrossRef] [PubMed]
318. Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.; Zou, Z. High-yield synthesis of ultrathin and uniform
Bi2WO6 square nanoplates benefitting from photocatalytic reduction of CO2 into renewable hydrocarbon fuel under visible light.
ACS Appl. Mater. Interfaces 2011, 3, 3594–3601. [CrossRef] [PubMed]
Atmosphere 2022, 13, 1958 40 of 42

319. Li, X.; Pan, H.; Li, W.; Zhuang, Z. Photocatalytic reduction of CO2 to methane over HNb3O8 nanobelts. Appl. Catal. A Gen. 2012,
413, 103–108.
320. Kočí, K.; Obalová, L.; Matějová, L.; Plachá, D.; Lacný, Z.; Jirkovský, J.; Šolcová, O. Effect of TiO2 particle size on the photocatalytic
reduction of CO2 . Appl. Catal. B Environ. 2009, 89, 494–502. [CrossRef]
321. Jensen, J.; Mikkelsen, M.; Krebs, F.C. Flexible substrates as basis for photocatalytic reduction of carbon dioxide. Sol. Energy Mater.
Sol. Cells 2011, 95, 2949–2958. [CrossRef]
322. Zhang, J.; Ren, M.; Li, X.; Hao, Q.; Chen, H.; Ma, X. Ni-based catalysts prepared for CO2 reforming and decomposition of methane.
Energy Convers. Manag. 2020, 205, 112419. [CrossRef]
323. Freeman, B.C.; Bhown, A.S. Assessment of the technology readiness of post-combustion CO2 capture technologies. Energy
Procedia 2011, 4, 1791–1796. [CrossRef]
324. Roh, K.; Bardow, A.; Bongartz, D.; Burre, J.; Chung, W.; Deutz, S.; Han, D.; Heßelmann, M.; Kohlhaas, Y.; König, A. Early-stage
evaluation of emerging CO2 utilization technologies at low technology readiness levels. Green Chem. 2020, 22, 3842–3859.
[CrossRef]
325. IEA. Energy Technology Perspectives 2020: Special Report on Carbon Capture, Utilisation and Storage; IEA: Paris, France, 2020.
326. Kearns, D.; Liu, H.; Consoli, C. Technology Readiness and Costs of CCS; Global CCS Institute: Melbourne, Australia, 2021.
327. Rafiee, A.; Khalilpour, K.R.; Milani, D.; Panahi, M. Trends in CO2 conversion and utilization: A review from process systems
perspective. J. Environ. Chem. Eng. 2018, 6, 5771–5794. [CrossRef]
328. Zhang, Z.; Pan, S.-Y.; Li, H.; Cai, J.; Olabi, A.G.; Anthony, E.J.; Manovic, V. Recent advances in carbon dioxide utilization. Renew.
Sustain. Energy Rev. 2020, 125, 109799. [CrossRef]
329. Agarwal, A.S.; Rode, E.; Sridhar, N.; Hill, D. Conversion of CO to Value Added Chemicals: Opportunities and Challenges. In
Handbook of Climate Change Mitigation and Adaptation; Springer: Cham, Switzerland, 2017; pp. 2487–2526.
330. Pal, T.K.; De, D.; Bharadwaj, P.K. Metal–organic frameworks for the chemical fixation of CO2 into cyclic carbonates. Coord. Chem.
Rev. 2020, 408, 213173. [CrossRef]
331. Kumaravel, V.; Bartlett, J.; Pillai, S.C. Photoelectrochemical conversion of carbon dioxide (CO2 ) into fuels and value-added
products. ACS Energy Lett. 2020, 5, 486–519. [CrossRef]
332. McGrath, O.J. Biological Conversion of Carbon Dioxide to Value-Added Chemicals; West Virginia University: Morgantown, WV, USA, 2021.
333. Nocito, F.; Dibenedetto, A. Atmospheric CO2 mitigation technologies: Carbon capture utilization and storage. Curr. Opin. Green
Sustain. Chem. 2020, 21, 34–43. [CrossRef]
334. Ayodele, B.V.; Khan, M.R.; Cheng, C.K. Syngas production from CO2 reforming of methane over ceria supported cobalt catalyst:
Effects of reactants partial pressure. J. Nat. Gas Sci. Eng. 2015, 27, 1016–1023. [CrossRef]
335. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernández, J.R.; Ferrari, M.-C.;
Gross, R.; Hallett, J.P. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [CrossRef]
336. MacDowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, A.; Jackson, G.; Adjiman, C.S.; Williams, C.K.; Shah, N.; Fennell, P. An
overview of CO2 capture technologies. Energy Environ. Sci. 2010, 3, 1645–1669. [CrossRef]
337. Styring, P.; Jansen, D.; de Coninick, H.; Reith, H.; Armstrong, K. Carbon Capture and Utilisation in the Green Economy; Centre for
Low Carbon Futures: New York, NY, USA, 2011.
338. Bell, A.; Marks, T. Carbon Management: Implications for R&D in the chemical sciences and technology. In A Workshop Report to
the Chemical Sciences Roundtable; National Academy Press: Washington, DC, USA, 2001.
339. Aresta, M. Carbon Dioxide as Chemical Feedstock; John Wiley & Sons: Hoboken, NJ, USA, 2010.
340. Li, Y.; Markley, B.; Mohan, A.R.; Rodriguez-Santiago, V.; Thompson, D.; Niekerk, D. Utilization of Carbon Dioxide from Coal-fired
Power Plant for the Production of Value-Added Products. In Design Engineering of Energy and Geo-Environmental Systems Course
(EGEE 580); Internal Document for a Course; College of Earth and Mineral Science: University Park, PA, USA, 2006.
341. Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a chemical feedstock: Opportunities and challenges. Dalton Trans. 2007,
28, 2975–2992. [CrossRef]
342. Downturn, A.M. Global Demand for Polycarbonate Growing Again, Says IHS Chemical Report; IHS Online Pressroom: Engelwood,
CO, USA, 2012.
343. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the valorization of exhaust carbon: From CO2 to chemicals, materials, and
fuels. Technological use of CO2 . Chem. Rev. 2014, 114, 1709–1742. [CrossRef]
344. Angunn, E.; Nada, A.; Gaëlle, B.-C. Evaluation of carbon dioxide utilisation concepts: A quick and complete methodology. Energy
Procedia 2014, 63, 8010–8016. [CrossRef]
345. Meylan, F.D.; Moreau, V.; Erkman, S. CO2 utilization in the perspective of industrial ecology, an overview. J. CO2 Util. 2015,
12, 101–108. [CrossRef]
346. Aresta, M. Carbon Dioxide Recovery and Utilization; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2003.
347. Metz, B.; Davidson, O.; de Coninck, H.; Loos, M.; Meyer, L. IPCC Special Report on Carbon Dioxide Capture and Storage; Cambridge
University Press: Cambridge, UK, 2005.
348. Hu, B.; Guild, C.; Suib, S.L. Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products. J.
CO2 Util. 2013, 1, 18–27. [CrossRef]
349. Ahmed, M. Greenhouse gas emissions and climate variability: An overview. In Quantification of Climate Variability, Adaptation and
Mitigation for Agricultural Sustainability; Springer: Cham, Switzerland, 2017; pp. 1–26.
Atmosphere 2022, 13, 1958 41 of 42

350. Fan, M.S.; Abdullah, A.Z.; Bhatia, S. Catalytic technology for carbon dioxide reforming of methane to synthesis gas. ChemCatChem
2009, 1, 192–208. [CrossRef]
351. Song, C. Global challenges and strategies for control, conversion and utilization of CO2 for sustainable development involving
energy, catalysis, adsorption and chemical processing. Catal. Today 2006, 115, 2–32. [CrossRef]
352. Matsubu, J.C.; Yang, V.N.; Christopher, P. Isolated metal active site concentration and stability control catalytic CO2 reduction
selectivity. J. Am. Chem. Soc. 2015, 137, 3076–3084. [CrossRef] [PubMed]
353. Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A.E.; Evans, J.; Senanayake, S.D.; Stacchiola, D.J.; Liu, P.; Hrbek, J.; Sanz, J.F. Highly
active copper-ceria and copper-ceria-titania catalysts for methanol synthesis from CO2 . Science 2014, 345, 546–550. [CrossRef]
354. Laurenczy, G. Hydrogen storage and delivery: The carbon dioxide–formic acid couple. CHIMIA Int. J. Chem. 2011, 65, 663–666.
[CrossRef]
355. Markewitz, P.; Kuckshinrichs, W.; Leitner, W.; Linssen, J.; Zapp, P.; Bongartz, R.; Schreiber, A.; Müller, T.E. Worldwide innovations
in the development of carbon capture technologies and the utilization of CO2 . Energy Environ. Sci. 2012, 5, 7281–7305. [CrossRef]
356. Foley, A. Sustainable Automotive Technologies 2013 Proceedings of the 5th International Conference ICSAT 2013, J. Wellnitz, A.
Subic, R. Trufin. Springer International Publishing, Switzerland (2014). £126 (hardback); £100.50 (eBook), ISBN: 978-3-319-01883-6
(hardback); ISBN: 978-3-319-01884-3 (eBook). J. Transp. Geogr. 2015, 45, 83–84. [CrossRef]
357. Kuld, S.; Conradsen, C.; Moses, P.G.; Chorkendorff, I.; Sehested, J. Quantification of zinc atoms in a surface alloy on copper in an
industrial-type methanol synthesis catalyst. Angew. Chem. 2014, 126, 6051–6055. [CrossRef]
358. Bansode, A.; Urakawa, A. Towards full one-pass conversion of carbon dioxide to methanol and methanol-derived products. J.
Catal. 2014, 309, 66–70. [CrossRef]
359. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H 2 for synthesis of CO, methanol and hydrocarbons: Challenges
and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [CrossRef]
360. Xiao, S.; Zhang, Y.; Gao, P.; Zhong, L.; Li, X.; Zhang, Z.; Wang, H.; Wei, W.; Sun, Y. Highly efficient Cu-based catalysts via
hydrotalcite-like precursors for CO2 hydrogenation to methanol. Catal. Today 2017, 281, 327–336. [CrossRef]
361. Kattel, S.; Yu, W.; Yang, X.; Yan, B.; Huang, Y.; Wan, W.; Liu, P.; Chen, J.G. CO2 Hydrogenation over Oxide-Supported PtCo
Catalysts: The Role of the Oxide Support in Determining the Product Selectivity. Angew. Chem. Int. Ed. 2016, 55, 7968–7973.
[CrossRef]
362. Brown, R.C. Introduction to thermochemical processing of biomass into fuels, chemicals, and power. In Thermochemical Processing
of Biomass: Conversion into Fuels, Chemicals and Power; John Wiley and Sons: Hoboken, NJ, USA, 2011; pp. 1–12.
363. Gangadharan, P.; Kanchi, K.C.; Lou, H.H. Evaluation of the economic and environmental impact of combining dry reforming
with steam reforming of methane. Chem. Eng. Res. Des. 2012, 90, 1956–1968. [CrossRef]
364. De Queiroz, F.A.O.; De Medeiros, J.L.; Alves, R.M.B. CO2 Utilization: A process systems engineering vision. In CO2 Separation
and Valorisation; IntechOpen: Rijeka, Croatia, 2014; pp. 35–88.
365. De Queiroz, F.A.O.; De Medeiros, J.L.; Alves, R.M.B. CO2 Utilization: A Process Systems Engineering Vision. In CO2 Sequestration
and Valorization; Morgado, C.D.R.V., Esteves, V.P.P., Eds.; IntechOpen: Rijeka, Croatia, 2014; pp. 44–90. [CrossRef]
366. Jiang, Z.; Liao, X.; Zhao, Y. Comparative study of the dry reforming of methane on fluidised aerogel and xerogel Ni/Al2O3
catalysts. Appl. Petrochem. Res. 2013, 3, 91–99. [CrossRef]
367. Schulz, L.A.; Kahle, L.C.; Delgado, K.H.; Schunk, S.A.; Jentys, A.; Deutschmann, O.; Lercher, J.A. On the coke deposition in dry
reforming of methane at elevated pressures. Appl. Catal. A Gen. 2015, 504, 599–607. [CrossRef]
368. Kaiser, P.; Unde, R.B.; Kern, C.; Jess, A. Production of liquid hydrocarbons with CO2 as carbon source based on reverse water-gas
shift and Fischer-Tropsch synthesis. Chem. Ing. Technol. 2013, 85, 489–499. [CrossRef]
369. Ross, J.R. Natural gas reforming and CO2 mitigation. Catal. Today 2005, 100, 151–158. [CrossRef]
370. Urlan, F.; Marcu, I.-C.; Sandulescu, I. Oxidative dehydrogenation of n-butane over titanium pyrophosphate catalysts in the
presence of carbon dioxide. Catal. Commun. 2008, 9, 2403–2406. [CrossRef]
371. Liu, H.; Zhang, Z.; Li, H.; Huang, Q. Intrinsic kinetics of oxidative dehydrogenation of propane in the presence of CO2 over
Cr/MSU-1 catalyst. J. Nat. Gas Chem. 2011, 20, 311–317. [CrossRef]
372. Estes, D.P.; Copéret, C. The Role of Proton Transfer in Heterogeneous Transformations of Hydrocarbons. CHIMIA Int. J. Chem.
2015, 69, 321–326. [CrossRef] [PubMed]
373. Cheng, Y.; Zhang, F.; Zhang, Y.; Miao, C.; Hua, W.; Yue, Y.; Gao, Z. Oxidative dehydrogenation of ethane with CO2 over Cr
supported on submicron ZSM-5 zeolite. Chin. J. Catal. 2015, 36, 1242–1248. [CrossRef]
374. Koirala, R.; Buechel, R.; Krumeich, F.; Pratsinis, S.E.; Baiker, A. Oxidative dehydrogenation of ethane with CO2 over flame-made
Ga-loaded TiO2. ACS Catal. 2015, 5, 690–702. [CrossRef]
375. Müller, K.; Baumgärtner, A.; Mokrushina, L.; Arlt, W. Increasing the equilibrium yield of oxidative dehydrogenation with CO2 by
secondary reactions. Chem. Eng. Technol. 2014, 37, 1261–1264. [CrossRef]
376. Wang, S.; Zhu, Z. Catalytic conversion of alkanes to olefins by carbon dioxide oxidative dehydrogenation a review. Energy Fuels
2004, 18, 1126–1139. [CrossRef]
377. Ansari, M.B.; Park, S.-E. Carbon dioxide utilization as a soft oxidant and promoter in catalysis. Energy Environ. Sci. 2012,
5, 9419–9437. [CrossRef]
378. Sun, H.; Zhang, Y.; Guan, S.; Huang, J.; Wu, C. Direct and highly selective conversion of captured CO2 into methane through
integrated carbon capture and utilization over dual functional materials. J. CO2 Util. 2020, 38, 262–272. [CrossRef]
Atmosphere 2022, 13, 1958 42 of 42

379. Mutz, B.; Carvalho, H.W.; Mangold, S.; Kleist, W.; Grunwaldt, J.-D. Methanation of CO2 : Structural response of a Ni-based
catalyst under fluctuating reaction conditions unraveled by operando spectroscopy. J. Catal. 2015, 327, 48–53. [CrossRef]
380. Younas, M.; Loong Kong, L.; Bashir, M.J.; Nadeem, H.; Shehzad, A.; Sethupathi, S. Recent advancements, fundamental challenges,
and opportunities in catalytic methanation of CO2 . Energy Fuels 2016, 30, 8815–8831. [CrossRef]
381. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation–
From fundamentals to current projects. Fuel 2016, 166, 276–296. [CrossRef]
382. Billig, E.; Decker, M.; Benzinger, W.; Ketelsen, F.; Pfeifer, P.; Peters, R.; Stolten, D.; Thrän, D. Non-fossil CO2 recycling—The
technical potential for the present and future utilization for fuels in Germany. J. CO2 Util. 2019, 30, 130–141. [CrossRef]
383. Bailera, M.; Lisbona, P.; Romeo, L.M.; Espatolero, S. Power to Gas projects review: Lab, pilot and demo plants for storing
renewable energy and CO2 . Renew. Sustain. Energy Rev. 2017, 69, 292–312. [CrossRef]
384. Xing, Y.; Ma, Z.; Su, W.; Wang, Q.; Wang, X.; Zhang, H. Analysis of research status of CO2 conversion technology based on
bibliometrics. Catalysts 2020, 10, 370. [CrossRef]
385. Wannakao, S.; Artrith, N.; Limtrakul, J.; Kolpak, A.M. Engineering transition-metal-coated tungsten carbides for efficient and
selective electrochemical reduction of CO2 to methane. ChemSusChem 2015, 8, 2745–2751. [CrossRef]
386. Park, M.; Kwak, B.S.; Jo, S.W.; Kang, M. Effective CH4 production from CO2 photoreduction using TiO2/x mol% Cu–TiO2
double-layered films. Energy Convers. Manag. 2015, 103, 431–438. [CrossRef]
387. Yasin, N.H.M.; Maeda, T.; Hu, A.; Yu, C.-P.; Wood, T.K. CO2 sequestration by methanogens in activated sludge for methane
production. Appl. Energy 2015, 142, 426–434. [CrossRef]
388. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to
hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [CrossRef]
389. Tomsett, A.; Hagiwara, T.; Miyamoto, A.; Inui, T. Highly active catalysts for CO2 methanation to provide the second reactor of
two stage process for high BTU SNG synthesis. Appl. Catal. 1986, 26, 391–394. [CrossRef]
390. Li, Z.; Shi, R.; Ma, Y.; Zhao, J.; Zhang, T. Photodriven CO2 Hydrogenation into Diverse Products: Recent Progress and Perspective.
J. Phys. Chem. Lett. 2022, 13, 5291–5303. [CrossRef]
391. Liu, Z.; Gao, X.; Liu, B.; Song, W.; Ma, Q.; Zhao, T.-S.; Wang, X.; Bae, J.W.; Zhang, X.; Zhang, J. Highly stable and selective layered
Co-Al-O catalysts for low-temperature CO2 methanation. Appl. Catal. B Environ. 2022, 310, 121303. [CrossRef]
392. Herrmann, F.; Grünewald, M.; Meijer, T.; Gardemann, U.; Feierabend, L.; Riese, J. Operating window and flexibility of a lab-scale
methanation plant. Chem. Eng. Sci. 2022, 254, 117632. [CrossRef]
393. Du, J.; Gao, J.; Gu, F.; Zhuang, J.; Lu, B.; Jia, L.; Xu, G.; Liu, Q.; Su, F. A strategy to regenerate coked and sintered Ni/Al2O3
catalyst for methanation reaction. Int. J. Hydrog Energy 2018, 43, 20661–20670. [CrossRef]
394. Burger, T.; Donaubauer, P.; Hinrichsen, O. On the kinetics of the co-methanation of CO and CO2 on a co-precipitated Ni-Al
catalyst. Appl. Catal. B Environ. 2021, 282, 119408. [CrossRef]
395. Kesavan, J.K.; Luisetto, I.; Tuti, S.; Meneghini, C.; Iucci, G.; Battocchio, C.; Mobilio, S.; Casciardi, S.; Sisto, R. Nickel supported on
YSZ: The effect of Ni particle size on the catalytic activity for CO2 methanation. J. CO2 Util. 2018, 23, 200–211. [CrossRef]
396. Schubert, M.; Pokhrel, S.; Thomé, A.; Zielasek, V.; Gesing, T.M.; Roessner, F.; Mädler, L.; Bäumer, M. Highly active Co–Al 2 O
3-based catalysts for CO2 methanation with very low platinum promotion prepared by double flame spray pyrolysis. Catal. Sci.
Technol. 2016, 6, 7449–7460. [CrossRef]
397. Li, Y.; Lu, G.; Ma, J. Highly active and stable nano NiO–MgO catalyst encapsulated by silica with a core–shell structure for CO2
methanation. RSC Adv. 2014, 4, 17420–17428. [CrossRef]

You might also like