Next Article in Journal
Stirring-Assisted In Situ Construction of Highly Dispersed MoS2/g-C3N4 Heterojunctions with Enhanced Edge Exposure for Efficient Photocatalytic Hydrogen Evolution
Previous Article in Journal
Algae to Biofuels: Catalytic Strategies and Sustainable Technologies for Green Energy Conversion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in Post-Combustion Carbon Dioxide Capture, Direct Air Capture, and Utilization

by
Abdullah Akhdhar
1,
Abdullah S. Al-Bogami
1,
Naeem Akhtar
2 and
Waleed A. El-Said
1,*
1
Department of Chemistry, College of Science, University of Jeddah, P.O. Box 80327, Jeddah 21589, Saudi Arabia
2
Institute of Chemical Sciences, Bahauddin Zakariya University (BZU), Multan 60800, Pakistan
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(9), 807; https://doi.org/10.3390/catal15090807 (registering DOI)
Submission received: 20 June 2025 / Revised: 1 August 2025 / Accepted: 17 August 2025 / Published: 25 August 2025
(This article belongs to the Section Environmental Catalysis)

Abstract

Carbon dioxide (CO2) released from natural or even anthropogenic sources may lead to an increase in the average global temperature and ultimately a climate shift. Thus, protecting the environment by reducing CO2 concentration is a global concern. The recent approach of integrating CO2 through capture, utilization, and storage seems to be an effective eradication technique. Even though a wide range of CO2 capture strategies have been successfully adopted, there is an urgent need to compare these based on their advantages and weaknesses to define the development direction for future perspectives. Several materials have been used for capturing CO2. Thus, we have elaborated and compared the current state of CO2 capture strategies, including post-combustion CO2 capture and direct air CO2 capture. Strategies adopted under post-combustion CO2 capture, including liquid- and solid-based sorbents, membrane-based separators, and electrochemical reduction, have been explained in detail, including their advantages, weaknesses, and potential risks. Thus, this review presents a thorough analysis of CO2 capture in terms of material developments and manufacturing techniques, with some research gaps for future development.

Graphical Abstract

1. Introduction

Despite the incipient climate crisis, the world continues to invest in fossil fuels, and by 2030, their use may rise by 2% annually, thus resulting in a temperature rise of 1.5 °C, according to the Production Gap Report 2020 [1]. Climate change patterns and an increase in the average world temperature are both significantly impacted by the growing level of carbon dioxide (CO2) in the atmosphere [2]. The average atmospheric concentration of CO2 in 2020 was over 400 ppm, the highest level in the last 800,000 years [3]. The burning of fossil fuels to meet the world’s energy needs is the primary source of CO2 emissions [4], thus affecting climate change [5]. Significant attempts have been made in recent years by applying various strategies to control CO2 levels even while supplying the growing energy demand on a worldwide scale. CO2 capture is one of the most advanced climate mitigation options for reducing aggregated CO2 concentration in the atmosphere and thereby limiting the detrimental consequences of climate change [6]. In general, CO2 capture systems comprise adsorption, biochemical, and membrane technologies that can efficiently collect CO2 from flue gases (post-combustion), capture CO2 upstream to avoid its release (pre-combustion), or even capture it from the ambient air [7]. Numerous materials have been thoroughly investigated for CO2 collection, including metal–organic frameworks (MOFs), zeolite, carbon, supported solid amines, ionic liquids, and other sorbents [8].
Amine-based solutions have been widely used due to their high CO2 absorption capacities and established technological frameworks [9,10,11]. Monoethanolamine (MEA) is the most widely used amine adsorbent for capturing CO2. Amine-based absorbents have been a staple in CO2 capture for the past five decades [9]. The mechanism is based on chemical absorption. However, amine-based adsorbents often suffer from high energy requirements for regeneration and from potential degradation over time [10].
Another significant challenge is how to handle the captured CO2. We can employ it to extract oil and create useful chemicals, fuels, and polymers; we can clean industrial waste using alkaline solutions; or we can infuse captured CO2 into geological formations and the ocean [12]. Geological storage and CO2 mineralization are the two primary methods for permanently storing the acquired CO2 [13]. In the former method, CO2 is pumped underground to be stored; in the latter method, CO2 reacts with minerals that contain calcium oxide (CaO) or magnesium oxide (MgO) to produce stable carbonates that can hold CO2 over millions of years [14]. CO2 utilization is as crucial for carbon neutrality as for CO2 capture and storage [15]. Numerous technologies have been applied for CO2 utilization, including hydrogenation, electrochemical, bio-electrochemical, and photochemical reduction, CO2 conversion to polymers, and so on [16]. Significant improvements have been made in scientific knowledge and technology development of different CO2 capture, storage, and usage solutions [17]. This review offers a thorough analysis of CO2 capture as well as recent developments in materials and manufacturing techniques. It also covers the benefits, drawbacks, and potential risks associated with various technological options and their techno-economics.

2. CO2 Emissions: Sources and Their Impact on the Environment

CO2 levels in the atmosphere rise due to a variety of CO2 emission sources, mainly divided into two categories: natural and human [18]. Natural causes include decomposition, animal and plant respiration, ocean discharge, and respiratory and volcanic activity. In contrast, the production of cement, clearing of forests, electricity, public transit, manufacturing, and the combustion of fossil fuels are all examples of anthropogenic emission sources [19]. Generally, energy sectors are responsible for one-third of the worldwide CO2 emissions, and approximately 70% of it comes from fossil fuels [20]. Another significant source of CO2 emissions is wildfires, which release an extra 7.3 to 9.5 gigatons (Gt) of CO2 into the atmosphere. It is noteworthy that total CO2 emissions from naturally occurring processes are about 20 times greater than anthropogenic CO2 emissions [21].

3. CO2 Capture Strategies

CO2 capture poses a plethora of challenges due to distinct sources of emission with varied properties, volumes, compositions, temporal or geographical patterns, and technological choices, as well as nanomaterials for CO2 extraction from flue streams [22]. During the Industrial Revolution (1750–2011), the emission of carbon dioxide into the air was about 2040 ± 310 Gt due to the widespread use of fossil fuels and deforestation [23]. According to the American Physical Society, the collection of CO2 will lower its atmospheric level by merely 1 ppmv [21]. To address this issue, it is necessary to create a variety of cutting-edge technologies that could both absorb and prevent CO2 emissions. The development of sophisticated technologies can prevent the CO2 emission to increase the efficiency of fossil fuel utilization and conversion [24]. A wide range of literature reviews is available describing a comprehensive collection of techniques for CO2 capture. For example, Mondal et al. presented contemporary CO2 capture and separation methods from thermal power plant flue gas [25], while Al-Mamoori et al. addressed the most recent achievements in combined CO2 collection and use, as well as the accompanying obstacles [26]. Similarly, to reduce CO2 gas emissions, several technologies have been explored, including pre-combustion, chemical looping, post-combustion, and oxyfuel combustion [27]. Modern technologies for CO2 capture include, but are not limited to, chemical/physical adsorption and absorption, bioremediation, and cryogenic and membrane separation [7]. Briefly, adsorption refers to CO2 molecules adhering to the surface of solid materials (e.g., zeolites, MOFs, activated carbon). In contrast, absorption involves dissolution of CO2 in or chemical binding of CO2 to liquid solvents (e.g., amines, ionic liquids). These mechanisms differ significantly in regeneration energy, operating conditions, and scalability.
The extent of CO2 in the flue stream is another essential concern in CO2 capture. It is predicted that, as the CO2 level in steam drops, the energy and cost burden for CO2 capture rises [28]. Furthermore, the chemical and thermal composition of other elements in the air stream influences the applicability and performance of the capture technique. Additionally, the energy and cost are also dependent upon fuel type, location, volume, CO2 level, and the kind of capture device used in the power plant [25]. Advantages and weaknesses are two major factors on which various applied strategies can be compared, and prospects can be adopted. Thus, various strategies, including post-combustion, sorbents, membrane-based separators, molecular sieves, electrochemical reduction, and direct air capture, have been explained, elaborated on, and compared.

3.1. Post-Combustion Strategy

Post-combustion strategy is a widely recognized technique for reducing CO2 emissions from energy-related sources. The post-combustion process involves removing CO2 from the flue gas produced by burning fossil fuels in the air. Then, the captured CO2 is compressed for transportation and geological storage [29]. This process is used on both new and retrofit boilers, heaters, turbines, and natural gas-fired or coal-fired power plants on a small scale. Challenges of post-combustion capture include the need for extensive equipment, concerns about solvent oxidation and thermal degradation, and health and environmental issues due to solvent losses from the absorber. Several different CO2 collecting systems are used following the combustion of fuel. Experimentally, they differ significantly in how they separate and capture CO2. They generally involve membrane separation, adsorption, mineralization, absorption, adsorption, and electrochemical separation using fuel cells, as shown in Figure 1. According to the evaluation of the possibilities for post-combustion carbon collection and storage, 57% of the options depended on absorption, 14% on adsorption, 8% on membranes, and 21% on mineralization or bio-fixation [30]. The absorption and desorption of CO2 by solvents and liquid sorbents are depicted in Figure 2. Regardless of the kind, CO2 collection and storage significantly increase financial costs.

3.1.1. CO2 Capture via Liquid Sorbents

Solvents were the first materials utilized in retrofitting existing coal-fired plants for active CO2 reduction, and are also a significant subgroup of materials devoted to CO2 capture [31]. Physical absorption or chemical adsorption are both viable options for post-combustion CO2 collection in liquid [32]. Both methods of absorption use a two-step process that depends upon the interfacial equilibrium among the liquid absorbent and flue gases (Figure 3). By adding a selective solvent in an absorptive tower, the product flue gas is stripped of its CO2 content on a preferential basis. The solvent is then recycled and reused in the subsequent CO2 capture procedures. Adsorption of CO2 usually requires a low temperature, while a rise in temperature leads to desorption due to the inverse relationship between temperature and solubility of gases in a fluid [33]. Thus, low temperatures are maintained. Liquid sorbents are divided into the following categories depending on their properties.
Amine-Based Liquid Sorbents
Amines are recognized as some of the most versatile solvents because of their reactive nitrogen atoms [35]. Additionally, they favor selective and reversible CO2 reactions. The two main advantages of amines are their affordability and low steam pressure [36]. Amine-based solvents can be divided into two primary groups: organic and inorganic solvents. The organic group mainly consists of polar solvents using methyl diethanolamine, diethanolamine, triethanolamine, and monoethanolamine [37]. The inorganic group mostly includes sodium and potassium carbonates combined with ammonia, with potassium carbonate being the most commonly used. These solvents are more stable, less costly, and have fewer environmental impacts. Because their contact with CO2 generates less heat compared to primary or secondary amines, they can also lower regeneration energy costs during the desorption process. However, their limited sensitivity to CO2 in flue gas streams restricts widespread use in absorbers. The most extensively researched and used amine is ethanolamine (MEA), considered the standard for all sorbent-based CO2 capture processes [38]. The absorption, compression, and refinement of MEA organic amines are commonly used methods to achieve CO2 capture goals [39]. MEA-based CO2 capture technology has become the most advanced and economically attractive approach. About 20 to 30% of typical CO2 capture systems rely on MEA [40]. Nevertheless, it requires significant energy regeneration due to thermal stripping at higher temperatures. Additionally, issues such as corrosion, high operational costs, solvent degradation, and greater power consumption pose challenges for amine-based CO2 capture. These limitations greatly restrict their widespread application. To address these problems, recent proposals include using amine mixtures and biphasic solvents.
The loading capability of three different CO2 collection setups, including the testing lab, pilot plant, and extended pilot system, was evaluated by utilizing the MDEA aqueous solution (Figure 4) [41]. In recent times, the usage of benzylamine (BZA) and a primary amine, both alone and in combination with other amines, has attracted great interest. Benzylamine (BZA), which offers attractive properties of large-scale production, like minimal corrosivity and strong resistance toward heat and oxidative degradation, can be coupled with ethanolamine (MEA) and can be used as a standard in the capture technique [42]. Similarly, Figure 5 illustrates the MEA-based deep eutectic solvent with an improved CO2 absorption capacity of around 0.2715 g g−1 [43]. Additionally, it has been reported that conventional MEA-based absorption operates at ~40–60 °C with regeneration energy of ~3.5 GJ/ton CO2 and CO2 loading of 0.4–0.5 mol/mol, while solid sorbents like MOFs exhibit higher loading (1–2 mmol/g) but require elevated regeneration temperatures (~150–200 °C).
Additionally, it has been reported that combining two primary amines may increase the rate of CO2 absorption; however, it does not significantly reduce the heat required for sorbent renewal [44]. Moreover, in contrast to secondary and tertiary amines, phase separation processes occur more efficiently in primary amine-based absorbents.
Similarly, the phase separation phenomenon was also observed in a mixture of polyamine/1-propanol/H2O absorbents. Thus, for the creation of novel biphasic solvents, a molecular knowledge of phase separation is essential. Additionally, amine-based CO2 capture involves a complex equilibrium network of reactions between CO2, water, and amine molecules. The reaction pathway depends on the type of amine (primary, secondary, tertiary) and the aqueous medium (Equation (1)). Generally, primary and secondary amines capture CO2 mainly via carbamate formation (Equations (2) and (3)), whereas tertiary amines facilitate bicarbonate formation (Equations (4) and (5)) due to their inability to form carbamates. The mechanism behind the absorption of CO2 at the amine surface can be explained [36].
CO2(g) → CO2(aq)
CO2(aq) + 2RNH2 → RNH3+ + RNHCOO
RNHCOO + H2O → RNH2 + HCO3
CO2 + H2O → HCO3 + H+
R3N + H+ → R3NH+
Furthermore, recently, the effect of multi-amine solvents with high CO2 absorption capacity was also assessed. Briefly, in the experiment, two multi-amine solvents, including 2-(2-aminoethylamino)ethanol (AEEA) and diethylenetriamine (DETA), along with the reference solvent monoethanolamine (MEA), were chosen as three amine solvents [45]. Compared to MEA, DETA was able to absorb 2.962 GJ/ton CO2 while reducing the reboiler duty by 16.8%. Thus, CO2 collection efficiency can be significantly improved with a multi-amine system. Additionally, the CO2 absorption/desorption performance of six novel solvent blends—comprising amine, sulfolane, methanol, diethanolamine (DEA), 2-amino-2-methyl-1-propanol (AMP), diisopropylamine (DPA), MDEA, and ethanolamine (MEA)—was evaluated using ATR-FTIR spectroscopy [46]. Results indicate that MEA, DEA, DIPA, and MDEA produce low or negligible amounts of MMC, while AMP has an intermediate performance [47]. These findings suggest that combining amines can somewhat reduce regeneration energy, but there is still a need to develop new absorbents for efficient CO2 capture. Additionally, piperazine (PZ) and its derivatives are other promising solvents in this category. These compounds outperform MEA in CO2 capture capacity and thermal stability, despite their higher vapor pressure [48].
Similarly, using a pilot-scale carbon capture process, the performance and features of a new polyamine-based solvent blend for post-combustion CO2 capture were examined. The proposed solvent was designed to differentiate the roles of a main amine, an auxiliary amine, and a reaction-rate enhancer. It consists of a mixture of three different amine types. Due to its ability to absorb large amounts of CO2, polyamine 3,3′-iminobis (N, N-dimethylpropylamine) was selected as the primary amine. The auxiliary amine used was 2-amino-2-methyl-1-propanol, while piperazine was included to accelerate the reaction. It was found that the proposed solvent performed steadily and required significantly less reboiler duty than the 30 mass% monoethanolamine (MEA) solvent.
Ammonia-Based Sorbents
Fluid ammonia (NH3) is considered a promising option because of its unique features, including a high capacity to capture CO2, low energy consumption, minimal operational complications, and the ability to remove acidic oxides and produce additional beneficial products [49]. Additionally, it does not suffer from corrosion issues like amine-based sorbents. As a result, aqueous ammonia, or NH3, has been widely used for CO2 capture from industrial gas streams. Both physical and chemical CO2 absorption can occur simultaneously when using an NH3 solution [50]. However, physical absorption appears to be the dominant process, which may depend on the gas-sorbent contact area—the process where CO2 molecules diffuse from the gas phase to the gas–liquid interface before dissolving [51]. Equations (6)–(12) illustrate how NH3 molecules react with dissolved CO2 to form CO2-rich ammonia salts.
CO2 (aq) + H2O ↔ HCO3 + H+
HCO3 ↔ CO32− + H+
NH3 (aq) + H2O ↔ NH4+ + OH
NH3 (aq) + HCO3 ↔ NH2COO + H2O
NH4+ + HCO3 ↔ NH4HCO3 (s)
NH4+ + NH2COO ↔ NH2COONH4 (s)
2NH4+ + CO32− + H2 ↔ (NH4)2CO3 H2O (s)
Despite heat being needed for the breakdown of NH4HCO3 or carbonate during the recycling of the NH3 solution, it has been claimed that the NH3 process substantially achieves higher energy efficiency and uses about 60% less energy, on average, than the MEA scrubbing method [52]. The overall estimated energy demand for the NH3-based process was calculated to be 1147 kJ/kg-CO2, which was found to be approximately 27% of the 4215 kJ/kg of CO2 calculated energy required for the MEA-based process utilizing 30% w/w MEA solution [21].
However, the industry appears to be most concerned about NH3 loss; therefore, the “chilled ammonia process (CAP)” has been created to solve this issue [46]. Ammonium bicarbonate (NH4HCO3) and ammonium carbonate ((NH4)2CO3) are prime products of this method, which involves the reaction of a gas stream with an aqueous solution of ammonia in a wet scrubber [53]. Recently, it has been discovered that the biodiesel byproduct glycerol can be effectively used as an additive for reducing NH3 vaporization. These findings show that increasing the glycerol content from 2.5 to 3.0% leads to a 4.8% increase in the mass transfer coefficient [54]. As a result, adding glycerol to NH3 raises the mass transfer coefficient, lowers NH3’s vapor pressure (acting as a green vapor suppressant), and minimizes NH3 evaporation in the absorption tower.
The addition of Ni(II) and piperazine in NH3 solution was also analyzed to reduce loss during CO2 collection (Figure 6) [34]. Furthermore, it was observed that the application of metal ion additives (Cu (II), Zn (II), and Ni (II)) can effectively decrease NH3 leakage in post-combustion capture [55]. NH3 emission may decrease to 26.2, 9.7, and 9.6% by adding divalent metal ions Ni, Cu, and Zn, respectively, which can also facilitate the CO2 desorption process and prevent NH3 leakage during regeneration. Due to the stronger complexation of metal ions with free NH3 compared to Cu(II), Zn, and Ni(II), good NH3 escape inhibition and substantial CO2 desorption enhancement are observed. Furthermore, aqueous solutions of NH3 and K2CO3 also appear to be viable solvents for CO2 collection from industrial exhaust emissions, with less NH3 loss [56].
According to theoretical and experimental research, complexing ammonia with metal ions significantly lowers the amount of free NH3, which minimizes evaporation loss but also lowers the response to CO2 [57]. Unfortunately, CO2 capture by this method may lead to the generation of an NH3 stream beforehand. It can be due to ammonia’s high energy requirement or high carbon footprint. Therefore, the development of a novel NH3-based energy-efficient system with less NH3 leakage along with high CO2 capture capacity is highly demanded.
Ionic Liquid-Based Sorbents
Ionic liquids (ILs), an evolving class of green solvents, have received a lot of interest recently in the domains of synthesis, catalysis, electrochemistry, biochemistry, and gas separation [58,59]. ILs are molten salts that are liquid under 100 °C and are made completely of organic cations and organic/inorganic anions [60,61]. The comparable solubility of CO2 and conventional ILs was originally described by Blanchard et al. in 1999 [62], by creating a novel method for the collection and isolation of CO2. Since then, several thorough investigations on CO2 capture by ILs have been carried out.
The mechanism of CO2 absorption by ILs is mainly based on physical absorption, except for the carboxylate-based IL [63,64,65]. The physisorption mechanism relies on the van der Waals force or hydrogen bonding between CO2 molecules and IL ions. The weak bonds between the CO2 molecules and IL ions facilitate easier desorption and cycling processes; thus, the CO2 molecules could be discharged by modifying the pressure or temperature, enabling the reuse of the IL. Figure 7 demonstrates the CO2 absorption process by ILs, in which the mixture of gases containing CO2 is inserted into the absorption column containing ILs. Firstly, the CO2 molecules are absorbed by the ILs, and the CO2-poor gas is released into the atmosphere. In the second step, the CO2 is regenerated in the stripper using a higher temperature, pressure swings, or inert gases [66].
Imidazolium-based ILs have attracted far more interest among the wide range of IL compositions due to their extraordinarily high solubility for CO2. The extent of CO2 solubility in ILs is largely determined by the reaction between the anion moiety of the ionic compound and CO2, while the cation plays more of a supporting role [67]. Lewis’s acid–base method is used to define this reaction, with the anion acting as an electron donor and CO2 as an electron acceptor [68].
The differences reported in CO2–cation interactions between ILs with hydrogen and methyl substitutions in the temperature gradient between 10 and 50 °C imply that this variation might have an impact. The maximum CO2 absorption was seen in ionic liquids with the [Tf2N] anion for all temperatures examined. Similar to this, at ambient temperature, a triethylenetetramine (TETA) cation-based IL displays a CO2 adsorption efficiency of 0.96 mol CO2/mol-IL, whereas the addition of 40% water boosted the capacity to 2.04 mol CO2/mol-IL [69]. Furthermore, absorbance ability has been demonstrated to be strongly related to the ionic structure and has followed the order (NO3) > (BF4) > (SO4) for the same cation, with 2(TETAH) and (NO3)2 comprising 40% moisture and exhibiting the greatest CO2 capacity [70]. These ammonia- and imidazolium-based ionic liquids share the same drawbacks in terms of lifespan issues, including high energy costs and huge carbon footprints. The main drawbacks of ILs’ application as a CO2 capture absorbent are their high viscosity and limited gas solubility at low CO2 partial pressures [71]. Furthermore, ILs offer solubility for CO2 capture, which is slightly less than that provided by amines. Thus, to overcome this issue, one option is to functionalize ILs with amine groups that can considerably accelerate the chemisorption process and boost the overall solubility of CO2. However, the synthesis of functionalized ILs requires stringent purification procedures. Thus, it becomes much more costly than the synthesis of ordinary ILs. Recently, it has been reported that the introduction of a surfactant or some other solvents to an IL sharply reduces its viscosity and boosts its efficiency for absorption [72]. As a result, new techniques are being explored that make use of binary or ternary IL mixtures, including IL blends with amines and IL blends with some other solvents such as ethanol, water, etc. Many workable combinations may be created by mixing more expensive, highly viscous ILs with other liquids or cheap amines [73]. The primary goal of the mixture is to benefit from the advantageous traits of the different solvents. Thus, such combinations can have the desirable CO2 capture characteristics of task-specific ILs without the disadvantages of high viscosity [74]. Briefly, adding 30–40 wt.% water to imidazolium-based ILs such as [C2C1im][NTf2] has been shown to reduce viscosity by nearly 50% and double the CO2 uptake capacity. Pinto et al. (2013) reported that the CO2 solubility increased from approximately 0.96 mol CO2/mol IL to 2.04 mol CO2/mol IL at ambient temperature when water was added, which could be attributed to improved mass transfer and reduced diffusional resistance [72]. Ethyl-3-methylimidazolium ethylsulfate ([C2C1im][EtSO4]), which shares the cation with [C2C1im][NTf2], and an IL with longer alkyl substituent chains in its cation ([C4C2im][EtSO4]) have been specifically chosen as ILs containing the ethylsulfate anion [75]. Even though ILs have better qualities, such as incredibly low vapor pressure and excellent selectivity for CO2 compared to other solvents, the price is still 10–20 times higher than the solvents that are readily available in stores. Therefore, the development of ILs with more competitive qualities than other commercial adsorbents is of great interest.

3.1.2. Solid Sorbents for CO2 Capture

Harmony between high gas stream temperature and CO2 capture process operating parameters is particularly appealing for better thermal control and improved efficiency [76]. The gas stream must be cooled down to levels that are suitable for the stability of the chosen liquid sorbent. However, solid sorbents, on the other hand, are typically more chemically stable, more cost-effective than liquid solvents, and consistent with the hot temperatures of flue gases generated by industrial operations and power plants. Therefore, solid sorbents have attracted great interest in capturing CO2. To achieve the best efficiency for CO2 absorption, various materials have been developed depending on the adsorption state (pre-combustion or post-combustion) [77]. In the past, zeolites and activated carbons (ACs) were two of the earliest solid adsorbents used for CO2 extraction. But numerous substances, such as metal-–organic frameworks (MOFs), polymers, and metal oxides, have been recently developed to increase the CO2 adsorption efficiency (Figure 8) [78,79,80]. Common criteria for assessing CO2 uptake on solid adsorbents include the following [81]:
  • High selectivity toward CO2;
  • Moderate renewal requirements;
  • Low cost and quick desorption and adsorption kinetics;
  • High CO2 operational capability.
Amine functionalization has often been used to enhance CO2 adsorption, either through post-synthesis grafting or impregnation during synthesis. Lewis acid/Lewis base interactions strengthen the bond between the adsorbent and CO2 molecules when nitrogen is added to a carbon surface, increasing the electron density and raising the material’s basicity. It has been shown that doping the network with nitrogen or other atoms offers some advantages over amine functionalization, especially in terms of stability and preventing leaching across regeneration cycles. Here, we discuss various types of solid sorbents [82].
Carbon-Based Materials
With several uses in catalysis, adsorption, and storage devices, carbon-based materials make up a significant fraction of high-performing adsorbents [83]. Carbon-based materials have several benefits, including inexpensive precursor materials and synthesis methods, variable pore size and surface area textural properties, selective adsorption, hydrophobicity, high-temperature stability, and simple recovery [84].
Due to their plentiful feedstock, affordable ingredients, large accessible surface area, and strong adsorptive properties, in numerous studies, activated carbons (ACs) have been demonstrated to be viable adsorbents in the carbon capture process [85]. The origin and make up of the raw materials can be linked to the adsorptive qualities and activation procedure of ACs. Numerous precursors of ACs have been identified, including N-rich chitosan, bagasse, rice husk, palm fiber and shell, coconut shell, waste biomass, and porous organic polymers, among many others [84]. These precursors require high-temperature calcination before being activated physically or chemically. Materials made of physically activated carbon have undergone pyrolysis and treatment with a stream of CO2, O2, or steam [86]. However, chemicals like salts, bases, or acids have been used to chemically treat ACs. The majority of the ACs for CO2 collection have been described for use in nitrogen doping or textural property adjustment, optimizing either one or both separately. Briefly, Sethia and Sayari (2015) created highly selective super-adsorbent ACs (CO2 capacity 5.39 mmol/g) for CO2 removal at 25 °C and 1 bar. High adsorption and selectivity are produced by the AC materials’ considerable nitrogen concentration (5.07 mmol/g), high surface area (1317 m2/g), and large pore volume (0.27 cm3/g) [87].
Materials made of ACs have benefits like low susceptibility to moisture, availability, and low cost. Nevertheless, these materials are better suited for high-pressure applications. By synthesizing ordered porous carbons and chemically altering adsorbent surfaces, ACs are created and adjusted to adjust surface areas and pore sizes to optimize low-pressure adsorption (pre-combustion) [83]. The manufacture of ordered mesoporous carbon (OMC) materials employs a variety of techniques. The most effective techniques for achieving high surface areas and narrow pore distributions are the soft template and hard template techniques [88].
The self-assembly of block copolymers, co-condensation, and carbonization are the foundations of the soft template technique. Such a method does away with the requirement to introduce pre-synthesized templates and then remove them. The hard template technique makes use of silica nanostructures as the template, filling them with carbon precursor, carbonizing them, and then removing the template with HF or NaOH [89].
Due to their large specific surface area, uniform pore size distribution, and many micrometer-sized pores, activated carbon fibers (ACFs) have been regarded as potential adsorbents. ACFs have a high adsorption capacity and quick rates of adsorption because of their small pore size dispersion. In contrast to powders and granular materials, ACFs’ fibrous shape makes them easier to handle [90]. For CO2 capture at atmospheric pressure, commercial ACFs modified by chemical activation using KOH have been tested. The optimal concentration of KOH increased the carbon fibers’ surface area while modifying the textural characteristics of ACFs [91]. Studying the impact of nitrogen doping on ACFs revealed that the porous structure of the ACF plays a major role in controlling CO2 uptake [92].
Due to its exceptional chemical and physical characteristics, graphene, an allotropic synthetic carbonaceous substance that is relatively new, has received a lot of interest. Large surface area (up to 2630 m2/g), mechanical strength, chemical stability, and strong thermal conductivity are some of these special qualities. Consequently, graphene has a wide range of uses, such as in supercapacitors, transistors, solar cells, touch screens, DNA sequencing, gas sensors, and the absorption of hazardous compounds [93]. However, because graphene layers are restacked during bulk manufacture, their distinctive features are suppressed. According to Chandra et al. (2012), graphene sheets functionalized with polypyrrole were chemically activated to add nitrogen [94]. Nitrogen-doped graphene (NG) has a significant CO2 adsorption capacity of 4.3 mmol/g, which has been measured under atmospheric pressure and ambient temperature [95]. Additional research on nitrogen-doped reduced graphene was performed by Kemp et al. (2013) employing polyaniline (PANI) to uniformly dope nitrogen in porous graphene structures [96]. The study showed that CO2 adsorption is controlled by the combination of N concentration and pore size distribution. The key benefits of PANI-doped graphene are its excellent selectivity and degree of reversibility in CO2 adsorption [95]. Interestingly, nitrogen doping modifies the electronic structure of graphene by offering sites such as pyridinic nitrogen, pyrrolic nitrogen, and graphitic nitrogen that act as Lewis bases, thus enhancing electrostatic interactions and binding affinity against acidic CO2 molecules and leading to stronger adsorption sites and improving both capacity and selectivity compared to pristine graphene. As a result, current research focuses on precisely controlling the type and distribution of nitrogen species and exploring co-doping strategies to further boost performance. Despite these advances, challenges remain in large-scale synthesis, controlling graphene layers, and maintaining high surface area with structural stability during cyclic adsorption–desorption. Takeuchi et al. (2017)’s investigation of CO2 adsorption on monolayer graphene on a silicon carbide substrate (001) at 30 K revealed a low required recovery energy (25–30 kJ/mol), demonstrating the physisorption of CO2 on monolayer graphene [97].
Silica-Based Materials
Being one of the most prevalent substances in the world, silica can be found in a wide variety of minerals and environmental conditions. With a large surface area and high porosity, silica is easily produced. These qualities can be tuned by making the right adjustments during the synthesis process. Particularly, mesoporous silica, with pores ranging from 2 to 50 nm, is employed in numerous applications, such as molecular sieves, drug delivery, catalytic supports, and energy storage [98]. It is a common practice to create mesoporous silica nanoparticles using the Stöber procedure. Tetraethylorthosilicate and water are combined in an alcoholic solvent to create the precursor, which is subsequently hydrolyzed into tiny particles that condense into a bigger structure [99].
The use of silica as a support for CO2 adsorption devices has demonstrated encouraging efficacy. Silica can be used as a substrate to load with amines to trap and adsorb CO2, since it is an inexpensive, easily produced porous material [100]. The performance of the combined adsorbent is significantly influenced by the type of porous silica utilized for CO2 adsorption. Due to its sporadic distribution of pore sizes and shapes, amorphous silica makes an inadequate support for CO2 adsorption. Functionalizing agents cannot reach all the pore volumes; hence, it is not employed. Not only is a substantially greater portion of the pore volume accessible with ordered mesoporous silica, but also the pore size, shape, and surface area of this material can also be easily modified by changing the synthesis, making it an ideal adsorbent support. The process of creating an amorphous silica framework around an organic template is frequently used to create mesoporous silica [101]. The template is subsequently eliminated by calcination or solvent extraction, leaving behind pores with a predetermined form and size.
Amine groups can be used to functionalize a mesoporous silica substrate and boost CO2 adsorption efficiency. All primary and secondary amines are capable of adsorbing CO2 [102]. The hydroxyl groups on the silica surface of the silica template cause a strong reaction with amines. To unravel the amine chains and distribute them uniformly across the silica framework, the hydroxyl groups form hydrogen bonds with the amines [83]. Selecting the proper amine will have an impact on the material’s ability to adsorb CO2 because different amine types have varied affinities for the gas. Silica loaded with primary or secondary amines will increase the potential loading of CO2 since primary and secondary amines react with CO2 more strongly than tertiary amines [103]. The ability to load CO2 is increased by increasing the loaded amine’s nitrogen weight fraction.
It is typical to create amine-functionalized silica adsorbents in two steps [104]. First, sodium silicate or tetraethylorthosilicate solutions are acidified with an inorganic acid to create porous silica. Second, amines such polyethylenimine (PEI), aminomethylpropanol (AMP), ethanolamine (MEA), and tetraethylenepentamine (TEPA) are supported on the produced porous silica, or amino functional groups are grafted onto it.
Alumina-Based Materials
Aluminum oxide can serve as a model for efficient CO2 adsorption when functionalized. High surface areas, huge pore volumes, customized mesoporosity, great crystallinity, and high thermal stability are all characteristics of alumina materials [105]. The material has strong potential for use as a CO2 capture sorbent if it possesses all the qualities listed above. Boehmite, a predecessor to mesoporous alumina, has demonstrated a significant affinity for CO2 using CO2 temperature-programmed desorption (TPD) [106]. Even greater CO2 affinity would result from functionalizing mesoporous alumina. To improve the ability of alumina to adsorb CO2, Feist et al. (2015) loaded it with La2O3. The surface area of the undoped alumina was not considerably changed by the loading of alumina with La2O3, showing that La2O3 does not obstruct any alumina pores [107].
Zeolite-Based Materials
Zeolite frameworks are networks of interconnected channels or cages that are microporous, having pores of the order of 0.5–1.2 nm. Zeolites are an appealing material to examine CO2 adsorption because of the resultant structure, which is a very stable material with high crystallinity, a high surface area, and strong adsorption sites [108]. Zeolites have been the subject of in-depth research on CO2 adsorption. Zeolites are molecular sieves because of their microporosity, which permits selective adsorption based on their pore size [109]. Only molecules that fit through the zeolite’s pores can be adsorbed. Zeolites can be customized for the adsorption of CO2 based on the molecular sieving principle due to the easily changeable pore size [110]. Zeolites’ basicity is raised by the presence of aluminum atoms in their silicate-based framework. Because aluminum has a lower electronegativity than silicon, this effect is produced [83]. Cations like Li, Na, or K can be added to the zeolite framework to make up for aluminum’s negative framework charges. The basicity of zeolite grows as the electronegativity of the additional cations decreases. The capacity of zeolite to adsorb CO2 increases with its base content. Additionally, electrostatic interactions between the framework’s cations and CO2 cause CO2 to be drawn into the zeolite pores [110].
A phenomenon known as “carbon gating” occurs in some zeolites, allowing CO2 to pass through the framework but no other substances [111]. Extra framework cations in cationic zeolites can cause narrow pores to get blocked, a phenomenon known as “gating” [112]. Large cages are frequently joined together by these apertures. By passing through the cations and going from cage to cage, only CO2 can diffuse through the materials. When attempting to adsorb CO2 from a mixture of gases, the gating effect provides excellent adsorption selectivity for CO2 over other tiny adsorbates, which is very helpful.
Porous Crystalline Solids
Metal–organic frameworks (MOFs), zeolite imidazolate frameworks (ZIFs), and covalent organic frameworks are examples of porous crystalline materials (COFs) with extraordinary tenability [113]. These substances, either through synthesis or computational design, have become more popular in CO2 to expedite the search for the most effective adsorbent using the rational adsorbent material design. Electrostatic interactions in MOFs are governed by the metal site type. The most recently researched MOFs with high CO2 adsorption capability, including HKUST-1 and M-MOF-74 (M = metal, e.g., Mg, Co), have metals like Mg, Zn, and Co in their structure [114]. Open metal sites of different metal ions are necessary for the development of novel structures.
Porous substances provide the opportunity for the development of energy efficiency and cost savings with the post-combustion CO2 collection technique. As a pure, cyclic option, CO2 adsorption on porous surfaces is appealing since it consumes less energy as well as a lower enthalpy for adsorption [115]. Adsorption techniques for CO2 collection and isolation include MOFs, carbon and silicon compounds, zeolites, and COFs [116,117,118,119,120,121,122]. Due to their inexpensive budget, great thermal durability, widespread accessibility, and minimal susceptibility to humidity, ACs have indeed been employed extensively as very effective CO2-trapping devices. In addition to being effective synthetic sorbents for capturing CO2, zeolites are also efficient and effective adsorbents for the elimination of CO2 [123].
Because of their exceptional pore structure, chemical tunability, and thermal stability, MOFs have drawn particular interest in the last two decades [124]. The efficacy of CO2 capture can be improved by adjusting elements, including open metal sites, the use of Lewis bases, polar functional groups to change organic ligands, post-synthesis modification, and pore size management [124,125,126,127]. The most significant cutting-edge techniques recently described for CO2 capture in MOFs are the chemistry of defect engineering, hydrophobicity, topology, breathing in MOFs, and functionalization in MOFs. Plenty of zeolite-templated and MOF-derived carbon compounds, as well as MOFs with strong CO2 capture performance, have been produced and published in recent years. These materials exhibit better CO2 capture capability than their equivalent counterpart materials [128,129,130]. Some GO/MOF (GO = graphene oxide) composites have been synthesized and evaluated for CO2 capture capability. Eight different MOF types have received most of the research attention: the MOF-74 series, CuBTC (also known as HKUST-1), UTSA-16, MOF-505, MOF-5, UiO-66, ZIF-8, and the MIL series, including MIL-101 and MIL-53 [131,132,133,134]. The hybridization of the MOF with the proper quantity of GO can considerably increase the CO2 adsorption capacity at low pressures, according to a general finding from various studies. For instance, a hybrid material based on UiO-66 demonstrated the greatest improvement, attaining a CO2 capacity 70% more than its equivalent pure MOF [135]. GO/CuBTC, with the maximum GO content (i.e., 65% wt.) and under the assumption of no stacking of GO, exhibits the best results in terms of key performance indicators, even though the increase is around 30–45% for GO/MOF materials employing Mg-MOF-74, CuBTC, or MOF-505 [134], with a selectivity for CO2/N2 adsorption of 120 at 313 K, a working capacity of 1.794 mmol/g at a desorption temperature of 443 K, and specific energy consumption of 0.534 GJ/ton-CO2, similar to amine scrubbing. Lin and co-workers report a highly stable and scalable MOF, termed Calgary Framework-20 (CALF-20), with outstanding high-humidity tolerance for long-lasting CO2 collection and direct steam regeneration at the pilot scale. Zinc (II) triazolates are stacked in square-grid layers and pillared by chelating oxalate linkers to create the three-dimensional extended framework known as CALF-20. The CO2 absorption of CALF-20 is 4.07 mmol/g for a v:v = 10:90 gas combination with a realized CO2/N2 IAST (ideal adsorbed solution theory) selectivity of 230 due to its intrinsic microporosity at 1.2 pressure (1 bar = 105 Pa) and 293 K [136]. Primary amine-functionalized IRMOF-74 was also studied for effective CO2 capture under dry and humid conditions (Figure 9) [125].
The IL/MOF composites have strong potential applications in CO2 collection and conversion [137]. On MOF-808 with high porosity, ionic liquid molecule layers (ILMLs) were constructed for gas adsorption. Due to their higher affinity for CO2, ILMLs act as an adhesive layer to capture CO2, while MOF-808 acts as a gas reservoir. At ambient temperatures, the ideal composite adsorbs a significant quantity of CO2 (3.00 mmol g−1), which is 2.6 times more than MOF-808 (1.15 mmol g−1). Notably, the CO2 capacity for each gram of ILMLs was 1000 times more than that of the bulk liquid, bypassing the gas–liquid interface constraint [138]. Selective CO2 capture capabilities were also demonstrated by lanthanide metal–organic frameworks (MOFs) named MOF-590 and -592 based on a benzoimidephenanthroline tetracarboxylate linker (Figure 10). Recently, a combination of MOF and COF was also studied for CO2 capture. The benefits of MOFs and COFs are combined in the MOF/COF hybrid materials, but they also seem to perform much better due to potential synergistic effects at the MOF/COF interface. The CO2 capacity of the reported NH2-UiO-66@Br-COFs hybrid materials was measured to be 169.5 mg/g at 273K and 1.0 bar, outperforming the equivalent single MOF and COF. This hybrid material outperformed the corresponding single MOF and COF. Furthermore, the hybrid materials’ vapor uptake was found to be 3.73 g/g, and it rose as the number of Br-COFs used for coating increased. This creates a new area for successful CO2 adsorption performance by a novel hybrid MOF/COF interface. Despite a lot of research having been conducted on molecular sieve-based CO2 capture, its large-scale application is still limited due to lower adsorption capacity at lower CO2 pressure and stability issues.

3.2. Membranes to Capture CO2

Many industries use membrane technologies for separation processes, including oxygen enrichment, removing CO2 and H2O from natural gas, and purifying H2. Extraction of CO2 through membranes is considered as a promising method for CO2 capture. Additionally, membrane technology offers several advantages, such as low maintenance and energy needs, minimal environmental impact, reliability, and ease of use. However, improving membrane selectivity remains a key challenge for membrane-based systems to compete with other absorption methods. Currently, only a few materials achieve enough CO2 selectivity to possibly reach 100%, although 50% selectivity is necessary to compete with absorption techniques. However, most of them show either low permeability or low selectivity [140].
Polyethylene oxide (PEO), due to its rubbery properties and intrinsic affinity for CO2 through interactions between polar ether groups and CO2, proved to be a potential material to solve this issue. However, PEO is a typical semi-crystalline polymer, with low permeability and high cost [141]. To overcome this challenge, the PEO membrane’s permeability was improved by using crosslinking, copolymerization, and chain branching to break the crystallization and increase chain flexibility. Unfortunately, these polymeric sorbents are not cost-effective for use in CO2 collection. As a potential sorbent for scalable and affordable CO2 collection, trimethylammonium-functionalized polyepichlorohydrin membrane (TPp) was created employing cheap, environmentally friendly raw ingredients and simple synthetic methods [142]. Depending on the source of the emission stream, flue gases often contain varying quantities of water vapor, carbon monoxide, sulfur oxides, nitrogen oxides, hydrogen sulfide, ammonia, oxygen, and argon. These components prevent CO2 separation and have a significant negative effect on the membrane’s lifespan due to plasticization and aging. However, the application of mixed matrix membranes (MMMs) by including certain inorganic elements into the membrane’s polymeric structure enhances the membrane’s mechanical features and helps in addressing the tradeoff issues. In this regard, silica, zeolites, carbon nanotubes, mixed-metal oxides, graphene oxide, carbon molecular sieves, attapulgite, and metal–organic frameworks are the fillers that are most frequently utilized [143].
A set of eight-member-ring zeolite membranes, with pore diameters of 3.6–4.0 Å, was a potential candidate to achieve the separation of CO2/CH4 and CO2/N2. High CO2 selectivity was demonstrated using DDR-type zeolite membranes (3.6 Å × 4.4 Å) over N2 and CH4, and the CO2 permeance was 3.5 × 10−8 mol (m2sPa)−1 [144]. However, the most common porous materials have poor interface compatibility with the polymer matrix, including zeolite and carbon molecular sieves [145]. The inclusion of organic ligands in MOFs improves the interaction with polymers and enables greater inorganic filler loading in the mixed-matrix membrane. This improvement occurs because the organic ligands (e.g., imidazolate or carboxylate groups) on the MOF surface form strong hydrogen-bonding or dipole–dipole interactions with polar functional groups in polymers, thus resulting in enhanced interfacial adhesion and uniform dispersion of fillers [146]. Additionally, this minimizes the interfacial voids that typically reduces the selectivity and enables higher MOF loading without sacrificing mechanical stability. Similarly, La(BTC)(H2O)(DMF)/Matrimid MMMs showed a 183% increase in CO2 permeability and improved selectivity from 34 to 48 compared to pristine membranes, confirming the advantage of organic ligand–polymer compatibility [146].
Therefore, the synthesis of MMMs by the inclusion of MOFs into the polymer to improve durability and resolve the permeability and selectivity has become a recent research interest. In this regard, recently, the P84 polymer matrix was doped uniformly with ZIF-302 nanocrystals to create ZIF-302/P84 MMMs for CO2/N2 separation. The tradeoff between the polymer membrane’s permeability and selectivity was broken, as seen by the greatly improved CO2 gas permeability and CO2/N2 separation factor, which were 5.2 and 46, respectively. Additionally, the long-term operating stability demonstrated that the ZIF-302/P84 MMMs’ CO2/N2 separation performance was maintained for more than 30 h at 3 bar and 60 °C [147].
MMMs created by adding a lanthanide-based MOF called La(BTC)(H2O)(DMF) as a nano-filler to the Matrimid matrix were also studied. The permeability of CO2 increased by 183% for the 30 wt.% loaded membranes compared to the pristine membrane. It has been observed that the selectivity of MMMs increased with increasing nano-filler loading. Briefly, rising from 34.1 to 48.45 for CO2/N2 and from 36.2 to 54.67 for CO2/CH4, the lack of membrane flaws, a better filler/polymer interface, and superior nano-filler dispersion in the polymer matrix were demonstrated [148]. The outcomes demonstrated that these membranes might be utilized further for industrial gas separation applications. However, its non-porous structure limits the further improvement in CO2 permeability. Porosity can be enhanced by the application of an electro-spun fiber composite membrane. These membranes have inadequate CO2 permeability, as fibers are made of an impermeable polymer. To overcome this challenge, recently, continuous and quick transport pathways with low resistance for gas molecules in the PEO matrix were built using highly permeable poly (1-trimethylsilyl-1-propyne) (PTMSP) electro-spun fibers [149]. PEO/PTMSP EFCM not only enhances CO2 permeability and adsorption but also selectivity. Thus, better CO2 permeability and improved selectivity of CO2 over N2 were also observed in ZIF-8@PAN-4/PEO MMM with the same ZIF-8 loading of around 16.6% wt.%; these improvements are boosted by 52.4% and 44.3%, respectively. Due to the low concentrations of CO2 in flue gases, a significant quantity of flue gas must be treated to achieve successful separation, which is the main drawback of this technique. The requirement of high-pressure gradients, the tradeoff between selectivity and permeability, operational issues, and low durability are limiting factors for large-scale applications. Therefore, there is a need to develop cost-effective membranes that overcome the aforementioned factors and also show stability at high temperatures of flue gases.

3.3. Electrochemical Techniques for CO2 Capture

Carboxylation processes can also be carried out via electrochemistry techniques, which have received a lot of attention as “greener” alternatives to chemical ones. Due to their benefits of low-temperature operation, excellent energy economy, and flexible plug-and-play operating mode, electrochemical CO2 capture techniques are receiving extensive attention. Direct redox-active complexation, transition metal-mediated complexation, pH swing, and molten carbonate are major components of the emerging electrochemical CO2 capture technique [150].
Figure 11 shows the basic parameters for pH-swing-based CO2 capture. The application of electrochemical CO2 capture powered by pH swing in carbonate-based capture devices has shown promising results. Proton-coupled electron transfer (PCET) processes were established by Watkins et al. through the creation of a CO2 separation membrane. To absorb and desorb CO2, redox-active carriers that engage in PCET reactions transport H+ from the anode to the cathode. Aziz et al. utilized a pH swing to take advantage of the CO2 hydration/dehydration balance generated by a PCET process. The mechanism of electrochemically recovered amine for CO2 collection, powered by pH swing through a PCET process, is broken down into four phases [151] as shown in Figure 12.
Sharifian et al. [152] and Renfrew et al. [153] review the advanced electrochemical technologies and find that they have acceptable energy consumption at low current densities when run at industrially desired current densities (>50 A/m2) [152]. Electrowinning techniques are used commercially for energy-efficient CO2 collection. Electrochemical CO2 capture systems with ammonia serving as the CO2 capture sorbent and copper serving as the electrochemical medium were studied and achieved a low energy need of 52 kJ/mol CO2 at anodic and cathodic current densities of 470 A/m2 and 2500 A/m2, respectively [154]. This energy performance was quite competitive with the most advanced CO2 capture systems, which typically call for electrical energy of greater than 100 kJ/mol CO2 at greater than 50 A/m2 [150].
A potential electrowinning-coupled CO2 capture (ECC) system employing ammonia (NH3) as a CO2 absorbent and copper (Cu) as an electrochemical mediator was used to develop the electrically driven strategy. To accomplish effective electrochemical CO2 desorption and NH3 regeneration at the desired energy consumption, the ECC method utilized the electro-active competitor of Cu(NH3)42+/Cu. Using a microkinetics process model, energy needs of 47.1–50.2 kJ/mol CO2 were attained [155]. An electrochemical CO2-capture cell with an ideal derivative (7,8-dihydroxyphenazine-2-sulfonic acid, abbreviated as DHPS) was also reported, which shows a superior low-electrolysis energy consumption of 0.49 GJ per ton of CO2 and an average current efficiency of 95.8% at 10 mA cm2 [156].
According to Jens et al., analysis costs can be reduced by about 46% by combining CO2 capture with utilization. Combined CO2 collection and conversion utilizing an electrolyte solution that includes both chemical and physical CO2 absorption solvents has been reported. This highlights the advantage of reducing constant CO2 feeding as a source of free CO2. In further detail, it involves converting CO2 into formate, glycolate, oxalate, and carbon monoxide in a solution of 2-amino-2-methyl-1-propanol (AMP) and propylene carbonate (PC) [157]. The suggested integrated CO2 capture and conversion system entails three primary stages: CO2 is first absorbed in the chemical capture solvent, then it is released within the electrolyzer, and finally, it is electrochemically converted on the spot. On the cathode, the freed CO2 is electrochemically changed into formate, carbon monoxide, or oxalate [158]. Although they have many advantages, electrochemical CO2 capture systems are still in the early stages of research and need a lot of work to become widely used.

3.4. Direct Capture of CO2

Lackner first proposed the idea of direct air capture (DAC) of CO2 in 1999 [159]. The development of DAC is a new method for tackling climate change that could lead to net-negative emissions by reducing atmospheric CO2 levels [160]. Sanz-Pérez et al. [159], Shi et al. [161], and Kelemen et al. [14] have reviewed the literature on DAC, which can involve various technologies such as mineral carbonation, absorption, and adsorption on membranes. Additionally, Adamu et al. [162] have reported photocatalysis technology and cryogenic separation, respectively. Eisaman et al. [163] and Sabatino et al. [164] have explored other approaches, like electrochemical methods or electrodialysis. Currently, the most developed and well-researched technologies are absorption and adsorption. Because of DAC’s potential to remove CO2 from the atmosphere, many review papers have been published about the sorbents used and developed for CO2 capture [165,166,167]. Each of these methods involves two steps: first, capturing CO2 with an air contactor; second, recycling the liquid or solid sorbent, including molecular sieves. The Clime-Works company in Hinwil, Switzerland, developed the first commercial DAC system based on adsorption. The organization also has pilot projects in Italy (Troia), where CO2 is used to produce methane; in Iceland (Hellisheidi), where CO2 is captured and stored for mineralization; and in Germany (Dresden), where CO2 is captured and used to generate energy resources. Other companies, such as Global Thermostat, Antecy, Hydrocell, and Skytree, also employ adsorption methods for removing CO2 from the air. A pilot plant based on absorption is operated by Carbon Engineering in Squamish, Canada [168].
Numerous technological and financial obstacles must be overcome by DAC technology for potential application. First, since atmospheric CO2 levels are very low (currently at 415 ppm), DAC sorbents need to capture CO2 firmly and to be selectively compared to other airborne components. Second, to impact the climate, DAC technologies must be deployed at a massive scale, potentially capturing up to 10 Gt of CO2 annually. DAC sorbents must be produced in large quantities (millions of tons per year), be chemically robust, and cost less than USD 10 per kilogram. Currently, only a few CO2 sorbents meet these strict criteria, most of which are based on alkaline media (such as NaOH or KOH), solid-supported amines, or moisture-swing anion-exchange resins [169]. Crystalline organic materials, such as hydrogen-bonded frameworks and metal–organic frameworks, offer a distinct and promising approach for DAC because of their well-defined, ordered structures. In preliminary tests, HKUST-1, Mg-MOF-74/Mg-dobdc, zeolite-13X, SIFSIX-3-Ni, and TEPA-SBA-15 were evaluated for their ability to extract CO2 under DAC conditions. These sorbents included HUMs, MOFs, zeolites, and amine-modified mesoporous silicates, representing four different types of porous materials. Under DAC conditions, temperature-programmed desorption (TPD) experiments showed that SIFSIX-3-Ni performed the best. Notably, the excellent behavior of SIFSIX-3-Ni HUM compared to other MOF sorbents suggests that further improvements in pore size and functionality could enhance DAC performance of this group of nanomaterials [170]. A precursor MOF suitable for multiple cycles of CO2 adsorption and release through chemical adsorption in actual DAC scenarios was created by functionalizing Mg-MOF-74 with ethylenediamine. Testing with simulated dry air indicated a significant CO2 adsorption capacity of 2.83 mmol/g at 0.39 mbar and 25 °C [171]. Additionally, these materials display a lower CO2 capture capacity that can be increased through amine functionalization [172]. Extensive research has also explored how different types of amines and their structures influence DAC performance [172]. Composites may also be functionalized without amines by adding Zn-OH groups to the pores, as illustrated in Figure 13. It has been demonstrated that removing CO2 from highly concentrated sources, such as fossil fuel combustion gases, can be effectively achieved using polymerized fiber contactors. Using PEI-functionalized polymer/silica fiber sorbents for CO2 collection, as shown in Figure 14, shows promising results.
Furthermore, another good substitute for MOFs is hydrogen-bonded frameworks (HBFs) [173]. Using the concepts of crystal engineering, the structures of HBFs may be accurately built and modified. As a different strategy, CO2 could be captured in hydrophilic media, provided that the absorption process results in the development of stable, solubilized HBF particles that contain CO2. In these situations, it is possible to recycle the water-soluble HBFs by first desorbing the CO2 from the separated particles, followed by filtering to separate the HBF-CO2 crystals from the mixture. A straightforward amine system, m-xylylenediamine (MXDA), provided DAC by crystallization of HBFs, providing evidence. MXDA-CO2 insoluble crystals were created when a solution of MXDA was exposed to the air and collected ambient CO2. This approach effectively blends the features of solid sorbent materials, including structural functionalization and reduced renewal costs, with the merits of liquid media, such as quick kinetics of CO2 capture, simple scaling, and relatively inexpensive [174,175].
The main disadvantage of this technique for large-scale applications is that most sophisticated sorbents have only been tested in a limited range of moisture and temperature conditions (>20 °C). The average annual temperature around the world ranges from minus thirty to fifty degrees, depending on the region, and almost nowhere do absolute humidity levels reach zero [176]. It will be quite challenging to utilize DAC widely because there are not many studies on it in humid situations and at sub-ambient temperatures (like 20 °C). Utilizing the minute temperature fluctuations made available by this weak chemisorption feature can result in significant energy savings. This research suggests that, to prepare DAC materials for prospective deployment in polar and temperate regions, considerable research on DAC materials that perform at low, sub-ambient temperatures is required. Table 1 provides a comparative overview of major CO2 capture approaches, summarizing capture efficiency, energy demand, cost, maturity (TRL), and key insights.

3.5. Chemical Looping Combustion

Although chemical looping combustion (CLC) is classified among oxyfuel combustion technologies, it can be regarded as a distinct technology [182]. Although the term “chemical looping combustion” was first introduced in 1987 by Ishida, Zheng, et al. [183], the CLC concept was patented by [184], and subsequently utilized to improve the combustion efficiency of power plants by [185]. The CLC process involves using solid oxygen instead of gaseous oxygen for the combustion reaction. The system comprises two reactors: (i) a fuel reactor where fuel is burned, and (ii) an air reactor where the oxygen carrier is reoxidized (Figure 15). CLC can be applied to gaseous, liquid, or solid fuels. It employs a solid oxygen carrier material (metal oxide) that fully oxidizes fuel to CO2 and H2O in the fuel reactor by selectively transporting oxygen from the air reactor to the fuel reactor. The metal oxide (MexOy) leaves the fuel reactor in a reduced state and is reoxidized again in the air reactor. Transition metal oxides (such as copper, cobalt, iron, manganese, and nickel) are promising oxygen carrier candidates [186,187]. Fe2O3 and Mn2O3 possess among the highest theoretical oxygen capacities, providing the highest theoretical ratio of oxygen moles per mole of metal. Recently, various CLC technologies have been developed, including (i) Chemical Looping Air Separation (CLAS) for O2 separation from air using metal oxide decomposition [188], (ii) Chemical Looping Reforming (CLR) for H2 production from natural gas [189], (iii) Chemical Looping Gasification (CLG) to convert biomass into syngas [190], (iv) Chemical Looping Oxygen Uncoupling (CLOU) for combusting solid fuels via metal oxide decomposition [191], (v) Self-sufficient Chemical Looping Reforming of Glycerol (CLRG) for H2 production rather than heat [192], and (vi) Syngas Chemical Looping Gasification (SCL) for producing H2 and electricity from coal-derived syngas [193]. The CLC technology offers several advantages, including (i) high efficiency [194], (ii) eliminating direct interaction between exhaust gases and combustion air [195], and (iii) avoiding dilution with N2, which can reduce costs and minimize NOx formation. However, because the reduction reaction of metal oxides is typically endothermic, it may increase overall costs due to the additional energy required.

3.6. Calcium Looping Combustion

Calcium looping (CaL), also known as regenerative carbon cycle (RCC), is one of the highly efficient post-combustion CO2 capture technologies used for capturing CO2 from flue gases in various industrial applications, such as cement plants. The CaL process stands out as a promising technology for CO2 capture, consisting of two main phases: carbonation and calcination. In this process, CaO-based materials serve as CO2 sorbents, which add value due to their cost-effectiveness and availability. CaL capture relies on two reversible chemical reactions [196]: (i) carbonation, where CaO-based materials react with CO2, and (ii) calcination, where CaCO3 decomposes at high temperatures (850–1000 °C) to regenerate CaO and release CO2 for utilization or storage [197]. The CaO can be reused in the carbonation process. The primary challenge of CaL technology is providing the heat necessary for the calcination process. To address the energy demands of calcination and improve capture efficiency, the CaL process has been integrated with other methods such as solar-assisted CaL, oxy-combustion, steam-assisted CaL, and CLC-assisted CaL [198,199,200,201]. Power plants, cement production facilities, and steel manufacturing plants that incorporate CaL technology benefit from several advantages in CO2 capture, including (i) directing flue gases to the CaL process, and (ii) utilizing the surplus waste heat generated during combustion, boiler operation, cooling systems, blast furnace activity, clinker production, and kiln operations for both carbonation and calcination stages. Consequently, these integrations enable CO2 capture with minimal modifications to existing combustion systems, making it a practical retrofit option for conventional industrial plants and playing a vital role in advancing a low-carbon energy future [202,203,204].

4. CO2 Utilization

The research establishment is averse to carbon dioxide capture and storage (CCS) due to issues with long-term responsibility, a lack of adequate storage capacity, the possibility of leaks, and limited support for coastal storage sites. Deep ocean storage would have an immediate effect on reducing pH, increasing water acidity, and possibly creating environmental imbalance. Implementation of sustainable development and carbon capture and utilization (CCU) is gaining popularity as a feasible alternative for CCS and is considered a long-term remedy to the CO2-related challenges. CCU involves the utilization of CO2 for purposes besides storage [205]. For applications that still need carbon sources, CO2 is a crucial raw material because it contributes to product structure and functionality. Utilizing carbon-free energy sources like electricity or hydrogen would be further justification. CO2 is one of the few carbon substitutes for fossil fuels. In recent years, CO2 utilization efforts have included chemical raw materials, coolants, stain removers, solvent media, injection tools, and compressed gas [206].
Among the CO2 utilization technologies, the conversion of captured CO2 into value-added products is attaining great attention [207]. CCU technologies consider CO2 as a raw material for producing numerous products such as methanol, methane, polymers, carbon fibers, fuels, building materials, etc. CO2 conversion into fuels and chemicals is achieved through several processes, e.g., chemical, biological, or electrochemical processes [208,209]. In chemical conversion, catalysts play a central role in the chemical conversion of CO2 into a value-added chemicals. Several catalysts, such as CuO/Al2O3, ZnO/Al2O3, Ru-PNP, Cu/ZnO/Al2O3, Ni/Al2O3, and Ni/CaAl2O4, were widely used in the CO2 chemical conversion. Briefly, methanol and formic acid are major examples of the chemicals produced through the chemical conversion of CO2, which could be used as a fuel [208]. In the case of biological conversion, microorganisms or enzymes were used to convert CO2 into biofuels (i.e., ethanol and butanol) or bioproducts (e.g., bioplastics and proteins) [210]. However, in the case of electrochemical conversion technologies, electric current is applied for the CO2 reduction to form several valuable end products, e.g., formate, ethylene, and methane [211]. Furthermore, CO2 mineralization (carbonation) is another CO2 conversion process, in which CO2 reacts with naturally occurring minerals (e.g., magnesium silicates and calcium silicates) to form stable carbonate compounds [212]. According to Figure 16, CO2 utilization can be classified into two categories including conversion and non-conversion. In CCU, the first step is capturing and separating CO2 at the source. The second step is its use, either directly (as a solvent, operating fluid, or heat-transfer medium) or through conversion into hydrocarbons, polymers, and chemicals [24].

4.1. Direct Utilization of CO2

Numerous products, including dry ice, carbonated beverages, firefighting equipment, solvents, industrial lubricants, welding media, and algal farms for photosynthesis, can directly employ CO2 [213,214]. Additionally, it may potentially improve geothermal systems and oil and gas recovery in large-scale operations. Additionally, one of the direct uses of CO2 is as a heat transfer fluid in many sectors, including the refrigerant and energy sectors [215]. Desalination is another application of CO2 that does not involve chemical conversion. In this process, CO2 and brine are mixed at high temperatures and pressures to generate hydrates that might subsequently be removed to produce potable water. In these applications, CO2 molecules do not interact or further split; they remain intact or dissolved in a solution [216,217]. These direct CO2 operations, however, are small-scale and have little impact on total CO2 reduction.

4.2. Utilization of CO2 by Conversion

The transformation of CO2 into efficient and environmentally friendly solar fuel offers a potential remedy for environmental problems, including climate change and the depletion of fossil fuels. Practical ways to reduce CO2 emissions are provided by technologies including CO2 capture, storage, and use. Although CO2 can be efficiently separated by CO2 collection and storage, it is more feasible to transform CO2 into fuels and chemicals. CO2 can be converted into important chemicals like carbonate, alcohol, salicylic acid, and other compounds [205,218]. The establishment of a C-C or C-H bond is necessary for the conversion process. Such bond creation requires energy input through the process of activating CO2 or other substrates, like reductive coupling. In contrast to C-H and C-C chemical bonds, the carbon–oxygen double bond is extremely stable. Chemical fixation of CO2 often requires significant energy input to carry out substrate activation reactions, as well as high reaction temperatures and pressures [219,220]. There are several methods for converting CO2. In a nutshell, there are two groups of CO2 conversion processes. One category involves processes that require little to no external energy, and the CO2 molecule, therefore, joins the other chemicals, which causes the reaction to take place. Typically, such operations are referred to as carboxylation processes. The resultant generation of lactones, ureas, carboxylates, isocyanates, and carbonates is included in this category [221,222]. The other conversion category involves processes that generate reduced CO2 and demand a large quantity of external energy. To break the bonds in CO2, the additional energy required for these processes is provided by temperature, electrons, or radioactive materials [223]. These methods are referred to as thermal, electrochemical, and photochemical reactions, respectively [224]. Therefore, decreasing the potential barrier to the production of C1 building block compounds requires a catalyst as well as extreme-pressure and extreme-temperature conditions. Chemicals, including urea, polymer building blocks, methanol, formic acid, methane, formaldehyde, dimethyl carbonate, salicylic acid, ethylene carbonate, cyclic carbonates, and polycarbonates, are produced during this catalyzed reaction of CO2 [222,225].

4.3. Enhanced Oil Recovery

The most cost-effective carbon capture, utilization, and storage (CCUS) method is the injection of CO2 into petroleum reservoirs to boost oil recovery (often known as CO2-EOR), as shown in Figure 17. Previously, CO2-EOR operations were intended to maximize the extraction of oil, with CO2 injection viewed as an expense [226,227]. As a result, various steps were taken to reduce the quantity of CO2 stored. The co-optimization of oil extraction and CO2 storage is necessary for CO2-EOR operations; however, given the increasing worldwide focus on carbon capture and sequestration, the displacement and sweeping capabilities of the inserted CO2 are crucial for overall oil recovery and CO2 storage [228]. CO2 displacement efficiency is excellent because of its relatively high solubility and minimal miscibility pressure with strong oil-expanding elements. However, because of its rapid mobility, gravitational separation, reservoir variability, and limited sweep efficiency, a significant amount of petroleum remains inside the reservoir, which restricts oil recovery and CO2 storage [229]. CO2 sweeping performance could be increased by introducing water and CO2 either simultaneously (SWAG injection) or sequentially (WAG injection) [210]. Oil recovery and CO2 storage are enhanced by water-saturated CO2 injection, as the presence of water reduces CO2 mobility. In this process, water constitutes just under 1% of the fluid. Compared to conventional CO2 injection, oil recovery increases by 7%, while an additional 1.14 megatons of CO2 is stored [230].
CO2 foaming and direct CO2 thickening agents can be used to restrict CO2 mobility. Several studies have been conducted on the impact of foam on the CO2 storage potential [231]. An uninterrupted thin fluid sheet known as lamella separates the discontinuous gas phase that makes up foam. To analyze foam production and degradation, concentrated CO2 was pumped into a lengthy sandstone core that had first been saturated with each surfactant. In comparison to the control study without a surfactant, the formation of foam and lowered water/CO2 surface forces during CO2–foam floods resulted in lower residual water saturations. Along with the better volumetric sweep anticipated by foam, this enhancement in microscopic displacement reveals a greater capability for sequestering CO2 than previously thought [232].
Figure 17. A general process flow diagram for EOR-CO2. Reprinted with permission from [233]. Copyright@2016, American Chemical Society.
Figure 17. A general process flow diagram for EOR-CO2. Reprinted with permission from [233]. Copyright@2016, American Chemical Society.
Catalysts 15 00807 g017

5. Analysis and Outlooks

CO2 capture undoubtedly faces a wide variety of obstacles across all frontiers, including difficulties with law and regulation, finances and the economy, the atmosphere and climate, human well-being and quality of life, social and socioeconomic imbalances, and international treaties despite being an accurate and efficient preventive measure. Technically speaking, the variety of methods, procedures, and composites used in CO2 capture pathways present significant obstacles, research gaps, and problems that must be resolved via basic identification and experimentation combined with vital technological advances. Even though many of these problems are addressed and explained in their appropriate parts, Table 1 briefly summarizes a few of the most obvious difficulties and opportunities.
With a basic understanding of CO2–sorbent engagement and management of surface binding processes that govern selective absorption of CO2 from combustion products, research into novel nanomaterials with enhanced sorption capacities should progress more rapidly. Innovative fluid and solid sorbents are needed to increase CO2 absorption capability while reducing environmental impacts. Although DAC consumes a lot of energy and is unquestionably the most expensive solution, it is nonetheless crucial because it completely removes ambient CO2 regardless of the type or source of release. The advancements made in post-combustion trapping solvents and nanomaterials will be extremely beneficial to DAC. By creating catalysts that are widely available on Earth and comprehending the basic mechanics for CO2 surface reduction, this precious carbon source may be utilized as well as recycled to create valuable items. It is now a costly prospect because the cost of CO2 cleanup is significantly higher than most cap-and-trade rates or carbon pricing. Although this inconsistency is exacerbated by legislation, it does not provide businesses, governments, and the corporate sector with the proper incentives to engage in the development of technological solutions. Instead, it forces sectors to opt for the least expensive production and pollution methods. This hole must be closed.
From Table 2, it is evident that ammonia-based technologies demonstrate lower energy consumption, while ionic liquids exhibit exceptionally high CO2 solubility and minimal regeneration energy requirements.

6. Conclusions

Mitigating rising atmospheric CO2 levels demands urgent deployment of capture, separation, and utilization technologies in a real sense. This review has analyzed the major strategies, including solvent-based absorption, solid sorbents, membranes, and DAC, and explains their mechanisms, performance metrics, and key materials. Solvent-based amines remain the most mature, with capture efficiencies of ~0.4–0.5 mol CO2 per mol amine but high regeneration energy (3–4 GJ/ton CO2). On the other hand, solid sorbents such as MOFs and activated carbons offer higher capacities (1–2 mmol/g) and structural tunability but remain limited by thermal stability and scalability. Similarly, membrane-based approaches provide modular, low-temperature operation and integration potential, while DAC offers a complementary pathway with structural–functional advantages but requires major cost reductions to approach the USD 100/ton CO2 target. A key perspective involves technology readiness levels (TRLs) and real-world deployment. Mature capture processes like calcium looping (CaL) and chemical looping combustion (CLC) have achieved TRL 6 and are supported by operational pilot and demonstration plants, bridging laboratory research and industrial application. In contrast, next-generation materials (e.g., MOFs, ionic liquids) are at TRL 2–3, where scalability, environmental impact, and cost remain unvalidated. Bridging this gap requires not only material innovation but also integration with process engineering, pilot-scale testing, and techno-economic analysis to accelerate the transition from bench to field.
Despite substantial progress, several challenges remain, including degradation under real flue gas conditions, a lack of long-term durability, and poorly characterized environmental footprints. Energy demands for regeneration and system complexity also hinder scale-up. Future research should prioritize AI-assisted design of low-energy, high-capacity sorbents, hybrid capture systems combining the strengths of solids, liquids, and membranes, and integration with renewable energy and CO2 conversion technologies to create circular carbon systems to guide policy and deployment. In conclusion, advancing CO2 capture technologies requires a dual focus, such as pioneering new materials and simultaneously maturing processes toward industrial-scale readiness. To achieve this, coordinated efforts between academia, industry, and policymakers will be crucial for achieving carbon neutrality and mitigating the impacts of climate change.

Author Contributions

All authors contributed to the draft of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia, grant number MoE-IF-UJ-R2-22-04100454-2.

Data Availability Statement

No datasets were generated during the current study.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia, for funding this research work through the project number MoE-IF-UJ-R2-22-04100454-2.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Erickson, P.; Lazarus, M.; Verkuijl, C.; Yehle, E. The Production Gap Report: The Discrepancy Between Countries’ Planned Fossil Fuel Production and Global Production Levels Consistent with Limiting Warming to 1.5 °C or 2 °C: 2020 Special Report. 2020. Available online: https://www.unep.org/resources/production-gap-report-2023 (accessed on 1 September 2020).
  2. Wuebbles, D.J.; Fahey, D.W.; Hibbard, K.A. Climate Science Special Report: Fourth National Climate Assessment; US Global Change Research Program: Washington, DC, USA, 2017; Volume I. [Google Scholar]
  3. Semenova, G. Global Environmental Problems in the World. In E3S Web of Conferences; EDP Sciences: Les Ulis, France, 2020. Available online: https://www.nrc.gov/docs/ML1900/ML19008A410.pdf (accessed on 14 December 2020).
  4. Consensus, S. Earth’s Climate is Warming; National Aeronautics and Space Administration (NASA): Washington, DC, USA, 2015. Available online: https://climate.nasa.gov/scientific-consensus (accessed on 20 June 2016).
  5. Wang, R.; Mirza, N.; Vasbieva, D.G.; Abbas, Q.; Xiong, D. The nexus of carbon emissions, financial development, renewable energy consumption, and technological innovation: What should be the priorities in light of COP 21 Agreements? J. Environ. Manag. 2020, 271, 111027. [Google Scholar] [CrossRef]
  6. Akimoto, K.; Sano, F.; Oda, J.; Kanaboshi, H.; Nakano, Y. Climate change mitigation measures for global net-zero emissions and the roles of CO2 capture and utilization and direct air capture. Energy Clim. Change 2021, 2, 100057. [Google Scholar] [CrossRef]
  7. Mukherjee, A.; Okolie, J.A.; Abdelrasoul, A.; Niu, C.; Dalai, A.K. Review of post-combustion carbon dioxide capture technologies using activated carbon. J. Environ. Sci. 2019, 83, 46–63. [Google Scholar] [CrossRef]
  8. Jelmy, E.J.; Thomas, N.; Mathew, D.T.; Louis, J.; Padmanabhan, N.T.; Kumaravel, V.; John, H.; Pillai, S.C. Impact of structure, doping and defect-engineering in 2D materials on CO2 capture and conversion. React. Chem. Eng. 2021, 6, 1701–1738. [Google Scholar] [CrossRef]
  9. Li, S.W.; Jiang, X.; Yang, Q.; Shao, L. Effects of amino functionalized polyhedral oligomeric silsesquioxanes on cross-linked poly(ethylene oxide) membranes for highly-efficient CO2 separation Chem. Eng. Res. Des. 2017, 122, 280–288. [Google Scholar] [CrossRef]
  10. Rana, A.; Andino, J.M. A Review of Materials for Carbon Dioxide Capture. Catalysts 2025, 15, 273. [Google Scholar] [CrossRef]
  11. Dziejarski, B.; Serafin, J.; Andersson, K.; Krzyżyńska, R. CO2 capture materials: A review of current trends and future challenges. Mater. Today Sustain. 2023, 24, 100483. [Google Scholar] [CrossRef]
  12. Tebbiche, I.; Mocellin, J.; Huong, L.T.; Pasquier, L.-C. 27—Circular Economy and Carbon Capture, Utilization, and Storage. In Biomass, Biofuels, Biochemicals; Pandey, A., Tyagi, R.D., Varjani, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 813–851. [Google Scholar]
  13. Raza, A.; Glatz, G.; Gholami, R.; Mahmoud, M.; Alafnan, S. Carbon mineralization and geological storage of CO2 in basalt: Mechanisms and technical challenges. Earth-Sci. Rev. 2022, 229, 104036. [Google Scholar] [CrossRef]
  14. Kelemen, P.B.; McQueen, N.; Wilcox, J.; Renforth, P.; Dipple, G.; Vankeuren, A.P. Engineered carbon mineralization in ultramafic rocks for CO2 removal from air: Review and new insights. Chem. Geol. 2020, 550, 119628. [Google Scholar] [CrossRef]
  15. Jiang, K.; Ashworth, P.; Zhang, S.; Liang, X.; Sun, Y.; Angus, D. China’s carbon capture, utilization and storage (CCUS) policy: A critical review. Renew. Sustain. Energy Rev. 2020, 119, 109601. [Google Scholar] [CrossRef]
  16. Nwabueze, Q.A.; Leggett, S. Advancements in the Application of CO2 Capture and Utilization Technologies—A Comprehensive Review. Fuels 2024, 5, 508–532. [Google Scholar] [CrossRef]
  17. Chen, S.; Liu, J.; Zhang, Q.; Teng, F.; McLellan, B.C. A critical review on deployment planning and risk analysis of carbon capture, utilization, and storage (CCUS) toward carbon neutrality. Renew. Sustain. Energy Rev. 2022, 167, 112537. [Google Scholar] [CrossRef]
  18. Rehman, A.; Ma, H.; Ozturk, I.; Murshed, M.; Dagar, V. The dynamic impacts of CO2 emissions from different sources on Pakistan’s economic progress: A roadmap to sustainable development. Environ. Dev. Sustain. 2021, 23, 17857–17880. [Google Scholar] [CrossRef]
  19. Yoro, K.O.; Daramola, M.O. Chapter 1—CO2 emission sources, greenhouse gases, and the global warming effect. In Advances in Carbon Capture; Rahimpour, M.R., Farsi, M., Makarem, M.A., Eds.; Woodhead Publishing: Sawston, UK, 2020; pp. 3–28. [Google Scholar]
  20. Liu, Z.; Ciais, P.; Deng, Z.; Davis, S.J.; Zheng, B.; Wang, Y.; Cui, D.; Zhu, B.; Dou, X.; Ke, P.; et al. Carbon Monitor, a near-real-time daily dataset of global CO2 emission from fossil fuel and cement production. Sci. Data 2020, 7, 392. [Google Scholar] [CrossRef]
  21. Gür, T.M. Carbon Dioxide Emissions, Capture, Storage and Utilization: Review of Materials, Processes and Technologies. Prog. Energy Combust. Sci. 2022, 89, 100965. [Google Scholar] [CrossRef]
  22. De Kleijne, K.; Hanssen, S.V.; Van Dinteren, L.; Huijbregts, M.A.J.; Van Zelm, R.; De Coninck, H. Limits to Paris compatibility of CO2 capture and utilization. One Earth 2022, 5, 168–185. [Google Scholar] [CrossRef]
  23. Karlsson, O.; Rocklöv, J.; Lehoux, A.P.; Bergquist, J.; Rutgersson, A.; Blunt, M.J.; Birnbaum, L.S. The human exposome and health in the Anthropocene. Int. J. Epidemiol. 2020, 50, 378–389. [Google Scholar] [CrossRef]
  24. Saleh, T.A. Nanomaterials and hybrid nanocomposites for CO2 capture and utilization: Environmental and energy sustainability. RSC Adv. 2022, 12, 23869–23888. [Google Scholar] [CrossRef]
  25. Yadav, S.; Mondal, S.S. A review on the progress and prospects of oxy-fuel carbon capture and sequestration (CCS) technology. Fuel 2022, 308, 122057. [Google Scholar] [CrossRef]
  26. Al-Mamoori, A.; Rownaghi, A.A.; Rezaei, F. Combined Capture and Utilization of CO2 for Syngas Production over Dual-Function Materials. ACS Sustain. Chem. Eng. 2018, 6, 13551–13561. [Google Scholar] [CrossRef]
  27. Kheirinik, M.; Ahmed, S.; Rahmanian, N. Comparative Techno-Economic Analysis of Carbon Capture Processes: Pre-Combustion, Post-Combustion, and Oxy-Fuel Combustion Operations. Sustainability 2021, 13, 13567. [Google Scholar] [CrossRef]
  28. Biermann, M.; Langner, C.; Roussanaly, S.; Normann, F.; Harvey, S. The role of energy supply in abatement cost curves for CO2 capture from process industry—A case study of a Swedish refinery. Appl. Energy 2022, 319, 119273. [Google Scholar] [CrossRef]
  29. Raganati, F.; Miccio, F.; Ammendola, P. Adsorption of Carbon Dioxide for Post-combustion Capture: A Review. Energy Fuels 2021, 35, 12845–12868. [Google Scholar] [CrossRef]
  30. Lai, L.S.; Jusoh, N.; Tay, W.H.; Yeong, Y.F.; Kiew, P.L. Rapid Prototyping for the Formulation of Monolith and Membrane for CO2 Removal. Sep. Purif. Rev. 2021, 51, 503–520. [Google Scholar] [CrossRef]
  31. 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. [Google Scholar] [CrossRef]
  32. de Riva, J.; Suarez-Reyes, J.; Moreno, D.; Díaz, I.; Ferro, V.; Palomar, J. Ionic liquids for post-combustion CO2 capture by physical absorption: Thermodynamic, kinetic and process analysis. Int. J. Greenh. Gas Control 2017, 61, 61–70. [Google Scholar] [CrossRef]
  33. Oexmann, J.; Kather, A. Minimising the regeneration heat duty of post-combustion CO2 capture by wet chemical absorption: The misguided focus on low heat of absorption solvents. Int. J. Greenh. Gas Control 2010, 4, 36–43. [Google Scholar] [CrossRef]
  34. Osman, A.I.; Hefny, M.; Abdel Maksoud, M.I.A.; Elgarahy, A.M.; Rooney, D.W. Recent advances in carbon capture storage and utilisation technologies: A review. Environ. Chem. Lett. 2021, 19, 797–849. [Google Scholar] [CrossRef]
  35. Xie, R.; Rutherford, M.; Peng, X. Formation of High-Quality I–III–VI Semiconductor Nanocrystals by Tuning Relative Reactivity of Cationic Precursors. J. Am. Chem. Soc. 2009, 131, 5691–5697. [Google Scholar] [CrossRef]
  36. Wang, R.; Zhao, H.; Wang, Y.; Qi, C.; Zhang, S.; Wang, L.; Li, M. Development of biphasic solvent for CO2 capture by tailoring the polarity of amine solution. Fuel 2022, 325, 124885. [Google Scholar] [CrossRef]
  37. El-Said, W.A.; Choi, J.-H.; Hajjar, D.; Makki, A.A.; Choi, J.-W. Fabrication of Hollow Nanocones Membrane with an Extraordinary Surface Area as CO2 Sucker. Polymers 2022, 14, 183. [Google Scholar] [CrossRef]
  38. Mostafavi, E.; Ashrafi, O.; Navarri, P. Assessment of process modifications for amine-based post-combustion carbon capture processes. Clean. Eng. Technol. 2021, 4, 100249. [Google Scholar] [CrossRef]
  39. Yang, Y.; Xu, W.; Wang, Y.; Shen, J.; Wang, Y.; Geng, Z.; Wang, Q.; Zhu, T. Progress of CCUS technology in the iron and steel industry and the suggestion of the integrated application schemes for China. Chem. Eng. J. 2022, 450, 138438. [Google Scholar] [CrossRef]
  40. Rao, A.B.; Rubin, E.S. Identifying Cost-Effective CO2 Control Levels for Amine-Based CO2 Capture Systems. Ind. Eng. Chem. Res. 2006, 45, 2421–2429. [Google Scholar] [CrossRef]
  41. Santos, S.P.; Gomes, J.F.; Bordado, J.C. Scale-Up Effects of CO2 Capture by Methyldiethanolamine (MDEA) Solutions in Terms of Loading Capacity. Technologies 2016, 4, 19. [Google Scholar] [CrossRef]
  42. Chen, G.; Chen, G.; Peruzzini, M.; Barzagli, F.; Zhang, R. Investigating the Performance of Ethanolamine and Benzylamine Blends as Promising Sorbents for Postcombustion CO2 Capture through 13C NMR Speciation and Heat of CO2 Absorption Analysis. Energy Fuels 2022, 36, 9203–9212. [Google Scholar] [CrossRef]
  43. Liu, X.; Ao, Q.; Shi, S.; Li, S. CO2 capture by alcohol ammonia based deep eutectic solvents with different water content. Mater. Res. Express 2022, 9, 015504. [Google Scholar] [CrossRef]
  44. Xiao, M.; Cui, D.; Yang, Q.; Liang, Z.; Puxty, G.; Yu, H.; Li, L.; Conway, W.; Feron, P. Role of mono- and diamines as kinetic promoters in mixed aqueous amine solution for CO2 capture. Chem. Eng. Sci. 2021, 229, 116009. [Google Scholar] [CrossRef]
  45. Kierzkowska-Pawlak, H. Kinetics of CO2 absorption in aqueous N, N-diethylethanolamine and its blend with N-(2-aminoethyl) ethanolamine using a stirred cell reactor. Int. J. Greenh. Gas Control 2015, 37, 76–84. [Google Scholar] [CrossRef]
  46. Yu, J.; Wang, S.; Yu, H. Modelling analysis of solid precipitation in an ammonia-based CO2 capture process. Int. J. Greenh. Gas Control. 2014, 30, 133–139. [Google Scholar] [CrossRef]
  47. Wagaarachchige, J.D.; Idris, Z.; Arstad, B.; Halstensen, M.; Jens, K.-J. Low-Viscosity Nonaqueous Sulfolane–Amine–Methanol Solvent Blend for Reversible CO2 Capture. Ind. Eng. Chem. Res. 2022, 61, 5942–5951. [Google Scholar] [CrossRef]
  48. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef]
  49. Atzori, F.; Barzagli, F.; Varone, A.; Cao, G.; Concas, A. CO2 absorption in aqueous NH3 solutions: Novel dynamic modeling of experimental outcomes. Chem. Eng. J. 2023, 451, 138999. [Google Scholar] [CrossRef]
  50. Aghel, B.; Gouran, A.; Behaein, S. Intensified biogas upgrading via various wastewater using microchannel. Chem. Eng. Process.-Process Intensif. 2022, 175, 108927. [Google Scholar] [CrossRef]
  51. Srivastava, R.K. Controlling SO2 Emissions—A Review of Technologies. 2000. Available online: http://purl.access.gpo.gov/GPO/LPS34639 (accessed on 10 June 2025).
  52. Chen, N. Morphology-Dependent Performance of Praseodymium-Based Materials via Controllable Synthesis for Energy Conversion and Storage. Ph.D. Thesis, UNSW Sydney, Sydney, Australia, 2020. Available online: http://hdl.handle.net/1959.4/67155 (accessed on 10 June 2025).
  53. Seddah, F.; Mustafa, D.; Bouziyah, A.; Battah, A. Carbon Dioxide Capture and Sequestration: Materials and Technology Potentials. 2020. Available online: https://hdl.handle.net/20.500.11888/16811 (accessed on 10 June 2025).
  54. Pragya, N.; Pandey, K.K.; Sahoo, P. A review on harvesting, oil extraction and biofuels production technologies from microalgae. Renew. Sustain. Energy Rev. 2013, 24, 159–171. [Google Scholar] [CrossRef]
  55. Li, L.; Conway, W.; Burns, R.; Maeder, M.; Puxty, G.; Clifford, S.; Yu, H. Investigation of metal ion additives on the suppression of ammonia loss and CO2 absorption kinetics of aqueous ammonia-based CO2 capture. Int. J. Greenh. Gas Control 2017, 56, 165–172. [Google Scholar] [CrossRef]
  56. Wang, F.; Zhao, J.; Miao, H.; Zhao, J.; Zhang, H.; Yuan, J.; Yan, J. Current status and challenges of the ammonia escape inhibition technologies in ammonia-based CO2 capture process. Appl. Energy 2018, 230, 734–749. [Google Scholar] [CrossRef]
  57. Chu, F.; Jon, C.; Yang, L.; Du, X.; Yang, Y. CO2 Absorption Characteristics in Ammonia Solution inside the Structured Packed. Ind. Eng. Chem. Res. 2016, 55, 3696–3709. [Google Scholar] [CrossRef]
  58. Hallett, J.P.; Welton, T. Room-Temperature Ionic Liquids: Solvents for Synthesis and Catalysis. 2. Chem. Rev. 2011, 111, 3508–3576. [Google Scholar] [CrossRef]
  59. Aryafard, M.; Minofar, B.; Cséfalvay, E.; Malinová, L.; Řeha, D. Novel room temperature ionic liquids and low melting mixtures based on imidazolium: Cheap ionic solvents for chemical and biological applications. J. Mol. Liq. 2021, 344, 117877. [Google Scholar] [CrossRef]
  60. Zhang, X.; Bai, L.; Zeng, S.; Gao, H.; Zhang, S.; Fan, M. Ionic Liquids: Advanced Solvents for CO2 Capture. In Energy Efficient Solvents for CO2 Capture by Gas-Liquid Absorption: Compounds, Blends and Advanced Solvent Systems; Budzianowski, W.M., Ed.; Springer International Publishing: Cham, Switzerland, 2017; pp. 153–176. [Google Scholar]
  61. Jiang, W.; Li, X.; Gao, G.; Wu, F.; Luo, C.; Zhang, L. Advances in applications of ionic liquids for phase change CO2 capture. Chem. Eng. J. 2022, 445, 136767. [Google Scholar] [CrossRef]
  62. Blanchard, L.A.; Hancu, D.; Beckman, E.J.; Brennecke, J.F. Green processing using ionic liquids and CO2. Nature 1999, 399, 28–29. [Google Scholar] [CrossRef]
  63. Cui, G.; Zhao, N.; Wang, J.; Wang, C. Computer-assisted design of imidazolate-based ionic liquids for improving sulfur dioxide capture, carbon dioxide capture, and sulfur dioxide/carbon dioxide selectivity. Chem. Asian J. 2017, 12, 2863–2872. [Google Scholar] [CrossRef] [PubMed]
  64. Simon, N.M.; Zanatta, M.; Neumann, J.; Girard, A.-L.; Marin, G.; Stassen, H.; Dupont, J. Cation-Anion-CO2 interactions in imidazolium based ionic liquid sorbents. ChemPhysChem 2018, 19, 2879–2884. [Google Scholar] [CrossRef]
  65. Sistla, Y.S.; Sridhar, V. Molecular understanding of carbon dioxide interactions with ionic liquids. J. Mol. Liq. 2021, 325, 115162. [Google Scholar] [CrossRef]
  66. Numpilai, T.; Pham, L.K.H.; Witoon, T. Advances in Ionic Liquid Technologies for CO2 Capture and Conversion: A Comprehensive Review. Ind. Eng. Chem. Res. 2024, 63, 19865–19915. [Google Scholar] [CrossRef]
  67. Ramdin, M.; de Loos, T.W.; Vlugt, T.J.H. State-of-the-Art of CO2 Capture with Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 8149–8177. [Google Scholar] [CrossRef]
  68. Cadena, C.; Anthony, J.L.; Shah, J.K.; Morrow, T.I.; Brennecke, J.F.; Maginn, E.J. Why Is CO2 So Soluble in Imidazolium-Based Ionic Liquids? J. Am. Chem. Soc. 2004, 126, 5300–5308. [Google Scholar] [CrossRef]
  69. Hu, H.; Li, F.; Xia, Q.; Li, X.; Liao, L.; Fan, M. Research on influencing factors and mechanism of CO2 absorption by poly-amino-based ionic liquids. Int. J. Greenh. Gas Control 2014, 31, 33–40. [Google Scholar] [CrossRef]
  70. Hu, P.; Zhang, R.; Liu, Z.; Liu, H.; Xu, C.; Meng, X.; Liang, M.; Liang, S. Absorption Performance and Mechanism of CO2 in Aqueous Solutions of Amine-Based Ionic Liquids. Energy Fuels 2015, 29, 6019–6024. [Google Scholar] [CrossRef]
  71. Kumar, S.; Cho, J.H.; Moon, I. Ionic liquid-amine blends and CO2BOLs: Prospective solvents for natural gas sweetening and CO2 capture technology—A review. Int. J. Greenh. Gas Control 2014, 20, 87–116. [Google Scholar] [CrossRef]
  72. Pinto, A.M.; Rodríguez, H.; Arce, A.; Soto, A. Carbon dioxide absorption in the ionic liquid 1-ethylpyridinium ethylsulfate and in its mixtures with another ionic liquid. Int. J. Greenh. Gas Control 2013, 18, 296–304. [Google Scholar] [CrossRef]
  73. Suleman, H.; Maulud, A.S.; Fosbøl, P.L.; Nasir, Q.; Nasir, R.; Shahid, M.; Nawaz, M.; Abunowara, M. A review of semi-empirical equilibrium models for CO2-alkanolamine-H2O solutions and their mixtures at high pressure. J. Environ. Chem. Eng. 2021, 9, 104713. [Google Scholar] [CrossRef]
  74. Alcantara, M.L.; de Carvalho, M.L.; Álvarez, V.H.; Ferreira, P.I.; Paredes, M.L.; Cardozo-Filho, L.; Silva, A.K.; Lião, L.M.; Pires, C.A.M.; Mattedi, S. High pressure vapor-liquid equilibria for binary carbon dioxide and protic ionic liquid based on ethanolamines+ butanoic acid. Fluid Phase Equilibria 2018, 460, 162–174. [Google Scholar] [CrossRef]
  75. Shaahmadi, F.; Shahraki, B.H.; Farhadi, A. The Solubility of Carbon Dioxide and Density for Binary Mixtures of 1-Butyl-3-methylimidazolium Acetate and 1-Butyl-3-methylimidazolium Tetrafluoroborate. J. Chem. Eng. Data 2019, 64, 584–593. [Google Scholar] [CrossRef]
  76. Murge, P.; Dinda, S.; Roy, S. Zeolite-based sorbent for CO2 capture: Preparation and performance evaluation. Langmuir 2019, 35, 14751–14760. [Google Scholar] [CrossRef]
  77. Fu, L.; Ren, Z.; Si, W.; Ma, Q.; Huang, W.; Liao, K.; Huang, Z.; Wang, Y.; Li, J.; Xu, P. Research progress on CO2 capture and utilization technology. J. CO2 Util. 2022, 66, 102260. [Google Scholar] [CrossRef]
  78. Wang, T.; Du, Y.; Yang, Y.; Jing, X.; Zhu, G. Imidazolium-Functionalized Ionic Porous Aromatic Frameworks for CO2 Capture and In Situ Conversion. Ind. Eng. Chem. Res. 2022, 61, 7284–7291. [Google Scholar] [CrossRef]
  79. Ding, M.; Jiang, H.-L. Incorporation of Imidazolium-Based Poly(ionic liquid)s into a Metal–Organic Framework for CO2 Capture and Conversion. ACS Catal. 2018, 8, 3194–3201. [Google Scholar] [CrossRef]
  80. Li, Y.; Zhang, X.; Lan, J.; Xu, P.; Sun, J. Porous Zn(Bmic)(AT) MOF with Abundant Amino Groups and Open Metal Sites for Efficient Capture and Transformation of CO2. Inorg. Chem. 2019, 58, 13917–13926. [Google Scholar] [CrossRef]
  81. Oschatz, M.; Antonietti, M. A search for selectivity to enable CO2 capture with porous adsorbents. Energy Environ. Sci. 2018, 11, 57–70. [Google Scholar] [CrossRef]
  82. Wang, J.; Huang, L.; Yang, R.; Zhang, Z.; Wu, J.; Gao, Y.; Wang, Q.; O’Hareb, D.; Zhong, Z. Recent advances in solid sorbents for CO2 capture and new development trends. Energy Environ. Sci. 2014, 7, 3478–3518. [Google Scholar] [CrossRef]
  83. Pardakhti, M.; Jafari, T.; Tobin, Z.; Dutta, B.; Moharreri, E.; Shemshaki, N.S.; Suib, S.L.; Srivastava, R. Trends in solid adsorbent materials development for CO2 capture. ACS Appl. Mater. Interfaces 2019, 11, 34533–34559. [Google Scholar] [CrossRef]
  84. Chakraborty, R.; Vilya, K.; Pradhan, M.; Nayak, A.K. Recent advancement of biomass-derived porous carbon based materials for energy and environmental remediation applications. J. Mater. Chem. A 2022, 10, 6965–7005. [Google Scholar] [CrossRef]
  85. Javed, M.; Zahoor, M.; Mazari, S.A.; Qureshi, S.S.; Sabzoi, N.; Jatoi, A.S.; Mubarak, N.M. An overview of effect of process parameters for removal of CO2 using biomass-derived adsorbents. Biomass Convers. Biorefin. 2021, 13, 4495–4513. [Google Scholar] [CrossRef]
  86. Mukherjee, A. Generation of Activated Carbon from Spent Coffee Grounds: Process Optimization, Kinetics and CO2 Capture. Ph.D. Thesis, University of Saskatchewan, Saskatoon, SK, Canada, 2022. Available online: https://harvest.usask.ca/items/cd27abca-b48b-4e0a-a82f-54f5600757b7 (accessed on 10 June 2025).
  87. Hao, W.; Björnerbäck, F.; Trushkina, Y.; Bengoechea, M.O.; Salazar-Alvarez, G.; Barth, T.; Hedin, N. High-performance magnetic activated carbon from solid waste from lignin conversion processes. 1. Their use as adsorbents for CO2. ACS Sustain. Chem. Eng. 2017, 5, 3087–3095. [Google Scholar] [CrossRef]
  88. Pavlenko, V.; Khosravi H, S.; Żółtowska, S.; Haruna, A.B.; Zahid, M.; Mansurov, Z.; Supiyeva, Z.; Galal, A.; Ozoemena, K.I.; Abbas, Q.; et al. A comprehensive review of template-assisted porous carbons: Modern preparation methods and advanced applications. Mater. Sci. Eng. R Rep. 2022, 149, 100682. [Google Scholar] [CrossRef]
  89. Saleem, A.; Zhang, Y.; Usman, M.; Haris, M.; Li, P. Tailored architectures of mesoporous carbon nanostructures: From synthesis to applications. Nano Today 2022, 46, 101607. [Google Scholar] [CrossRef]
  90. Ghavipanjeh, F.; Yavari, M.; Beygzadeh, M.; Sheikholeslami, Z. Facile fabrication of flexible super hydrophilic activated carbon sponge for continuous capillary adsorption. J. Water Process Eng. 2022, 49, 102945. [Google Scholar] [CrossRef]
  91. Penchah, H.R.; Ghaemi, A.; Jafari, F. Piperazine-modified activated carbon as a novel adsorbent for CO2 capture: Modeling and characterization. Environ. Sci. Pollut. Res. 2022, 29, 5134–5143. [Google Scholar] [CrossRef]
  92. Machida, M.; Yoo, P.; Amano, Y. Adsorption of nitrate from aqueous phase onto nitrogen-doped activated carbon fibers (ACFs). SN Appl. Sci. 2019, 1, 323. [Google Scholar] [CrossRef]
  93. Prasad, R.; Desai, C.; Prakash, S.O.; Prasad, S.; Bhat, T.; Kamble, B.; Sarvalkar, P.; Kanthe, A.; Kale, R.; Banga, S.; et al. A Critical Review on Design and Development of Carbonaceous Materials for Veterinary Medicine. ES Food Agrofor. 2022, 9, 15–38. [Google Scholar]
  94. Saleh, M.; Chandra, V.; Kemp, K.C.; Kim, K.S. Synthesis of N-doped microporous carbon via chemical activation of polyindole-modified graphene oxide sheets for selective carbon dioxide adsorption. Nanotechnology 2013, 24, 255702. [Google Scholar] [CrossRef]
  95. Kemp, K.C.; Chandra, V.; Saleh, M.; Kim, K.S. Reversible CO2 adsorption by an activated nitrogen doped graphene/polyaniline material. Nanotechnology 2013, 24, 235703. [Google Scholar] [CrossRef]
  96. Yang, L.; Hu, J.; He, L.; Tang, J.; Zhou, Y.; Li, J.; Ding, K. One-pot synthesis of multifunctional magnetic N-doped graphene composite for SERS detection, adsorption separation and photocatalytic degradation of Rhodamine 6G. Chem. Eng. J. 2017, 327, 694–704. [Google Scholar] [CrossRef]
  97. He, Y. Transition Metal Oxide Nanostructures in Heterogeneous Catalysis; Yale University: New Haven, CT, USA, 2020. [Google Scholar]
  98. Aguirre-Cruz, G.; Legorreta-Garcia, F.; Stanciu, L.; Aguirre-Alvarez, G. Synthesis of hierarchical silica zeolites for heterogenous catalysis and adsorption. Microporous Mesoporous Mater. 2022, 111, 112274. [Google Scholar] [CrossRef]
  99. Semeykina, V.; Zharov, I. Medium controlled aggregative growth as a key step in mesoporous silica nanoparticle formation. J. Colloid Interface Sci. 2022, 615, 236–247. [Google Scholar] [CrossRef]
  100. Qi, L.; Yang, W.; Zhang, L.; Liu, Q.; Fei, Z.; Chen, X.; Zhang, Z.; Tang, J.; Cui, M.; Qiao, X. Reinforced CO2 Capture on Amine-Impregnated Organosilica with Double Brush-like Additives Modified. Ind. Eng. Chem. Res. 2022, 61, 14859–14867. [Google Scholar] [CrossRef]
  101. Gebretatios, A.G.; Pillantakath, A.R.K.K.; Witoon, T.; Lim, J.-W.; Banat, F.; Cheng, C.K. Rice husk waste into various template-engineered mesoporous silica materials for different applications: A comprehensive review on recent developments. Chemosphere 2022, 310, 136843. [Google Scholar] [CrossRef]
  102. Wang, S.; Qiu, X.; Chen, Y.; Chen, S. Preparation and structure tuning of CO2 adsorbent based on in-situ amine-functionalized hierarchical porous polymer. Microporous Mesoporous Mater. 2022, 330, 111585. [Google Scholar] [CrossRef]
  103. Li, T.; Yang, C.; Tantikhajorngosol, P.; Sema, T.; Tontiwachwuthikul, P. Experimental investigations of CO2 absorption and catalyst-aided CO2 desorption performance of several different amines blending with a promoter. Chem. Eng. Sci. 2022, 264, 118177. [Google Scholar] [CrossRef]
  104. Fan, Y.; Jia, X. Progress in Amine-Functionalized Silica for CO2 Capture: Important Roles of Support and Amine Structure. Energy Fuels 2022, 36, 1252–1270. [Google Scholar] [CrossRef]
  105. Wu, K.; Ye, Q.; Wang, L.; Meng, F.; Dai, H. Mesoporous alumina-supported layered double hydroxides for efficient CO2 capture. J. CO2 Util. 2022, 60, 101982. [Google Scholar] [CrossRef]
  106. Silva, D.P.d.S.; Silva, D.C.M.; Ribeiro, T.R.S.; Solano, J.R.S.; da Silva, B.J.B.; Altino, S.A.; da Silva, A.O.S. Mesoporous Alumina Synthesis Using Carboxylic Acids to Enhance Performance in CO2 Adsorption. J. Environ. Chem. Eng. 2022, 10, 108928. [Google Scholar] [CrossRef]
  107. Feist, B.J.; Hill, J.M. CO2 and H2S adsorption on γ-Al2O3-supported lanthanum oxide. Energy Fuels 2015, 29, 6049–6056. [Google Scholar] [CrossRef]
  108. Zhou, K.; Mousavi, B.; Luo, Z.; Phatanasri, S.; Chaemchuen, S.; Verpoort, F. Characterization and properties of Zn/Co zeolitic imidazolate frameworks vs. ZIF-8 and ZIF-67. J. Mater. Chem. A 2017, 5, 952–957. [Google Scholar] [CrossRef]
  109. Adegoke, K.A.; Oyedotun, K.O.; Ighalo, J.O.; Amaku, J.F.; Olisah, C.; Adeola, A.O.; Iwuozor, K.O.; Akpomie, K.G.; Conradie, J. Cellulose derivatives and cellulose-metal-organic frameworks for CO2 adsorption and separation. J. CO2 Util. 2022, 64, 102163. [Google Scholar] [CrossRef]
  110. Fu, D.; Davis, M.E. Carbon dioxide capture with zeotype materials. Chem. Soc. Rev. 2022, 51, 9340–9370. [Google Scholar] [CrossRef] [PubMed]
  111. 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. [Google Scholar] [CrossRef]
  112. Coudert, F.-X.; Kohen, D. Molecular insight into CO2 “trapdoor” adsorption in zeolite Na-RHO. Chem. Mater. 2017, 29, 2724–2730. [Google Scholar] [CrossRef]
  113. Yaghi, O.M.; Kalmutzki, M.J.; Diercks, C.S. Introduction to Reticular Chemistry: Metal-Organic Frameworks and Covalent Organic Frameworks; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  114. Usman, M.; Iqbal, N.; Noor, T.; Zaman, N.; Asghar, A.; Abdelnaby, M.M.; Galadima, A.; Helal, A. Advanced Strategies in Metal-Organic Frameworks for CO2 Capture and Separation. Chem. Rec. 2022, 22, e202100230. [Google Scholar] [CrossRef]
  115. Soo, X.Y.D.; Lee, J.J.C.; Wu, W.-Y.; Tao, L.; Wang, C.; Zhu, Q.; Bu, J. Advancements in CO2 capture by absorption and adsorption: A comprehensive review. J. CO2 Util. 2024, 81, 102727. [Google Scholar] [CrossRef]
  116. Sharma, A.; Jindal, J.; Mittal, A.; Kumari, K.; Maken, S.; Kumar, N. Carbon materials as CO2 adsorbents: A review. Environ. Chem. Lett. 2021, 19, 875–910. [Google Scholar] [CrossRef]
  117. Milner, P.J.; Siegelman, R.L.; Forse, A.C.; Gonzalez, M.I.; Runčevski, T.; Martell, J.D.; Reimer, J.A.; Long, J.R. A Diaminopropane-Appended Metal–Organic Framework Enabling Efficient CO2 Capture from Coal Flue Gas via a Mixed Adsorption Mechanism. J. Am. Chem. Soc. 2017, 139, 13541–13553. [Google Scholar] [CrossRef]
  118. Milner, P.J.; Martell, J.D.; Siegelman, R.L.; Gygi, D.; Weston, S.C.; Long, J.R. Overcoming double-step CO2 adsorption and minimizing water co-adsorption in bulky diamine-appended variants of Mg2(dobpdc). Chem. Sci. 2018, 9, 160–174. [Google Scholar] [CrossRef]
  119. Ahmed, R.; Liu, G.; Yousaf, B.; Abbas, Q.; Ullah, H.; Ali, M.U. Recent advances in carbon-based renewable adsorbent for selective carbon dioxide capture and separation—A review. J. Clean. Prod. 2020, 242, 118409. [Google Scholar] [CrossRef]
  120. Hu, X.; Liu, L.; Luo, X.; Xiao, G.; Shiko, E.; Zhang, R.; Fan, X.; Zhou, Y.; Liu, Y.; Zeng, Z.; et al. A review of N-functionalized solid adsorbents for post-combustion CO2 capture. Appl. Energy 2020, 260, 114244. [Google Scholar] [CrossRef]
  121. Ochedi, F.O.; Liu, Y.; Adewuyi, Y.G. State-of-the-art review on capture of CO2 using adsorbents prepared from waste materials. Process Saf. Environ. Prot. 2020, 139, 1–25. [Google Scholar] [CrossRef]
  122. Zulys, A.; Yulia, F.; Muhadzib, N. Nasruddin Biological Metal–Organic Frameworks (Bio-MOFs) for CO2 Capture. Ind. Eng. Chem. Res. 2021, 60, 37–51. [Google Scholar] [CrossRef]
  123. Gutierrez-Ortega, A.; Montes-Morán, M.A.; Parra, J.B.; Sempere, J.; Nomen, R.; Gonzalez-Olmos, R. Comparative study of binderless zeolites and carbon molecular sieves as adsorbents for CO2 capture processes. J. CO2 Util. 2022, 61, 102012. [Google Scholar] [CrossRef]
  124. 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. [Google Scholar] [CrossRef]
  125. Flaig, R.W.; Popp, T.M.O.; Fracaroli, A.M.; Kapustin, E.A.; Kalmutzki, M.J.; Altamimi, R.M.; Fathieh, F.; Reimer, J.A.; Yaghi, O.M. The Chemistry of CO2 Capture in an Amine-Functionalized Metal–Organic Framework under Dry and Humid Conditions. J. Am. Chem. Soc. 2017, 139, 12125–12128. [Google Scholar] [CrossRef]
  126. Fracaroli, A.M.; Furukawa, H.; Suzuki, M.; Dodd, M.; Okajima, S.; Gándara, F.; Reimer, J.A.; Yaghi, O.M. Metal–Organic Frameworks with Precisely Designed Interior for Carbon Dioxide Capture in the Presence of Water. J. Am. Chem. Soc. 2014, 136, 8863–8866. [Google Scholar] [CrossRef]
  127. Liu, B.; Zhang, R.; Pan, C.-Y.; Jiang, H.-L. Unprecedented Li+ Exchange in an Anionic Metal–Organic Framework: Significantly Enhanced Gas Uptake Capacity. Inorg. Chem. 2017, 56, 4263–4266. [Google Scholar] [CrossRef]
  128. Shi, Z.; Tao, Y.; Wu, J.; Zhang, C.; He, H.; Long, L.; Lee, Y.; Li, T.; Zhang, Y.-B. Robust Metal–Triazolate Frameworks for CO2 Capture from Flue Gas. J. Am. Chem. Soc. 2020, 142, 2750–2754. [Google Scholar] [CrossRef]
  129. Bien, C.E.; Chen, K.K.; Chien, S.-C.; Reiner, B.R.; Lin, L.-C.; Wade, C.R.; Ho, W.S.W. Bioinspired Metal–Organic Framework for Trace CO2 Capture. J. Am. Chem. Soc. 2018, 140, 12662–12666. [Google Scholar] [CrossRef] [PubMed]
  130. Bermeo, M.; Vega, L.F.; Abu-Zahra, M.R.M.; Khaleel, M. Critical assessment of the performance of next-generation carbon-based adsorbents for CO2 capture focused on their structural properties. Sci. Total Environ. 2022, 810, 151720. [Google Scholar] [CrossRef] [PubMed]
  131. Wang, J.; Cui, S.; Li, Z.; Wen, S.; Ning, P.; Lu, S.; Lu, P.; Huang, L.; Wang, Q. Comprehensive investigation of dynamic CO2 capture performance using Mg/DOBDC as precursor to fabricate a composite of metallic organic framework and graphene oxide. Chem. Eng. J. 2021, 415, 128859. [Google Scholar] [CrossRef]
  132. Chen, Y.; Lv, D.; Wu, J.; Xiao, J.; Xi, H.; Xia, Q.; Li, Z. A new MOF-505@GO composite with high selectivity for CO2/CH4 and CO2/N2 separation. Chem. Eng. J. 2017, 308, 1065–1072. [Google Scholar] [CrossRef]
  133. Cao, Y.; Zhao, Y.; Lv, Z.; Song, F.; Zhong, Q. Preparation and enhanced CO2 adsorption capacity of UiO-66/graphene oxide composites. J. Ind. Eng. Chem. 2015, 27, 102–107. [Google Scholar] [CrossRef]
  134. Zhao, H.; Bahamon, D.; Khaleel, M.; Vega, L.F. Insights into the performance of hybrid graphene oxide/MOFs for CO2 capture at process conditions by molecular simulations. Chem. Eng. J. 2022, 449, 137884. [Google Scholar] [CrossRef]
  135. Le, V.N.; Vo, T.K.; Yoo, K.S.; Kim, J. Enhanced CO2 adsorption performance on amino-defective UiO-66 with 4-amino benzoic acid as the defective linker. Sep. Purif. Technol. 2021, 274, 119079. [Google Scholar] [CrossRef]
  136. Wang, J.-X.; Liang, C.-C.; Gu, X.-W.; Wen, H.-M.; Jiang, C.; Li, B.; Qian, G.; Chen, B. Recent advances in microporous metal–organic frameworks as promising adsorbents for gas separation. J. Mater. Chem. A 2022, 10, 17878–17916. [Google Scholar] [CrossRef]
  137. Jiang, Y.; Li, D.; Zhao, Y.; Sun, J. Hydrogen bond donor functionalized poly(ionic liquids)@MIL-101 for the CO2 capture and improving the catalytic CO2 conversion with epoxide. J. Colloid Interface Sci. 2022, 618, 22–33. [Google Scholar] [CrossRef] [PubMed]
  138. Zhao, M.; Ban, Y.; Yang, W. Assembly of ionic liquid molecule layers on metal–organic framework-808 for CO2 capture. Chem. Eng. J. 2022, 439, 135650. [Google Scholar] [CrossRef]
  139. Nguyen, H.T.D.; Tran, Y.B.N.; Nguyen, T.C.; Gándara, F.; Nguyen, P.T.K. A Series of Metal–Organic Frameworks for Selective CO2 Capture and Catalytic Oxidative Carboxylation of Olefins. Inorg. Chem. 2018, 57, 13772–13782. [Google Scholar] [CrossRef]
  140. Li, S.; Chang, S.-M.; Yin, M.-J.; Zhang, W.-H.; Sun, W.-S.; Shiue, A.; An, Q.-F. Build up ‘highway’ in membrane via solvothermal annealing for high-efficient CO2 capture. J. Membr. Sci. 2022, 652, 120444. [Google Scholar] [CrossRef]
  141. Hossain, I.; Park, S.; Husna, A.; Kim, Y.; Kim, H.; Kim, T.-H. PIM-PI-1 and Poly(ethylene glycol)/Poly(propylene glycol)-Based Mechanically Robust Copolyimide Membranes with High CO2-Selectivity and an Anti-aging Property: A Joint Experimental–Computational Exploration. ACS Appl. Mater. Interfaces 2021, 13, 49890–49906. [Google Scholar] [CrossRef]
  142. Nicotera, I.; Policicchio, A.; Conte, G.; Agostino, R.G.; Rehman, M.H.U.; Lufrano, E.; Simari, C. Quaternized polyepichlorohydrin-based membrane as high-selective CO2 sorbent for cost-effective carbon capture. J. CO2 Util. 2022, 63, 102135. [Google Scholar] [CrossRef]
  143. Sood, A.; Thakur, A.; Ahuja, S.M. Recent ameliorations in membrane based carbon capture technologies. Mater. Today Proc. 2022, 62, 6514–6529. [Google Scholar] [CrossRef]
  144. Wang, B.; Huang, W.; Zhu, Y.; Zhou, R.; Xing, W. Ultra-permeable high-selective SAPO-34 membranes for efficient CO2 capture. J. Membr. Sci. 2022, 650, 120420. [Google Scholar] [CrossRef]
  145. Zagho, M.M.; Hassan, M.K.; Khraisheh, M.; Al-Maadeed, M.A.A.; Nazarenko, S. A review on recent advances in CO2 separation using zeolite and zeolite-like materials as adsorbents and fillers in mixed matrix membranes (MMMs). Chem. Eng. J. Adv. 2021, 6, 100091. [Google Scholar] [CrossRef]
  146. Qian, J.; Song, E.; Lian, H.; Jiang, J.; Wang, C.; Pan, Y. High-performance ZIF-302 mixed-matrix membranes for efficient CO2 capture. Korean J. Chem. Eng. 2022, 39, 1020–1027. [Google Scholar] [CrossRef]
  147. Zheng, W.; Li, Z.; Sun, T.; Ruan, X.; Dai, Y.; Li, X.; Zhang, C.; He, G. PAN electrospun nanofiber skeleton induced MOFs continuous distribution in MMMs to boost CO2 capture. J. Membr. Sci. 2022, 650, 120330. [Google Scholar] [CrossRef]
  148. Bano, S.; Tariq, S.R.; Anjum, T.; Najam, M.; Usman, M.; Yasin, M.; Shafi, H.Z.; Khan, A.L. Development of highly permselective Mixed Matrix Membranes comprising of polyimide and Ln-MOF for CO2 capture. Chemosphere 2022, 307, 136051. [Google Scholar] [CrossRef]
  149. Zheng, W.; Liu, Z.; Ding, R.; Dai, Y.; Li, X.; Ruan, X.; He, G. Constructing continuous and fast transport pathway by highly permeable polymer electrospun fibers in composite membrane to improve CO2 capture. Sep. Purif. Technol. 2022, 285, 120332. [Google Scholar] [CrossRef]
  150. Wang, C.; Jiang, K.; Yu, H.; Yang, S.; Li, K. Copper electrowinning-coupled CO2 capture in solvent based post-combustion capture. Appl. Energy 2022, 316, 119086. [Google Scholar] [CrossRef]
  151. Jiang, W.; Tang, L.; Wang, Y.; Song, X.; Zhao, Z.; Lan, C.; Wu, Y.; Liu, T.; Xie, H. Electrochemical proton-coupled electron transfer driven amine regeneration for CO2 capture. Chem. Eng. J. 2025, 520, 166315. [Google Scholar]
  152. Sharifian, R.; Wagterveld, R.M.; Digdaya, I.A.; Xiang, C.; Vermaas, D.A. Electrochemical carbon dioxide capture to close the carbon cycle. Energy Environ. Sci. 2021, 14, 781–814. [Google Scholar] [CrossRef]
  153. Renfrew, S.E.; Starr, D.E.; Strasser, P. Electrochemical Approaches toward CO2 Capture and Concentration. ACS Catal. 2020, 10, 13058–13074. [Google Scholar] [CrossRef]
  154. Legrand, L.; Shu, Q.; Tedesco, M.; Dykstra, J.E.; Hamelers, H.V.M. Role of ion exchange membranes and capacitive electrodes in membrane capacitive deionization (MCDI) for CO2 capture. J. Colloid Interface Sci. 2020, 564, 478–490. [Google Scholar] [CrossRef]
  155. Wang, C.; Li, K.; Yu, H.; Yang, S.; Jiang, K. Electrochemical behavior of Cu-mediated electrowinning-coupled CO2 capture. Electrochim. Acta 2022, 422, 140571. [Google Scholar] [CrossRef]
  156. Xie, H.; Wu, Y.; Liu, T.; Wang, F.; Chen, B.; Liang, B. Low-energy-consumption electrochemical CO2 capture driven by biomimetic phenazine derivatives redox medium. Appl. Energy 2020, 259, 114119. [Google Scholar] [CrossRef]
  157. Gao, N.; Quiroz-Arita, C.; Diaz, L.A.; Lister, T.E. Intensified co-electrolysis process for syngas production from captured CO2. J. CO2 Util. 2021, 43, 101365. [Google Scholar] [CrossRef]
  158. Pérez-Gallent, E.; Vankani, C.; Sánchez-Martínez, C.; Anastasopol, A.; Goetheer, E. Integrating CO2 Capture with Electrochemical Conversion Using Amine-Based Capture Solvents as Electrolytes. Ind. Eng. Chem. Res. 2021, 60, 4269–4278. [Google Scholar] [CrossRef]
  159. Sanz-Pérez, E.S.; Murdock, C.R.; Didas, S.A.; Jones, C.W. Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, 11840–11876. [Google Scholar] [CrossRef]
  160. Castro-Muñoz, R.; Ahmad, M.Z.; Malankowska, M.; Coronas, J. A new relevant membrane application: CO2 direct air capture (DAC). Chem. Eng. J. 2022, 446, 137047. [Google Scholar] [CrossRef]
  161. Shi, X.; Xiao, H.; Azarabadi, H.; Song, J.; Wu, X.; Chen, X.; Lackner, K.S. Sorbents for the direct capture of CO2 from ambient air. Angew. Chem. Int. Ed. 2020, 59, 6984–7006. [Google Scholar] [CrossRef]
  162. Adamu, A.; Russo-Abegão, F.; Boodhoo, K. Process intensification technologies for CO2 capture and conversion—A review. BMC Chem. Eng. 2020, 2, 2. [Google Scholar] [CrossRef]
  163. Elhenawy, S.E.M.; Khraisheh, M.; AlMomani, F.; Walker, G. Metal-Organic Frameworks as a Platform for CO2 Capture and Chemical Processes: Adsorption, Membrane Separation, Catalytic-Conversion, and Electrochemical Reduction of CO2. Catalysts 2020, 10, 1293. [Google Scholar] [CrossRef]
  164. Sabatino, F.; Mehta, M.; Grimm, A.; Gazzani, M.; Gallucci, F.; Kramer, G.J.; Annaland, M. van S. Evaluation of a Direct Air Capture Process Combining Wet Scrubbing and Bipolar Membrane Electrodialysis. Ind. Eng. Chem. Res. 2020, 59, 7007–7020. [Google Scholar] [CrossRef]
  165. Zhu, Z.; Xie, W.; Wu, J.; Miao, Y.; Xiang, C.; Chen, C.; Ge, B.; Gan, Z.; Yang, F.; Zhang, M.; et al. Recent Advances in Direct Air Capture by adsorption. Chem. Soc. Rev. 2022, 51, 6574. [Google Scholar] [CrossRef]
  166. Sodiq, A.; Abdullatif, Y.; Aissa, B.; Ostovar, A.; Nassar, N.; El-Naas, M.; Amhamed, A. A Review on Progress Made in Direct Air Capture of CO2. Environ. Technol. Innov. 2023, 29, 102991. [Google Scholar] [CrossRef]
  167. Kotowicz, J.; Niesporek, K.; Baszczeńska, O. Advancements and Challenges in Direct Air Capture Technologies: Energy Intensity, Novel Methods, Economics, and Location Strategies. Energies 2025, 18, 496. [Google Scholar] [CrossRef]
  168. Leonzio, G.; Mwabonje, O.; Fennell, P.S.; Shah, N. Environmental performance of different sorbents used for direct air capture. Sustain. Prod. Consum. 2022, 32, 101–111. [Google Scholar] [CrossRef]
  169. Custelcean, R. Direct air capture of CO2 via crystal engineering. Chem. Sci. 2021, 12, 12518–12528. [Google Scholar] [CrossRef] [PubMed]
  170. Kumar, A.; Hua, C.; Madden, D.G.; O’Nolan, D.; Chen, K.-J.; Keane, L.-A.J.; Perry, J.J.; Zaworotko, M.J. Hybrid ultramicroporous materials (HUMs) with enhanced stability and trace carbon capture performance. Chem. Commun. 2017, 53, 5946–5949. [Google Scholar] [CrossRef]
  171. Piscopo, C.G.; Loebbecke, S. Strategies to Enhance Carbon Dioxide Capture in Metal-Organic Frameworks. ChemPlusChem 2020, 85, 538–547. [Google Scholar] [CrossRef]
  172. Bien, C.E.; Liu, Q.; Wade, C.R. Assessing the Role of Metal Identity on CO2 Adsorption in MOFs Containing M–OH Functional Groups. Chem. Mater. 2020, 32, 489–497. [Google Scholar] [CrossRef]
  173. Sujan, A.R.; Pang, S.H.; Zhu, G.; Jones, C.W.; Lively, R.P. Direct CO2 Capture from Air using Poly(ethylenimine)-Loaded Polymer/Silica Fiber Sorbents. ACS Sustain. Chem. Eng. 2019, 7, 5264–5273. [Google Scholar] [CrossRef]
  174. Williams, N.J.; Seipp, C.A.; Brethomé, F.M.; Ma, Y.-Z.; Ivanov, A.S.; Bryantsev, V.S.; Kidder, M.K.; Martin, H.J.; Holguin, E.; Garrabrant, K.A.; et al. CO2 capture via crystalline hydrogen-bonded bicarbonate dimers. Chem 2019, 5, 719–730. [Google Scholar] [CrossRef]
  175. Stamberga, D.; Einkauf, J.D.; Liu, M.; Custelcean, R. Direct Air Capture of CO2 via Reactive Crystallization. Cryst. Growth Des. 2024, 24, 4556–4562. [Google Scholar] [CrossRef]
  176. Wang, J.; Fu, R.; Wen, S.; Ning, P.; Helal, M.H.; Salem, M.A.; Xu, B.B.; El-Bahy, Z.M.; Huang, M.; Guo, Z.; et al. Progress and current challenges for CO2 capture materials from ambient air. Adv. Compos. Hybrid Mater. 2022, 5, 2721–2759. [Google Scholar] [CrossRef]
  177. Jiang, W.; Lin, Y.; Sun, C.; Sun, Y.; Zhu, Y. Comparative Review for Enhancing CO2 Capture Efficiency with Mixed Amine Systems and Catalysts. Molecules 2024, 29, 4618. [Google Scholar] [CrossRef] [PubMed]
  178. Arias, B.; Criado, Y.A.; Méndez, A.; Marqués, P.; Finca, I.; Abanades, J.C. Pilot Testing of Calcium Looping at TRL7 with CO2 Capture Efficiencies toward 99%. Energy Fuels 2024, 38, 14757–14764. [Google Scholar] [CrossRef]
  179. Hack, J.; Maeda, N.; Meie, D.M. Review on CO2 Capture Using Amine-Functionalized Materials. ACS Omega 2022, 7, 39520–39530. [Google Scholar] [CrossRef]
  180. Mantripragada, H.C.; Rubina, E.S. Chemical Looping for Pre-combustion CO2 Capture Performance and Cost Analysis. Energy Procedia 2013, 37, 618–625. [Google Scholar] [CrossRef]
  181. Küng, L.; Aeschlimann, S.; Charalambous, C.; McIlwaine, F.; Young, J.; Shannon, N.; Strassel, K.; Maesano, C.; Kahsar, R.; Pike, D.; et al. A roadmap for achieving scalable, safe, and low-cost direct air carbon capture and storage. Energy Environ. Sci. 2023, 16, 4280–4304. [Google Scholar] [CrossRef]
  182. Mukherjee, S.; Kumar, P.; Yang, A.; Fennell, P. Energy and exergy analysis of chemical looping combustion technology and comparison with pre-combustion and oxy-fuel combustion technologies for CO2 capture. J. Environ. Chem. Eng. 2015, 3, 2104–2114. [Google Scholar] [CrossRef]
  183. Ishida, M.; Zheng, D.; Akehata, T. Evaluation of a chemical-looping-combustion power-generation system by graphic exergy analysis. Energy 1987, 12, 147–154. [Google Scholar] [CrossRef]
  184. Abuelgasim, S.; Wang, W.; Abdalazeez, A. A brief review for chemical looping combustion as a promising CO2 capture technology: Fundamentals and progress. Sci. Total Environ. 2021, 764, 142892. [Google Scholar] [CrossRef]
  185. Ritcher, H.; Knoche, K. Reversibility of Combustion Process, Efficiency and Costing. Second Law Anal. Process 1983, 235, 71. [Google Scholar]
  186. Galinsky, N.L.; Huang, Y.; Shafiefarhood, A.; Li, F. Iron oxide with facilitated O2− transport for facile fuel oxidation and CO2 capture in a chemical looping scheme. ACS Sustain. Chem. Eng. 2013, 1, 364–373. [Google Scholar] [CrossRef]
  187. Lisi, L.; Mancino, G.; Cimino, S. Chemical looping oxygen transfer properties of Cu-doped lanthanum oxysulphate. Int. J. Hydrog. Energy 2015, 40, 2047–2054. [Google Scholar] [CrossRef]
  188. Wang, W.J.; Zhang, B.; Wang, G.P.; Li, Y.H. O−2 release of Mn-based oxygen carrier for chemical looping air separation (CLAS): An insight into kinetic studies. Aerosol Air Qual. Res. 2016, 16, 453–463. [Google Scholar] [CrossRef]
  189. Nazir, S.M.; Bolland, O.; Amini, S. Full plant scale analysis of natural gas fired power plants with pre-combustion CO2 capture and chemical looping reforming (CLR). In Proceedings of the 13th International Conference on Greenhouse Gas Control Technologies, Ghgt-13 2017, Lausanne, Switzerland, 14–18 November 2016; Dixon, T., Laloui, L., Twinning, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 2146–2155. [Google Scholar]
  190. Aghabararnejad, M.; Patience, G.S.; Chaouki, J. TGA and kinetic modeling of CO, Mn and Cu oxides for chemical looping gasification (CLG). Can. J. Chem. Eng. 2014, 92, 1903–1910. [Google Scholar] [CrossRef]
  191. Skulimowska, A.; Felice, L.D.; Kaminska-Pietrzak, N.; Celinska, A.; Plawecka, M.; Hercog, J.; Krauz, M.; Aranda, A. Chemical looping with oxygen uncoupling (CLOU) and chemical looping combustion (CLC) using copper-enriched oxygen carriers supported on fly ash. Fuel Process. Technol. 2017, 168, 123–130. [Google Scholar] [CrossRef]
  192. Wang, W.J. Thermodynamic investigation on hydrogen production via self-sufficient chemical looping reforming of glycerol (CLRG) using metal oxide oxygen carriers. J. Energy Inst. 2014, 87, 152–162. [Google Scholar] [CrossRef]
  193. Salkuyeh, Y.K.; Saville, B.A.; MacLean, H.L. Techno-economic analysis and life cycle assessment of hydrogen production from natural gas using current and emerging technologies. Int. J. Hydrog. Energy 2017, 42, 18894–18909. [Google Scholar] [CrossRef]
  194. Olaleye, A.K.; Wang, M.H. Techno-economic analysis of chemical looping combustion with humid air turbine power cycle. Fuel 2014, 124, 221–231. [Google Scholar] [CrossRef]
  195. Hossain, M.M.; de Lasa, H.I. Chemical-looping combustion (CLC) for inherent CO2 separations-a review. Chem. Eng. Sci. 2008, 63, 4433–4451. [Google Scholar] [CrossRef]
  196. Ortiz, C.; Romano, M.C.; Valverde, J.M.; Binotti, M.; Chacartegui, R. Process integration of calcium-looping thermochemical energy storage system in concentrating solar power plants. Energy 2018, 155, 535–551. [Google Scholar] [CrossRef]
  197. Colelli, G.; Chacartegui, R.; Ortiz, C.; Carro, A.; Arena, A.P.; Verda, V. Life cycle and environmental assessment of calcium looping (CaL) in solar thermochemical energy storage, Energy Convers. Manage 2022, 257, 115428. [Google Scholar] [CrossRef]
  198. Alalwan, H.A.; Alminshid, A.H. CO2 capturing methods: Chemical looping combustion (CLC) as a promising technique. Sci. Total Environ. 2021, 788, 147850. [Google Scholar] [CrossRef] [PubMed]
  199. Arias, B.; Diego, M.E.; Méndez, A.; Alonso, M.; Abanades, J.C. Calcium looping performance under extreme oxy-fuel combustion conditions in the calciner. Fuel 2018, 222, 711–717. [Google Scholar] [CrossRef]
  200. Han, R.; Wang, Y.; Xing, S.; Pang, C.; Hao, Y.; Song, C.; Liu, Q. Progress in reducing calcination reaction temperature of calcium-looping CO2 capture technology: A critical review. Chem. Eng. J. 2022, 450, 137952. [Google Scholar] [CrossRef]
  201. Tregambi, C.; Montagnaro, F.; Salatino, P.; Solimene, R. A model of integrated calcium looping for CO2 capture and concentrated solar power. Solar Energy 2015, 120, 208–220. [Google Scholar] [CrossRef]
  202. Duan, L.; Feng, T.; Jia, S.; Yu, X. Study on the performance of coal-fired power plant integrated with Ca-looping CO2 capture system with recarbonation process. Energy 2016, 115, 942–953. [Google Scholar] [CrossRef]
  203. Kumari, P.; Amoodi, N.A.; Dumée, L.F.; Hajaj, A.A. Integrated calcium looping technologies for enhanced CO2 valorisation—A critical review. Carbon Capture Sci. Technol. 2025, 14, 100371. [Google Scholar] [CrossRef]
  204. Manovic, V.; Anthony, E.J. Integration of Calcium and Chemical Looping Combustion using Composite CaO/CuO-Based Materials. Environ. Sci. Technol. 2011, 45, 10750–10756. [Google Scholar] [CrossRef]
  205. Baena-Moreno, F.M.; Rodríguez-Galán, M.; Vega, F.; Alonso-Fariñas, B.; Arenas, L.V.A.; Navarrete, B. Carbon capture and utilization technologies: A literature review and recent advances. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 41, 1403–1433. [Google Scholar] [CrossRef]
  206. Gulzar, A.; Gulzar, A.; Ansari, M.B.; He, F.; Gai, S.; Yang, P. Carbon dioxide utilization: A paradigm shift with CO2 economy. Chem. Eng. J. Adv. 2020, 3, 100013. [Google Scholar] [CrossRef]
  207. Nagireddi, S.; Agarwal, J.R.; Vedapuri, D. Carbon Dioxide Capture, Utilization, and Sequestration: Current Status, Challenges, and Future Prospects for Global Decarbonization. ACS Eng. Au 2024, 4, 22–48. [Google Scholar] [CrossRef]
  208. Zanatta, M. Materials for Direct Air Capture and Integrated CO2 Conversion: Advancement, Challenges, and Prospects. ACS Mater. Au 2023, 3, 576–583. [Google Scholar] [CrossRef]
  209. Podder, J.; Patra, B.R.; Pattnaik, F.; Nanda, S.; Dalai, A.K. A Review of Carbon Capture and Valorization Technologies. Energies 2023, 16, 2589. [Google Scholar] [CrossRef]
  210. Kondaveeti, S.; Abu-Reesh, I.M.; Mohanakrishna, G.; Bulut, M.; Pant, D. Advanced Routes of Biological and Bio-electrocatalytic Carbon Dioxide (CO2) Mitigation Toward Carbon Neutrality. Front. Energy Res. 2020, 8, 94. [Google Scholar] [CrossRef]
  211. Li, M.; Hua, B.; Wang, L.-C.; Sugar, J.D.; Wu, W.; Ding, Y.; Li, J.; Ding, D. Switching of metal-oxygen hybridization for selective CO2 electrohydrogenation under mild temperature and pressure. Nat. Catal. 2021, 4, 274–283. [Google Scholar] [CrossRef]
  212. Wang, F.; Dreisinger, D. Carbon mineralization with concurrent critical metal recovery from olivine. Proc. Natl. Acad. Sci. USA 2022, 119, e2203937119. [Google Scholar] [CrossRef]
  213. 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. [Google Scholar] [CrossRef]
  214. Ho, H.-J.; Iizuka, A.; Shibata, E. Carbon Capture and Utilization Technology without Carbon Dioxide Purification and Pressurization: A Review on Its Necessity and Available Technologies. Ind. Eng. Chem. Res. 2019, 58, 8941–8954. [Google Scholar] [CrossRef]
  215. Mikulčić, H.; Skov, I.R.; Dominković, D.F.; Alwi, S.R.W.; Manan, Z.A.; Tan, R.; Duić, N.; Mohamad, S.N.H.; Wang, X. Flexible Carbon Capture and Utilization technologies in future energy systems and the utilization pathways of captured CO2. Renew. Sustain. Energy Rev. 2019, 114, 109338. [Google Scholar] [CrossRef]
  216. Heeschen, K.U.; Deusner, C.; Spangenberg, E.; Priegnitz, M.; Kossel, E.; Strauch, B.; Bigalke, N.; Luzi-Helbing, M.; Haeckel, M.; Schicks, J.M. Production Method under Surveillance: Laboratory Pilot-Scale Simulation of CH4–CO2 Exchange in a Natural Gas Hydrate Reservoir. Energy Fuels 2021, 35, 10641–10658. [Google Scholar] [CrossRef]
  217. Qureshi, M.F.; Khandelwal, H.; Usadi, A.; Barckholtz, T.A.; Mhadeshwar, A.B.; Linga, P. CO2 hydrate stability in oceanic sediments under brine conditions. Energy 2022, 256, 124625. [Google Scholar] [CrossRef]
  218. Chauvy, R.; Meunier, N.; Thomas, D.; De Weireld, G. Selecting emerging CO2 utilization products for short-to mid-term deployment. Appl. Energy 2019, 236, 662–680. [Google Scholar] [CrossRef]
  219. Kumar, A.; Raizada, P.; Thakur, V.K.; Saini, V.; Khan, A.A.P.; Singh, N.; Singh, P. An overview on polymeric carbon nitride assisted photocatalytic CO2 reduction: Strategically manoeuvring solar to fuel conversion efficiency. Chem. Eng. Sci. 2021, 230, 116219. [Google Scholar] [CrossRef]
  220. Grim, R.G.; Huang, Z.; Guarnieri, M.T.; Ferrell, J.R.; Tao, L.; Schaidle, J.A. Transforming the carbon economy: Challenges and opportunities in the convergence of low-cost electricity and reductive CO2 utilization. Energy Environ. Sci. 2020, 13, 472–494. [Google Scholar] [CrossRef]
  221. Bragoni, V.; Gooßen, L.J. Development of Sustainable Catalytic Methods for the Chemical Utilisation of Carbon Dioxide and for the Valorisation of Cashew Nut Shell Liquid. Ph.D. Thesis, Ruhr-Universität Bochum, Bochum, Germany, 2020. [Google Scholar] [CrossRef]
  222. Lopes, E.J.; Ribeiro, A.P.; Martins, L.M. New trends in the conversion of CO2 to cyclic carbonates. Catalysts 2020, 10, 479. [Google Scholar] [CrossRef]
  223. George, A.; Shen, B.; Craven, M.; Wang, Y.; Kang, D.; Wu, C.; Tu, X. A Review of Non-Thermal Plasma Technology: A novel solution for CO2 conversion and utilization. Renew. Sustain. Energy Rev. 2021, 135, 109702. [Google Scholar] [CrossRef]
  224. Kamkeng, A.D.N.; Wang, M.; Hu, J.; Du, W.; Qian, F. Transformation technologies for CO2 utilisation: Current status, challenges and future prospects. Chem. Eng. J. 2021, 409, 128138. [Google Scholar] [CrossRef]
  225. Beydoun, K.; Klankermayer, J. Recent Advances on CO2 Utilization as C1 Building Block in C-N and C-O Bond Formation. In Organometallics for Green Catalysis; Dixneuf, P.H., Soulé, J.-F., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 39–76. [Google Scholar]
  226. Liu, Y.; Rui, Z. A storage-driven CO2 EOR for a net-zero emission target. Engineering 2022, 18, 79–87. [Google Scholar] [CrossRef]
  227. Mandal, A.; Ojha, K. Enhanced Oil Recovery; CRC Press: Boca Raton, FL, USA, 2023. [Google Scholar] [CrossRef]
  228. Xiaolong, C.; Yiqiang, L.; Xiang, T.; Huan, Q.; Xuebing, S.; Jianghao, L. Effect of gravity segregation on CO2 flooding under various pressure conditions: Application to CO2 sequestration and oil production. Energy 2021, 226, 120294. [Google Scholar] [CrossRef]
  229. Kumar, N.; Sampaio, M.A.; Ojha, K.; Hoteit, H.; Mandal, A. Fundamental aspects, mechanisms and emerging possibilities of CO2 miscible flooding in enhanced oil recovery: A review. Fuel 2022, 330, 125633. [Google Scholar] [CrossRef]
  230. Ajoma, E.; Saira; Sungkachart, T.; Ge, J.; Le-Hussain, F. Water-saturated CO2 injection to improve oil recovery and CO2 storage. Appl. Energy 2020, 266, 114853. [Google Scholar] [CrossRef]
  231. Solbakken, J.S.; Aarra, M.G. CO2 mobility control improvement using N2-foam at high pressure and high temperature conditions. Int. J. Greenh. Gas Control 2021, 109, 103392. [Google Scholar] [CrossRef]
  232. Føyen, T.; Brattekås, B.; Fernø, M.A.; Barrabino, A.; Holt, T. Increased CO2 storage capacity using CO2-foam. Int. J. Greenh. Gas Control 2020, 96, 103016. [Google Scholar] [CrossRef]
  233. Dai, Z.; Viswanathan, H.; Middleton, R.; Pan, F.; Ampomah, W.; Yang, C.; Jia, W.; Xiao, T.; Lee, S.-Y.; McPherson, B.; et al. CO2 Accounting and Risk Analysis for CO2 Sequestration at Enhanced Oil Recovery Sites. Environ. Sci. Technol. 2016, 50, 7546–7554. [Google Scholar] [CrossRef] [PubMed]
  234. Zhou, Q.; Liu, L.; Croiset, E.; Tan, Z.; Liu, Q.; Yang, J. Modeling study of the heat of absorption and solid precipitation for CO2 capture by chilled ammonia. RSC Adv. 2019, 9, 20075–20086. [Google Scholar] [CrossRef]
  235. Lv, Z.; Wang, T.; Qiao, K.; Yang, L.; Du, X. Experimental study on carbon dioxide absorption by aqueous ammonia with nickel and chromium ions in bubbling tower at low temperatures. Chem. Eng. Res. Des. 2022, 179, 298–307. [Google Scholar] [CrossRef]
  236. Ali, S.A.; Mulk, W.U.; Ullah, Z.; Khan, H.; Zahid, A.; Shah, M.U.H.; Shah, S.N. Recent Advances in the Synthesis, Application and Economic Feasibility of Ionic Liquids and Deep Eutectic Solvents for CO2 Capture: A Review. Energies 2022, 15, 9098. [Google Scholar] [CrossRef]
  237. Gu, H.; Song, G.; Niu, M.; Zhao, S.; Gao, Y.; Li, F. Sr2CeO4 as a robust high temperature sorbent for CO2 capture with near 100% sorbent conversion efficiency. Chem. Eng. J. 2022, 441, 135942. [Google Scholar] [CrossRef]
  238. Ren, L.; Li, M.; Wang, S.; Li, L.; Wu, Y. Effects of Oxygen-Vacancy-Promoted Ion Diffusion on CO2 Capture Behavior of CaO-Based Sorbents. Ind. Eng. Chem. Res. 2022, 61, 5527–5535. [Google Scholar] [CrossRef]
  239. Choi, H.L.; Jeong, Y.; Lee, H.; Bae, T.-H. High-Performance Mixed-Matrix Membranes Using a Zeolite@MOF Core–Shell Structure Synthesized via Ion-Exchange-Induced Crystallization and Post-Synthetic Conversion. JACS Au 2024, 4, 253–262. [Google Scholar] [CrossRef]
  240. Jiang, W.; Liu, W.; Wang, Y.; Zhao, Z.; Li, Q.; Wu, Y.; Liu, T.; Xie, H. Electrochemically Regenerated Amine for CO2 Capture Driven by a Proton-Coupled Electron Transfer Reaction. Ind. Eng. Chem. Res. 2022, 61, 13578–13588. [Google Scholar] [CrossRef]
  241. Rim, G.; Kong, F.; Song, M.; Rosu, C.; Priyadarshini, P.; Lively, R.P.; Jones, C.W. Sub-Ambient Temperature Direct Air Capture of CO2 using Amine-Impregnated MIL-101 (Cr) Enables Ambient Temperature CO2 Recovery. JACS Au 2022, 2, 380–393. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The most regularly practiced CO2 post-combustion collection systems.
Figure 1. The most regularly practiced CO2 post-combustion collection systems.
Catalysts 15 00807 g001
Figure 2. A general representation of CO2 captured by liquid and solid sorbents. (A Comparative Assessment of Emerging Solvents and Adsorbents for Mitigating CO2 Emissions from the Industrial Sector by Using Molecular Modeling Tools) Copyright © 2020 Bahamon, Alkhatib, Alkhatib, Builes, Sinnokrot, and Vega. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY).
Figure 2. A general representation of CO2 captured by liquid and solid sorbents. (A Comparative Assessment of Emerging Solvents and Adsorbents for Mitigating CO2 Emissions from the Industrial Sector by Using Molecular Modeling Tools) Copyright © 2020 Bahamon, Alkhatib, Alkhatib, Builes, Sinnokrot, and Vega. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY).
Catalysts 15 00807 g002
Figure 3. Key stages involved in post-combustion CO2 collection in liquids [34] (under a Creative Commons License).
Figure 3. Key stages involved in post-combustion CO2 collection in liquids [34] (under a Creative Commons License).
Catalysts 15 00807 g003
Figure 4. CO2 capture by methyldiethanolamine (MDEA) aqueous solution. Copyright under CC BY [41].
Figure 4. CO2 capture by methyldiethanolamine (MDEA) aqueous solution. Copyright under CC BY [41].
Catalysts 15 00807 g004
Figure 5. The collection of CO2 on alcohol and ammonia-based solvents [43]. Copyright@2022. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 License.
Figure 5. The collection of CO2 on alcohol and ammonia-based solvents [43]. Copyright@2022. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 License.
Catalysts 15 00807 g005
Figure 6. Effect of Ni and PZ on the reduction in ammonia loss [34]. Licensed under a Creative Commons Attribution 4.0 International License.
Figure 6. Effect of Ni and PZ on the reduction in ammonia loss [34]. Licensed under a Creative Commons Attribution 4.0 International License.
Catalysts 15 00807 g006
Figure 7. Flow diagram showing CO2 capture by ionic liquids.
Figure 7. Flow diagram showing CO2 capture by ionic liquids.
Catalysts 15 00807 g007
Figure 8. Solid sorbents for efficient CO2 capture: (a) In situ polymerization of imidazolium-based poly (ionic liquids) in MOF (MIL-101) shows good CO2 capture capability due to synergistic effect of Lewis acid and base sites in MOF and ILs. Reprinted with permission from [59], Copyright@2018, American Chemical Society. (b) Mixed ligand Zn(Bmic)(AT) MOF synthesized by solvothermal method shows good CO2 adsorption due to multiple open metal sites, an uncoordinated oxygen atom, and amino functional groups. Reprinted with permission from [60], Copyright@2019, American Chemical Society. (c) Two ionic PAFs (iPAF-167 and iPAF-168), with incorporated imidazolium groups, exhibit excellent CO2 capture and conversion properties, respectively.
Figure 8. Solid sorbents for efficient CO2 capture: (a) In situ polymerization of imidazolium-based poly (ionic liquids) in MOF (MIL-101) shows good CO2 capture capability due to synergistic effect of Lewis acid and base sites in MOF and ILs. Reprinted with permission from [59], Copyright@2018, American Chemical Society. (b) Mixed ligand Zn(Bmic)(AT) MOF synthesized by solvothermal method shows good CO2 adsorption due to multiple open metal sites, an uncoordinated oxygen atom, and amino functional groups. Reprinted with permission from [60], Copyright@2019, American Chemical Society. (c) Two ionic PAFs (iPAF-167 and iPAF-168), with incorporated imidazolium groups, exhibit excellent CO2 capture and conversion properties, respectively.
Catalysts 15 00807 g008
Figure 9. CO2 capture by utilizing amine-rich Mg-IRMOF-74-III under dry and humid conditions via chemisorption [125] available via Creative Commons Attribution 4.0 International License.
Figure 9. CO2 capture by utilizing amine-rich Mg-IRMOF-74-III under dry and humid conditions via chemisorption [125] available via Creative Commons Attribution 4.0 International License.
Catalysts 15 00807 g009
Figure 10. Lanthanide metal–organic frameworks (MOFs) for efficient CO2 capture and conversion: (a) CO2 separation via MOF-592 from a gas mixture, and (b) oxidative carboxylation process catalyzed by MOF-590. Reprint with permission [139] Copyright@2018, American Chemical Society.
Figure 10. Lanthanide metal–organic frameworks (MOFs) for efficient CO2 capture and conversion: (a) CO2 separation via MOF-592 from a gas mixture, and (b) oxidative carboxylation process catalyzed by MOF-590. Reprint with permission [139] Copyright@2018, American Chemical Society.
Catalysts 15 00807 g010
Figure 11. Metrics for pH-swing–based CO2 capture systems [152]. Licensed under a Creative Commons Attribution 3.0 Unported License.
Figure 11. Metrics for pH-swing–based CO2 capture systems [152]. Licensed under a Creative Commons Attribution 3.0 Unported License.
Catalysts 15 00807 g011
Figure 12. Key stages of electrochemical CO2 capture by the PCET process: (1) electrochemical acidification, (2) desorption of CO2, (3) electrochemical alkalization, and (4) absorption of CO2. Reprinted with permission from [151]. Copyright@2022, American Chemical Society.
Figure 12. Key stages of electrochemical CO2 capture by the PCET process: (1) electrochemical acidification, (2) desorption of CO2, (3) electrochemical alkalization, and (4) absorption of CO2. Reprinted with permission from [151]. Copyright@2022, American Chemical Society.
Catalysts 15 00807 g012
Figure 13. The reaction of CO2 with Zn-MOF fractionized with Zn-OH for direct air capture. Creative Commons Attribution-NonCommercial 3.0 Unported License from the Royal Society of Chemistry [169].
Figure 13. The reaction of CO2 with Zn-MOF fractionized with Zn-OH for direct air capture. Creative Commons Attribution-NonCommercial 3.0 Unported License from the Royal Society of Chemistry [169].
Catalysts 15 00807 g013
Figure 14. Amine-functionalized silica fibers for direct air capture [173].
Figure 14. Amine-functionalized silica fibers for direct air capture [173].
Catalysts 15 00807 g014
Figure 15. A schematic diagram shows that the CLC process consists of two reactors (fuel reactor (reduction) and air reactor (reoxidation)).
Figure 15. A schematic diagram shows that the CLC process consists of two reactors (fuel reactor (reduction) and air reactor (reoxidation)).
Catalysts 15 00807 g015
Figure 16. Two primary classes of CO2 utilization and their corresponding products.
Figure 16. Two primary classes of CO2 utilization and their corresponding products.
Catalysts 15 00807 g016
Table 1. Comparative overview of major CO2 capture technologies with performance, energy, cost, and maturity (TRL).
Table 1. Comparative overview of major CO2 capture technologies with performance, energy, cost, and maturity (TRL).
TechnologyCO2
Capture (%)
Energy (GJ/ton)Cost (USD/ton)TRLKey InsightRef.
Amine Absorption (MEA)85–953–440–807–8Mature, widely deployed, but energy-intensive[177]
Solid Sorbents (MOFs, ACs)70–902–350–1003–4High selectivity,
scalability challenges
[178]
Membrane Separation70–852–350–1205–6Compact design, low flux/fouling issues[179]
Calcium Looping (CaL)90–952.5–330–606–7Pilot-scale maturity, cost-effective potential[180]
Direct Air Capture (DAC)50–705–8300–6004–5Negative emissions, very high cost[181]
Table 2. Different types of sorbents based on their advantages and limitations.
Table 2. Different types of sorbents based on their advantages and limitations.
TechnologyAdvantagesLimitationsDevelopment and Future ProspectsReferences
Amine-based sorbentsMost widely used technology for CO2 capture
It has a high selectivity, a quick absorption rate, a big cycle capacity, and a high removal efficiency for CO2.
Able to collect CO2 from the flue gas stream even at very low partial pressure
It demands significant regeneration energy due to thermal stripping at a higher temperature. Corrosion and deterioration of absorbents are other limiting factors.Need to develop novel biphasic solvents and polyamine solvents for energy-efficient CO2 capture[36]
Ammonia-based sorbentsIt has a great capacity to collect CO2 with less energy consumption. Absorption capacity declines because of ammonia leakage.The application of chilled ammonia and the addition of transition metals can reduce ammonia leakage, but decrease efficiency. [234,235]
Ionic liquids as adsorbentsIt has extraordinarily high solubility for CO2, minimal vapor pressure, a low specific heat capacity, and strong thermal stability.
It requires minimal regeneration energy.
Expensive solvents, high viscosity, and reduced fuel absorption at lower CO2 partial pressuresPhase change and blends with other solvents may reduce the effective cost.[236]
Oxides and mineralsChemically stable, more cost-effective, and requires low regeneration energy
It can collect CO2 at an ambient temperature.
They are widely available in nature.
Suffer from rapid loss of reactivity under cyclic operation due to agglomeration, pore structure collapse, and a decrease in active sitesNeed to develop a highly stable novel sorbent[237]
Molecular sievesThey have less energy consumption, lower enthalpy, and tunable structures.Limited due to a lesser adsorption capacity at lower CO2 pressureSurface modifications, building interaction between the adsorbent and CO2[238]
MembranesLow maintenance and energy input requirements, a small environmental impact, excellent selectivity and dependability, and ease of operationRequire high-pressure gradients, the tradeoff between selectivity and permeability, and the low durabilityTuning and altering the characteristics of materials by surface treatment, pore size and porosity management, process engineering, and doping to enhance permeability[239]
Electrochemical capturingThe low-temperature operation, excellent energy economy, and flexible plug-and-play operating modeStill in the early stages of research and needs a lot of work to become widely usedA cost-effective system developed to capture CO2 and give valuable products[240]
Direct air capture (DAC)Can capture CO2 from a smaller and distributed point source, with no need for unnecessary transportHigh setup cost, lack of experiments at ambient and sub-ambient temperaturesDeveloping materials that show promising results over a wide range of temperatures[241]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akhdhar, A.; Al-Bogami, A.S.; Akhtar, N.; El-Said, W.A. Progress in Post-Combustion Carbon Dioxide Capture, Direct Air Capture, and Utilization. Catalysts 2025, 15, 807. https://doi.org/10.3390/catal15090807

AMA Style

Akhdhar A, Al-Bogami AS, Akhtar N, El-Said WA. Progress in Post-Combustion Carbon Dioxide Capture, Direct Air Capture, and Utilization. Catalysts. 2025; 15(9):807. https://doi.org/10.3390/catal15090807

Chicago/Turabian Style

Akhdhar, Abdullah, Abdullah S. Al-Bogami, Naeem Akhtar, and Waleed A. El-Said. 2025. "Progress in Post-Combustion Carbon Dioxide Capture, Direct Air Capture, and Utilization" Catalysts 15, no. 9: 807. https://doi.org/10.3390/catal15090807

APA Style

Akhdhar, A., Al-Bogami, A. S., Akhtar, N., & El-Said, W. A. (2025). Progress in Post-Combustion Carbon Dioxide Capture, Direct Air Capture, and Utilization. Catalysts, 15(9), 807. https://doi.org/10.3390/catal15090807

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop