Next Article in Journal / Special Issue
Surface-Assisted Laser Desorption/Ionization Mass Spectrometry Analysis of Latent Fingermarks Using Greenly Synthesized Silver Nanoparticles
Previous Article in Journal
The Antibacterial Performance of Implant Coating Made of Vancomycin-Loaded Polymer Material: An In Vitro Study
Previous Article in Special Issue
Vapor–Gas Deposition of Polymer Coatings on Metals from Azeotropic Solutions of Organosilanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Porous Carbon for CO2 Capture Technology: Unveiling Fundamentals and Innovations

by
Gazi A. K. M. Rafiqul Bari
* and
Jae-Ho Jeong
*
School of Mechanical Smart and Industrial Engineering, Gachon University, 1342 Seongnam-daero, Sujeong-gu, Seongnam-si 13120, Gyeonggi-do, Republic of Korea
*
Authors to whom correspondence should be addressed.
Surfaces 2023, 6(3), 316-340; https://doi.org/10.3390/surfaces6030023
Submission received: 28 July 2023 / Revised: 27 August 2023 / Accepted: 15 September 2023 / Published: 18 September 2023
(This article belongs to the Collection Featured Articles for Surfaces)

Abstract

:
Porous carbon is an emerging material for the capture of CO2 from point sources of emissions due to its high structural, mechanical, and chemical stability, along with reusability advantages. Currently, research efforts are mainly focused on high- or medium-pressure adsorption, rather than low-pressure or DAC (direct air capture) conditions. Highly porous and functionalized carbon, containing heteroatoms (N, O, etc.), is synthesized using different activation synthesis routes, such as hard template, soft template, and chemical activation, to achieve high CO2 capture efficiency at various temperatures and pressure ranges. Fundamental pore formation mechanisms with different activation routes have been evaluated and explored. Higher porosity alone can be ineffective without the presence of proper saturated diffusion pathways for CO2 transfer. Therefore, it is imperative to emphasize more rational multi-hierarchical macro-/meso-/micro-/super-/ultra-pore design strategies to achieve a higher utilization efficiency of these pores. Moreover, the present research primarily focuses on powder-based hierarchical porous carbon materials, which may reduce the efficiency of the capture performance when shaping the powder into pellets or fixed-bed shapes for applications considered. Therefore, it is imperative to develop a synthesis strategy for pelletized porous carbon and to explore its mechanistic synthesis route and potential for CO2 capture.

1. Introduction

The industrialization and urbanization of humans have resulted in a significant increase in CO2 emissions, estimated to be over 36 billion tons at present [1]. As a consequence, the current CO2 concentration levels in the environment have reached ~420 ppm, compared to ~280 ppm during preindustrial times (Figure 1) [2,3]. This increase is certain to cause hugely detrimental effects on human life through climate change. It requires immediate efforts to adopt a net-zero strategy to mitigate CO2 effects by efficiently capturing CO2 from point sources (such as oil refineries, heavy industries, cement, fossil fuel power plants, and industrial process plants) or through direct air capture (DAC) [4,5,6].
Over the years, CO2 capture or separation from flue gases has seen the development of the chemical absorption technique utilizing liquid alkanolamine solution process (scrubbing) on an industrial scale [7,8]. The liquid amine absorption technique has reached an industrial state-level advancement with considerable CO2 capture capacity (1.5 mmol g−1) and absorption rate (0.05–0.18 mol m−2 s−1) at a 4–25 vol.% concentration level [7,9,10]. Primary and secondary amines form carbamate with comparatively higher reaction kinetics with CO2, requiring high enthalpy absorption reactions of 80–90 and 70–75 kJ mol−1 CO2, respectively. In the case of tertiary amines, they form bicarbonate with CO2 with sluggish reaction kinetics, requiring 40–55 kJ mol−1 CO2 absorption enthalpy through a base-catalyzed hydration mechanism [11,12,13].
Consequently, solvent regeneration (stripping) requires a huge amount of energy (3–5 GJ t−1 CO2 captured), leading to large-scale industrial energy consumption and adding to the 70% additional cost, which is the main obstacle. However, the acceptable value for industrial application is 0.4–0.7 GJ t−1 CO2 [14,15,16]. Furthermore, the liquid amine absorption process faces other hurdles, such as liquid amine degradation due to high heating, solvent loss, equipment corrosion of reactors (steel/carbon), production of carcinogenic toxic products (nitramines, nitrosamines), and the utilization of large amounts of NaOH or KOH to restore CO2 absorption capacity of amines, resulting in a secondary pollutant to alkaline wastewater [17,18].
To address the drawbacks of the liquid amine absorption process mentioned above, the physical adsorption technique has become a focal point. CO2 physisorption on porous sorbent surfaces provides lower adsorption enthalpy (15–50 kJ mol−1) [19,20]. Consequently, the regeneration of CO2 gas molecules requires lower energy consumption and provides a completely clean surface for reusability. Solid adsorbents, such as MOFs (metal–organic frameworks), COFs (covalent–organic frameworks), zeolites, and carbon, offer alternative options for CO2 sorption from point sources, countering the drawbacks of the aqueous liquid amine absorption technique (84 kJ mol−1, USD 45–80 per ton of CO2 capture) [21]. Compared to the liquid amine absorption technique, solid adsorption offers lower heat of adsorption (MOF: 50 kJ mol−1, COF: 30–50 kJ mol−1, zeolites: 30–50 kJ mol−1, carbon: 10–35 kJ mol−1) [11,20,22]. Comparing the approximate capital costs/operation and maintenance expenses, including abandonment costs, for CO2 capture using polymeric membranes, chemical absorption, and physical absorption reveals values of USD 640/290, 839/365, and 590/271 million, respectively [23,24,25]. Depending on the specific characteristics and properties (surface area, pore volume, porosity, interconnected pore structure, surface functionality, etc.) of the solid adsorbent, an efficient level of capture and separation from flue gas can be achieved after post-combustion capture [26,27].
One of the technical criteria for solid adsorbent capture is the consideration of capture under dry conditions to ensure a reasonable capture efficiency. However, most of the post-combustion flue gas remains wet (5–7 vol.% water) at 40–80 °C, necessitating an energy-intensive drying process to remove moisture from the multi-component flue gas flow [27,28]. Ultimately, removing moisture from the gas flow or drying it backfires in terms of technical feasibility on an industrial scale. Additionally, wet condition CO2 capture significantly reduces the efficiency of MOF-, COF-, and zeolite-based solid adsorbents due to lower hydrothermal stability [27,29]. In comparison, carbon-based materials exhibit comparable stability under moisture and thermal conditions due to the structural and mechanical rigidity of the carbon structure.
Among the solid adsorbents, carbon-based adsorbents are considered promising candidates for CO2 capture technology due to their abundant sources, low cost, and scalable preparation process [30]. Additionally, carbon-based materials are known for their good thermal and chemical stability, low energy regeneration, and ability to maintain adsorption performance even under high moisture conditions [31,32]. They also offer the benefit of tunable opportunities for textural properties and surface functionalization [33].

2. Carbon as a Promising Candidate for CO2 Capture Technology

Carbon is considered a promising candidate for CO2 capture technology due to its capacity to offer a clean surface for reutilization, along with its thermal, mechanical, chemical, and moisture stability [34]. Its simple preparation strategy and low cost make it suitable for industrial applications [35].
A large variety of precursor sources, primarily biomass-based, such as rice husk, coconut shells, shrimp shells, sawdust, microalgae, carrot, olive stones, almond shells, kiwi peel, sugar beet pulp, palm stone saccharides, and glucose, are available [36]. Different activation strategies, including chemical methods (using urea, KOH, NaOH, ZnCl2, H3PO4, K2CO3, H2SO4, HNO3, HCl, H3PO4/KOH, NaOH, ZnCl2, CaCl2, K2CO3/H2O2, and KMnO4), physical methods (using oxidizing gases like O2, CO2, steam, air, and NH3), and both hard (using molten salt, eutectic salt, MOF, silica, zeolite) and soft (using organic precursors like melamine and cyanuric acid) approaches, can be employed to tailor the pore size and shape of the carbon structure [21,37].
Various templating methods, including gas templates, hard templates, soft templates, and chemical activation, are employed to design ultra-super-micropores for low-pressure CO2 capture and macro–mesopores for high-pressure CO2 capture [1]. Heteroatom doping on carbon and surface modification induces hetero-surface functionalized groups on the carbon structure to control surface chemistry [38,39]. CO2 molecules, being of acidic nature and possessing a quadrupole moment, have a competitive advantage over other gases for surface sorption due to the surface chemistry. Essentially, weak van der Waals forces induce the physisorption of CO2 molecules on the surface, which can be intensified by functional groups, such as acid–base interactions, H-bonds, or electrostatic interactions. As a result, this leads to a higher adsorption capacity for CO2 compared to other gas mixtures, making it more selective towards CO2 molecules [40,41].
During carbonization, a higher degree of aromaticity imparts a hydrophobic nature to carbon, allowing it to function effectively under hydrated conditions [42]. However, while heteroatom doping reduces polyaromatic condensation and reduce mechanical and chemical stability, it is essential to seek a rational design strategy for synthesizing carbon materials [43,44].

2.1. Carbon Products Preparation Mechanism

A carbon synthesis technique utilizes various technical processes such as pyrolysis (400–1100 °C), sol–gel, and hydrothermal processes [45]. Hydrothermal technology utilizes water as a medium in a closed reactor vessel to undergo processes such as hydrolysis, dehydration, decarboxylation, concentration polymerization, and aromatization [46]. It aligns with the principles of thermochemical conversion, making it a sustainable and green process similar to the natural formation of carbon. Hydrothermal methods offer the opportunity to produce various initial intermediate products that can later be combined with the general pyrolysis process, enabling the diversification of morphological and structural properties [47,48]. Pyrolysis is a thermochemical conversion process where materials decompose into solid (char), liquid (oil), and gas (syngas) as the temperature increases. The properties of pyrolysis products depend on various factors, including temperature, residence time, heating rate, and whether fast or slow pyrolysis is employed. In the conventional pyrolysis process, a furnace’s heating wall transfers heat to the material’s surface, resulting in a higher yield of char compared to microwave pyrolysis [49,50].
Fundamental carbon formation mechanisms bear the dehydration, intramolecular condensation, decarboxylation, and polymerization for oxygen-containing functional groups in the different temperature ranges (Figure 2) [1,51]. Various techniques or different sources are utilized in the formation of different intermediate products to provide the different characteristic carbon products. Initially, various sources or technical processes yield diverse intermediate products through dehydration, intramolecular condensation, or initial polymerization, resulting in an intermediate structural framework [52]. With increasing temperature, aromatic polycondensation steps occur, generating multiple gas components like NH3, CH4, CO, CO2, H2, water vapor, or structural degradation, which create structural, internal, and surface pores. Additionally, pressure modifies or influences the shape of the carbon [53,54].

2.2. Factors Influencing CO2 Adsorption on the Carbon Surface

The number of CO2 molecules adsorbed on the surface depends on several factors, including surface area, cumulative pore volume, pore sizes, pore interconnectivity, pore shape, and surface chemistry [55]. The low-pressure or high-pressure capture efficiency depends on the textural properties of the adsorbents. It is suggested that low-pressure conditions are overwhelmed by ultra-micropores (0.35–0.70 nm) to super-micropores (0.70–0.90 nm). At high-pressure conditions, pores size greater than micropores is more favorable [56,57]. Also, it is important to consider the competition among the different gases (CO2, N2, NOx, SO2, H2O vapor, etc.) to be separated or captured at a point source [58]. The kinetic diameter of CO2 (0.33 nm) is smaller than the kinetic diameters of nitrogen (0.36 nm) and oxygen (0.34 nm); CO2 adsorption favors the smaller than micro-size pores (<2 nm), while N2 adsorption is more favorable at mesoporous volumes (>2 nm) [59]. Surface chemistry of the adsorbents influences the CO2 sorption amounts, selectivity, dynamic kinetics due to CO2 molecule polarizability, and quadrupole nature. A CO2 molecule has higher polarizability (26.3 × 10−25 cm3) and quadrupole moment (13.4 × 10−40 C m2) in comparison to a N2 molecule of 17.6 × 10−25 cm3 and 4.7 × 10−40 C m2 [10,20,26].
CO2 physisorption on a carbon surface occurs due to the weak van der Waals force. Heteroatoms or heteroatom-containing functional groups influence the CO2 adsorption on the surface of carbon, leading to more adsorption potential, depending on the types of interactions (H-bond, electrostatic interaction, Lewis acid–base interaction) (Figure 3) [60,61,62]. There is still an ongoing debate about which types of interactions mainly occur or which forces are important in the CO2 physisorption on the carbon surface. It can be assumed that higher polarizability and quadrupole moment of CO2 molecules are more favorable for CO2 molecules to separate from the other gases (N2, O2, H2, CH4, etc.) [63].

3. Strategies for Modifying Carbon Structures

A three-dimensional interconnected porous carbon synthesis strategy has been developed over the years to enhance the surface area, porosity, and gas diffusion rate of the components [37]. Various activation techniques for the carbon structure have been explored, including chemical activation, soft and hard template methods, and in situ metal ion activation strategies. The basic synthesis strategy, challenges, and minute factors that influence the structure of carbon are discussed in the following section.

3.1. Chemical Activation

To produce highly porous and efficient carbon products, chemical activation of the carbon during carbonization has shown a potential utilization technique [65]. In the process, carbon precursors or carbon sources are simply mixed with an activating agent, and carbonization proceeds through different temperatures (350~800 °C) to achieve a porous carbon. Different acidic (e.g., HCl, FeCl3, ZnCl2, H3PO4, H2SO4, and HNO3), alkaline (e.g., KOH, NaOH, CaCO3 and K2CO3), or oxidant substances (e.g., H2O2, and KMnO4) have previously been utilized as activating agents to activate the carbon during carbonization [66,67,68]. Utilizing such an activating agent provides a very effective way to reduce the carbonization temperature and induce porosity or shape modification during the formation process. However, it cannot omit the extra step of removing the corrosive chemical from the structure using water.
Different groups have tried to determine the activation mechanism, nearly in complete agreement. Most commonly, KOH was used as the chemical activation agent, being a strong base [65,69]. Most groups predict the carbonization reaction as Equation (1) [70]:
2C + 6KOH → 2K2CO3 + 2K+ + 3H2,
Afterward, some groups have shown that volatile groups from the precursor (CO, CO2, and H2O), upon heating, react with KOH and form K2CO3:
2KOH + CO2 → 2K2CO3 + H2O,
The metallic potassium (K), resulting from the reaction of the K2CO3 with the carbon species, as well as its spatial positioning or penetration into the internal structure of carbon, promotes porosity. This internal porosity, upon removal during the washing step, provides the internal pore network:
K2CO3 + 2C → 2K + 3CO,
A proper ratio of precursors to activating agents influences the textural, mechanical, and structural properties of the carbon products. An optimally high ratio of the activating agent (approximately 4) enhances the porosity of the carbon by intercalating its structure and resulting in expansion of the pores. However, with a higher ratio (>6), the excessive enhancement breaks the pore walls and reduces the overall porosity. A carbon nanosphere was hydrothermally prepared from glucose [71]. This process controlled the spherical shape of the precursor. Subsequent heat treatment and activation using a KOH solution were employed to introduce finely tuned ultra-micropores into the structure (Figure 4).

3.2. Physical Activation

Physical activation is environmentally friendly, cheaper, simpler and requires no chemical agent. The gas template controls the textural properties [72,73]. During pyrolysis (300–1100 °C), degradation products generate different kinds of gas, such as CO2, CO, CH4, NO, H2O, and NH3, depending on the precursor’s source [37]. Sometimes, external sources of media like CO2, NH3, steam, He, Ar, air, or a mixture of gases (binary mixture) are utilized to activate the carbon surface (Figure 5) [74,75,76]. Several efforts have been made to increase the pore volume and surface area by utilizing external steam during pyrolysis [77]. Steam, along with the pyrolysis temperature, interacts with carbon to produce C(O) and H2, which are sources of various activated gases such as CO, CO2, and CH4. Consequently, this process promotes the formation of new pores and the expansion of pore sizes (Figure 5) [64,78]. The steam activation mechanism of the carbon surface is described in the following reactions [36,79]:
C + H2O → C(O) + H2,
C(O) → CO + C,
CO + H2O → CO2 + H2,
C + 2H2O → CO2 + 2H2,
C + CO2 → 2CO,
C + 2H2 → CH4,
CH4 + H2O → CO + 3H2,
Jung-Heo showed that steam activation of carbon from cellulose fiber expanded the existing ultra-micropores and produced additional ultra-micropores (Figure 6) [80]. The graphitic structure shattered due to the activation process, resulting in the size of the pores being larger than 2 nm. The activation process increased the specific surface area and pore volume to 1018 m2 g−1 and 0.43 cm3 g−1 from 452 m2 g−1 and 0.199 cm3 g−1, respectively. Consequently, the CO2 capture performance reached 4.41 mmol g−1 at 0 °C, 1 bar (1.21 mmol g−1 at 25 °C, 15% CO2) compared to without activation, which was 3.18 mmol g−1 at 0 °C, 1 bar [80].
Air activation is one of the efficient ways to simultaneously produce micropores and mesopores on the carbon surface while inducing O-containing functional groups such as C=O, C-O, -OH, and -COOH [81,82]. It is important to choose the activation time carefully. Initially, micropore formation on the carbon surface occurs predominantly, and, over time, it leads to an increase in mesopores due to the destruction of micropores [81,83]. The air activation mechanism of the carbon surface is described in the following reactions [84]:
2C + O2 (g) → 2C(O),
C + O2 (g) → CO2,
C(O) + O2 (g) → CO (g) + CO2,
In comparison to chemical activation, physical activation has less control over porosity development than chemical activation. The flow rate of N2 or other gases influences the textural properties of the carbonized products (10–200 mL min−1). During the carbonization process, the gas template of the degradation products provides the basic porosity and shapes of the carbonized products [85]. The faster or slower removal of the gas template depends on the gas flow rate of the carbonization medium. The effects of the flow rate on textural properties and shapes need to be explored. Additionally, the activation time (0.5~72 h) plays a crucial role in determining the textural properties, as higher activation time and degree promote higher porosity in the structure. At the same time, the porosity range differs based on the activation time or temperature range of carbonization. Smaller pores (ultra-to-micropores, ultra-pores: 0.35–0.70 nm, super-micropores: 0.70–0.90 nm, micropores: 0.90–2 nm) tend to shift to the range of higher pores with increasing temperature due to pore wall breakages that induce the conversion of super-micropores to mesopores [34,85].

3.3. Metal Ion Activation

The chemical and physical activation strategies are highly efficient for microporous carbon synthesis. Pores smaller than 1 nm (ultra-to-super-micropores) are particularly effective for low pressure CO2 uptake. Several research efforts have focused on the single-ion activation route to produce carbon with ultra-micropores in the structure [86]. In this process, monodispersed alkali metal ions (Li+, K+, Na+, Rb+, Cs+) are introduced into the carbon precursors of phenolic resin through a reaction between alkali hydroxide and the acidic groups of the resin during the solution process [87]. Monodispersed metal ions induce ultra-micropores in the carbon structure and provide the opportunity to tune pore sizes at the angstrom or sub-angstrom level (0.6–0.76 nm), depending on their individual activation strength and metal ionic size. The technique of metal ion activation has made significant progress through the in situ homogeneous activation process. In this method, metal salts are combined with organic precursors, such as alkali metal salts of carboxylic phenolic resin, through a hydrothermal process to produce xerogel. Under a temperature of 400 °C, these compounds undergo decomposition, resulting in the creation of metal carbonate (M2CO3), accompanied by the release of water vapor and CO2 gas, which act as templates. Moreover, the subsequent thermal decomposition of the metal carbonate (M2CO3) facilitates the generation of metal oxide (M2O) through a redox reaction between the metal carbonate (M2CO3) and carbon. This reaction is depicted by the equation M2CO3 + C → M2O + 2CO.
Furthermore, the metal oxide (M2O) significantly etches the carbon framework through a vigorous redox interaction between the metal oxide and the carbon framework. This is illustrated by the equation M2O + C → 2M + CO. As a result, the produced metallic component intercalates the crystalline graphitic lattice of the carbon. Consequently, this process leads to the expansion of the interlayer spacing within the lattice, thereby stabilizing the carbon structure. Subsequent to the washing process, in which the metallic portion is removed by water, the structure no longer reverts to its previous nonporous state. Instead, it retains the expanded lattice configuration, resulting in nanopores smaller than 1 nm. The precise size of these pores (spacing) can be finely tuned based on the dimensions of the metal ion, allowing for adjustment from Li+ to Cs+. Zhou et al. produced N-doped, uniformly ultra-microporous carbon material of approximately 0.5 nm using a metal activation process on K+-exchanged meta-aminophenol–formaldehyde resin [86]. This process led to a notable CO2 capture performance of 1.67 mmol g−1 under low pressure (0.15 bar, 25 °C).

3.4. Hard Template Activation

A hard template is utilized to associate with gaseous and solid carbon structures during high-temperature pyrolysis by duplicating or molecular printing on the surface of the template structure. The carbon can maintain its microstructure during the activation treatment process. Hard templates such as zeolites (zeolite x/y), porous metals (Zn–Co, Cu–Ni, Ni–Mn, etc.), metal powder (Ni/Cu powder), metal foam (Ni, Cu, Fe, etc.), silica (silica sphere, SBA-15, MCM-48), and MOFs (MOF-5, ZIF-8) have been explored to produce interconnected porous carbon structures [88,89]. Hard templates can provide effective and stable effects under confined conditions in spaces. In general, zeolites (0.1–10 nm), MOFs (0.1–50 nm), silica (10–1000 nm), salts (1 nm–10 µm), porous metals (100 nm–10 µm), metal powder (100 nm–1 µm), and metal foam (1 µm–10 µm) are utilized to infuse various ranges of porosity into the carbon structure [89]. Recently, Joseph et al. successfully synthesized interconnected hierarchical carbon by utilizing Coca-Cola, a soft drink, combined with the hard template KIT-6 (mesoporous silica nanoparticles) and activating the carbon using ZnCl2 [90]. This method resulted in an impressively high specific surface area of 2003 m2 g−1. In this process, the hard template acts as the source of the mesoporous structure, while ZnCl2 plays a pivotal role in enhancing micropores within the carbon structure and reducing the number of mesopores. However, an excessive use of the activation agent ZnCl2 can potentially lead to damage in both mesopore and micropore structures (Figure 7).
A molten salt, or a eutectic mixture of salts such as LiBr/KBr (348 °C), LiI/KI (275 °C), LiCl/KCl (353 °C), is utilized as a high-temperature solvent to synthesize porous carbon (Figure 8) [52,91,92,93]. In the synthesis process, the solvent intermediate products are carbonized under diluted conditions, resembling water under hydrothermal conditions, providing the opportunity for the formation of structured carbon [94]. Carbohydrate polymerization in the presence of strongly interacting metal ionic species creates nanopores in the carbon material, and the pore size can be tuned depending on the ionic species or the precursor-to-salt ratio. A proper choice of cation and counter anion controls the pore size. Additionally, the miscibility in the reaction medium with intermediate products, depending on their polarizability, influences the textural properties [95].
In situ hard salt templates of NaCl, ZnCl2, Na2CO3, and CaCO3 are utilized to enable large-scale production of carbon from glucose, sucrose, and biomass. This is due to their cheap and simple removal process through water washing steps. The silica and zeolite templates require HF or NaOH, while metal powder, porous metal, metal foam, and MOFs need an acidic solution for removal after shaping the carbon structure. Such types of aggressive removal steps sometimes damage the objective pores’ structure to some extent [89,96].
Overall, hard template carbon offers greater mechanical stability and robustness when compared to reactive or soft template synthesis carbon. This type of carbon displays elevated surface area and structural integrity. Moreover, it enables the production of various ordered structures, including thin-walled carbon layers, sieving-specific pore sizes, ordered mesoporous carbon, and close-packed ordered hexagonal or cubic mesoporous carbon. It also leads to carbon with bimodal porosity [97,98]. Nonetheless, the utilization of hard-templated carbon comes with notable drawbacks. A significant challenge involves the necessity of eliminating the hard template from the carbon through etching, achieved by an aggressive acid removal process like, when using a silica template, washing with HF or HCl. This extra step consumes time and resources. Additionally, hard-templated carbon is constrained in terms of pore size range, which becomes critical when aiming for a multi-hierarchical pore structure [97,99,100]. It is worth noting that even a high-surface-area carbon product lacking multi-hierarchical porosity may not deliver efficient gas capture performance [1,101]

3.5. Soft Template Activation

A soft template, derived from an organic co-polymer along with carbon precursors, is employed in the preparation of meso-/macroporous carbon products [102]. This soft template results in a supramolecular arrangement, combining carbon precursors, block co-polymers, and polymers to form a mesophase. The stability of this mesophase is achieved through either a catalytic or thermal treatment. Subsequent thermal decomposition removes the template, inducing porosity in the carbon structure. To create the mesophase, a carbon donor or carbon precursor is combined with a structure-directing organic or co-polymeric agent. Both components are necessary for mesophase formation. It is crucial for the template to withstand the thermal carbonization process while also facilitating the crosslinking of carbon with polymers [102,103,104].
The size and morphology of carbon mesopores can be readily controlled through the implementation of soft templating effects. The creation of supramolecular products, assembled through molecular cooperation, can be regulated in terms of their size, shape, and orientation by adjusting the ratio of carbon donors to directing co-polymers [105]. Additionally, factors such as the solubility of components in the solution, the solution ratio, and polarizability effects play crucial roles in the modification of soft templating products [106,107].
The mesophase structure primarily forms through hydrogen bonding between template polymers and carbon precursors. For instance, the hydrogen-bonded molecular cooperative assembled complex of melamine and cyanuric acid (MCA), along with various other complexes like MCA and glucose (MG), Pluronic F108, Pluronic F127, P123, and ZIF-8, contribute to the formation of the mesophase structure through H-bonding [1,108,109]. Subsequent thermal treatment leads to the compaction and shrinkage of materials, resulting in the isotropic formation of mesopores ranging from 3 to 30 nm in size. The template is either partially or completely removed at temperatures between 300 and 400 °C. Simultaneously, the carbonization process also involves the aromatic polycondensation of carbon products [1,110].

4. Significant Role of Surface Functional Groups in Porous Carbon

The surface functional group on the carbon structure plays a critical role in CO2 capture technology. Pure carbon bears adsorption energy of 6.6 kJ mol−1 for CO2, which is not sufficient to capture CO2 molecules and separate them from a mixture of gases (flue gas), even though CO2 molecules have a higher quadrupole moment. This becomes particularly important at low pressure or during direct air capture, where higher adsorption energy is required to effectively compete with other gases and moisture [101,111]. It is essential to introduce exposed heteroatom functional groups (such as N, B, O, P, S, etc.) on the surfaces of micro-/meso-/macropores of the carbon to enhance the amount of CO2 captured and the kinetics [22,112,113,114].
Different precursors or mixtures of hetero-precursors are utilized to prepare and induce various surface functional groups on the carbon [115,116]. Built-in functional groups efficiently improve the adsorption of CO2 molecules through acid–base interactions, H-bond interactions, or electrostatic quadrupole interactions. Numerical quantification shows that O-containing, N-containing, and N-/O-co-modified functional groups on carbon improve the CO2 adsorption energy to −14.3 to −22.6 kJ mol−1, −22.1 to −27.1 kJ mol−1, and −28.9 kJ mol−1, respectively [117,118,119].
Heteroatom doping or impregnation on the carbon structure influences its structural and mechanical stability. Heteroatom insertion counterbalances the structural or mechanical properties of the carbon by reducing the degree of polyaromatic condensation during carbonization at over 600 °C and destroying the pore structure, which poses challenges for maintaining the chemical, mechanical, and thermal stability of the carbon framework [1,120].
Xing et al. demonstrated the importance of H-bonding in N-doped activated carbon [31]. The CO2 adsorption capacity is roughly proportional to the N content of the activated carbon. The presence of the N atom on the carbon lattice modifies the electronic state of the hydrogen atom in the graphene layer, intensifying the interaction between the CO2 molecules and the carbon surface. The affinity for CO2 molecules through H-bonding is measured by the binding energy ΔE (kJ mol−1), where a higher value of binding energy denotes a stronger affinity for CO2 molecules. A CO2 molecule’s binding energy without N-containing carbon is 1.26 kJ mol−1, which is much lower than the average binding energy of N-containing carbon, which is 7.84 kJ mol−1. The binding energies with NH2, pyridinic N, and CH are 11.2, 10.5, and 9.1 kJ mol−1, respectively. The findings of this study cannot correlate with or explain the traditional acid–base interaction of CO2 molecules with carbon surfaces [31].
Wu et al., through DFT analysis, compared the importance of configuration types of N and O atoms in the carbon structure for the CO2 physisorption mechanism on the carbon surface at the molecular level [118]; the pyridinic carbon surface shows the highest attraction to CO2 molecules (Eads = −21.4 kJ mol−1) (Figure 9). Additionally, they explored O doping on the carbon structure, which enhances the CO2 adsorption energy (Eads = −9.7 to –17.0 kJ mol−1) compared to bare carbon (Eads = −5.4 kJ mol−1). They also observed a charge density difference with and without the N/O functional group, indicating no significant electron transfer between a CO2 molecule and a carbon surface, which establishes physisorption as the adsorption nature of functionalized carbon.
Bari et al. investigated the precise role of surface functional groups in CO2 sorption on the carbon surface using the CO2 temperature-programmed desorption profile (CO2-TPD) [1]. Initially, the study adsorbed CO2 at 0.2 bar and desorbed all physiosorbed CO2 from the porous carbon by purging with helium gas for an hour. The experiment then compared the roles of N- and O-containing functional groups in the temperature range of 50–400 °C through temperature-programmed desorption (Figure 10).
N-containing functional groups (pyridinic N, pyrrolic N, graphitic N, and oxidized N) in the carbon desorb approximately 24% of CO2 within the temperature range of 60–180 °C due to weak intermolecular interactions such as hydrogen bonding, acid–amine, or electrostatic interactions. They also demonstrate that O-containing functional groups (C-OH, (CO)OR, R = H, C) act as active sites for CO2 uptake by forming bicarbonate, carbamic acid, and carbamate due to strong chemical bonds (~100 kJ mol−1) [121]. Such chemisorbed CO2 is desorbed at temperatures above 200 °C [122,123,124].
Jianfei et al. successfully synthesized N- and P-co-doped porous carbon by employing chitosan aerogel, phytic acid, and dicyandiamide as precursor materials. Subsequently, they activated the carbon using NaNO3, thereby circumventing the need for the utilization of harsh alkalis like NaOH or KOH (Figure 11) [125]. This innovative method yielded an impressive CO2 capture performance of 5.31 mmol g−1 at a pressure of 1 bar and a temperature of 25 °C. Notably, when examining the carbon sample lacking P doping, a CO2 capture capacity of 1.89 mmol g−1 was observed at 1 bar and 25 °C. These results strongly suggest that the synergistic interaction between N- and P-doping induces modifications in the surface chemistry, consequently augmenting the CO2 capture performance [125].
Similarly, Davood et al. conducted a study involving the incorporation of various heteroatoms (B, N, P, S) into activated porous carbon to enhance CO2 capture performance [126]. A comparison was made among these elements in terms of their impact. Among the considered elements (B, N, P, and S), phosphorus (P)-doping exhibited the highest efficiency in enhancing CO2 capture, yielding an uptake of 7.13 mmol g−1 at 1 bar and 20 °C (Figure 12a). In this context, it is worth noting that the carbon atom (C) has a higher electronegativity (2.5) than the phosphorus (P) atom (2.1), which stands in contrast to the situation in C–N interactions, where the nitrogen (N) atom possesses an electronegativity of 3.0. This discrepancy results in the phosphorus atom displaying a relatively more positive electrostatic charge, while the carbon atom withdraws bonding electrons. Simultaneously, due to the larger atomic radius of phosphorus, the graphene structure of carbon experiences more significant distortion upon interaction. The distribution of the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) is predominantly centered around the phosphorus atom. This arrangement greatly facilitates Lewis acid–base interactions with CO2 molecules, thereby contributing significantly to the overall CO2 capture performance [126].
A hydrochar-based carbon material was synthesized through the combination of glucose and ethylene diamine, and it exhibited efficient CO2 capture performance compared to carbon derived from urea activation by KOH and carbon derived from urea activation by ZIF8, specifically in the low-pressure range of 0 to 1 bar (Figure 12b) [127]. To assess the influence of pore structure and chemical properties on CO2 adsorption capacity, a machine learning approach was employed using a dataset comprising 1594 CO2 adsorption data points. The findings of the study highlight that, under low-pressure conditions (0.1–1 bar), ultra-micropores and carbon-containing nitrogen (N) functional groups have the most significant impact on CO2 capture performance (Figure 12c,d). As the pore size increases, the amount of CO2 adsorption decreases, as smaller pore sizes provide a greater likelihood of interaction between the pore walls and CO2 molecules. Notably, the machine learning outcomes emphasize that pore sizes smaller than 0.7 nm play a critical role in capturing CO2 at 1 bar. Regarding functional groups, N-containing groups such as pyridinic N, pyrrolic N, and quaternary N exhibit substantial effects on CO2 uptake, surpassing the effects of O-containing functional groups like hydroxyls, carboxylic acids, and ketones [127].

5. Envisioning Future Research Prospects and Forging a Focused Strategy

It is interesting to note that, until now, the development of multi-hierarchical porous carbon has mainly focused on powder-based highly efficient carbon products (Table 1).
However, for practical industrial-scale applications, it is necessary to shape the powder into pellets [147]. Achieving a well-defined fixed shape and particle size before practical application requires additional fabrication steps, such as extrusion, granulation, pressing, and binder inclusion [148].
After shaping the powder into fixed pellets for practical use, there is a high concern for pressure drops across the reactor bed, which can limit mass and heat transfer [149,150,151,152]. Moreover, the inclusion of a binder or other fabrication processes can lead to changes or reductions in surface area, pore size, and pore volume [153,154,155]. This can also cause a significant reduction in the amount of active components and efficient functional activity, thereby affecting structural and material stability [156,157].
To overcome these drawbacks, research trends require a shift towards a pelletization-based synthesis strategy as the ultimate solution to obtain well-preserved fixed-shaped and functionally hierarchical porous carbon products [158,159]. However, only a few research efforts have focused on this approach, and they are still far from achieving practical efficient performance [160].
It could be assumed that pelletized-state carbonization, rather than powder state carbonization or pyrolysis, will effectively improve, modify, or influence the morphological state of carbon products, due to the close contact with the carbon intermediate. It could also be assumed that the gas template is more aggressively utilized at the molecular level during the precursor’s carbonization.
Figure 13 shows that, when precursors are carbonized at higher temperatures in powder form, the self-degraded or mass-loss gas template may be able to escape easily from the pyrolyzed medium. On the other hand, in a close-packed or void-filling state, the degradation products of the gas template interact more intensely and effectively at the molecular level, occupying the space left by the mass loss and providing more efficient textural properties. The proposed synthesis mechanism could reduce the technical complexity of activation by chemical or secondary physical activation (hard template, soft organic template, steam, air, NH3, He, Ar, etc.) processes. Alternatively, it could be combined with the existing processes to explore new possibilities for obtaining more efficient carbon products.
Different activation processes come with their own set of advantages and disadvantages (Table 2). Achieving a multi-hierarchical porous carbon is crucial when considering CO2 capture efficiency, selectivity, kinetics, and a wide pressure range (including DAC, medium, and high pressures). To optimize low-pressure capture efficiency, an abundance of ultra-to-super-micropores is highly anticipated, while high-pressure scenarios tend to benefit from the presence of meso-/macropores.
To ensure both swift capture efficiency and effective utilization of ultra-micropores, it is essential to integrate them harmoniously with meso-/macropores in the overall morphological design or engineering. Furthermore, the inclusion of heteroatom-containing compositions is paramount in creating an efficient and selective adsorbent, leveraging surface chemistry. This task is challenging, as it requires the introduction of functional groups while maintaining a high degree of aromatic condensation within carbon products, in order to achieve a structurally stable carbon end-product. Combining all necessary conditions within a single-pot synthesis strategy proves to be a formidable challenge, both in terms of attainment and the engineering of an efficient pore structure. Combining metal ion activation with soft templating and pelletization followed by carbonization in a one-pot process has the potential to produce multi-hierarchical carbon products.

6. Summary

Porous carbon-based materials surpass many limitations and exhibit outstanding CO2 adsorption/desorption capabilities, as well as excellent clean surface cycling durability, mechanical strength, chemical resistance, and moisture stability. Various fundamental strategies have been developed to induce the efficient formation of hierarchical porous carbon products. However, applying different pressure ranges to capture CO2 within multi-porosity structures poses a challenging task. This is because ultra-/super-micropores are favorable for low-pressure and low-quantity CO2 adsorption, while meso-/macropores are more suited for high-pressure conditions.
Additionally, when considering the selectivity of ultra-to-micropores for CO2 adsorption, the meso-/macropores region competes with other gases. Considering the molecular sieve effects, even with a higher surface area and a large number of ultra-/super-micropores, there is an unfortunate limitation in their ability to provide effective adsorption and proper kinetics due to improper diffusion pathways. Therefore, it becomes essential to devise a proper and rational design strategy to achieve a multi-hierarchical pore structure encompassing macro-/meso-/micro-/super-/ultra-pores. Moreover, the use of heteroatom-containing sources, along with inappropriate strategies for activation and templating, hinders the degree of aromatic condensation during carbonization, leading to unbalanced formation of macro-/mesopores.
The concept of merging soft templates and metal ion activation, alongside the innovative pelletization method, holds the prospect of promising potential across a diverse spectrum of practical applications (Figure 14). The ability of a soft template to direct the structure helps in incorporating heteroatoms into the carbon structure from diverse multi-component precursor sources. This plays a crucial role in maintaining the appropriate level of carbon polycondensation during the carbonization process, contributing to the mechanical strength of the structure and facilitating the development of meso- and macropores. The effectiveness of capturing CO2 and performing low-pressure capture relies on the presence of ultra-to-super-micropores. Metal ion activation, in this context, facilitates the efficient creation of these ultra-to-super-micropores within the structure. Additionally, the challenges related to structural and functional degradation that often arise when shaping synthesized efficient powder for application can effectively be addressed through the proposed pre-pelletization strategy. It can be inferred that the pre-pelletization strategy makes effective use of gas templates, which share similarities with the physical activation strategy. This approach holds the potential to provide a solution to the limitations associated with post-synthesis shaping while preserving the desired structural and functional attributes.

Author Contributions

Conceptualization: G.A.K.M.R.B. and J.-H.J.; Methodology: G.A.K.M.R.B.; Validation: G.A.K.M.R.B. and J.-H.J.; Writing—Original Draft: G.A.K.M.R.B.; Writing—Review and Editing: G.A.K.M.R.B. and J.-H.J.; Resources: G.A.K.M.R.B. and J.-H.J.; Supervision: G.A.K.M.R.B. and J.-H.J.; Funding Acquisition: J.-H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) (20223030020070, Development of an X-ray-based non-destructive inspection platform for maintaining the blade lightning system), and also supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) (2021202080023B, Development and demonstration of thermoelectric power generation system for marine application by waste heat utilization).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bari, G.A.R.; Kang, H.J.; Lee, T.G.; Hwang, H.J.; An, B.H.; Seo, H.W.; Ko, C.H.; Hong, W.H.; Jun, Y.S. Dual-templating-derived porous carbons for low-pressure CO2 capture. Carbon Lett. 2023, 33, 811–822. [Google Scholar] [CrossRef]
  2. Singh, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar] [CrossRef]
  3. Ahmad, H.H.; Saleem, F.; Arif, H. Evaluation of Catastrophic Global Warming due to Coal Combustion, Paradigm of South Asia. Int. J. Innov. Sci. Technol. 2021, 3, 198–207. [Google Scholar] [CrossRef]
  4. Pedraza, J.; Zimmermann, A.; Tobon, J.; Schomäcker, R.; Rojas, N. On the road to net zero-emission cement: Integrated assessment of mineral carbonation of cement kiln dust. Chem. Eng. J. 2021, 408, 127346. [Google Scholar] [CrossRef]
  5. Janakiram, S.; Santinelli, F.; Costi, R.; Lindbråthen, A.; Nardelli, G.M.; Milkowski, K.; Ansaloni, L.; Deng, L. Field trial of hollow fiber modules of hybrid facilitated transport membranes for flue gas CO2 capture in cement industry. Chem. Eng. J. 2021, 413, 127405. [Google Scholar] [CrossRef]
  6. Fujikawa, S.; Selyanchyn, R.; Kunitake, T. A new strategy for membrane-based direct air capture. Polym. J. 2021, 53, 111–119. [Google Scholar] [CrossRef]
  7. Rochelle, G.T. Amine Scrubbing for CO2 Capture. Science 2009, 325, 1652–1654. [Google Scholar] [CrossRef] [PubMed]
  8. Singh, A.; Sharma, Y.; Wupardrasta, Y.; Desai, K. Selection of amine combination for CO2 capture in a packed bed scrubber. Resour. -Effic. Technol. 2016, 2, S165–S170. [Google Scholar] [CrossRef]
  9. Li, X.; Wang, S.; Chen, C. Experimental study of energy requirement of CO2 desorption from rich solvent. Energy Procedia 2013, 37, 1836–1843. [Google Scholar] [CrossRef]
  10. Patel, H.A.; Byun, J.; Yavuz, C.T. Carbon Dioxide Capture Adsorbents: Chemistry and Methods. ChemSusChem 2017, 10, 1303–1317. [Google Scholar] [CrossRef] [PubMed]
  11. Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.; Xie, B.; Toe, C.Y.; Zhu, X.; Wang, J.; et al. Industrial carbon dioxide capture and utilization: State of the art and future challenges. Chem. Soc. Rev. 2020, 49, 8584–8686. [Google Scholar] [CrossRef]
  12. Wang, M.; Joel, A.S.; Ramshaw, C.; Eimer, D.; Musa, N.M. Process intensification for post-combustion CO2 capture with chemical absorption: A critical review. Appl. Energy 2015, 158, 275–291. [Google Scholar] [CrossRef]
  13. Zhang, S.; Du, M.; Shao, P.; Wang, L.; Ye, J.; Chen, J.; Chen, J. Carbonic Anhydrase Enzyme-MOFs Composite with a Superior Catalytic Performance to Promote CO 2 Absorption into Tertiary Amine Solution. Environ. Sci. Technol. 2018, 52, 12708–12716. [Google Scholar] [CrossRef]
  14. Luis, P. Use of monoethanolamine (MEA) for CO2 capture in a global scenario: Consequences and alternatives. Desalination 2016, 380, 93–99. [Google Scholar] [CrossRef]
  15. MacDowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, A.; Jackson, G.; Adjiman, C.S.; Williams, C.K.; Shah, N.; Fennell, P. An overview of CO2 capture technologies. Energy Environ. Sci. 2010, 3, 1645–1669. [Google Scholar] [CrossRef]
  16. Sabatino, F.; Grimm, A.; Gallucci, F.; van Sint Annaland, M.; Kramer, G.J.; Gazzani, M. A comparative energy and costs assessment and optimization for direct air capture technologies. Joule 2021, 5, 2047–2076. [Google Scholar] [CrossRef]
  17. Barzagli, F.; Mani, F.; Peruzzini, M. A Comparative Study of the CO2 Absorption in Some Solvent-Free Alkanolamines and in Aqueous Monoethanolamine (MEA). Environ. Sci. Technol. 2016, 50, 7239–7246. [Google Scholar] [CrossRef]
  18. Mazari, S.A.; Ali, B.S.; Jan, B.M.; Saeed, I.M. Degradation study of piperazine, its blends and structural analogs for CO2 capture: A review. Int. J. Greenh. Gas Control 2014, 31, 214–228. [Google Scholar] [CrossRef]
  19. Chang, B.; Shi, W.; Yin, H.; Zhang, S.; Yang, B. Poplar catkin-derived self-templated synthesis of N-doped hierarchical porous carbon microtubes for effective CO2 capture. Chem. Eng. J. 2019, 358, 1507–1518. [Google Scholar] [CrossRef]
  20. 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]
  21. Wang, X.; He, T.; Hu, J.; Liu, M. The progress of nanomaterials for carbon dioxide captureviathe adsorption process. Environ. Sci. Nano. 2021, 8, 890–912. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Cano, Z.P.; Luo, D.; Dou, H.; Yu, A.; Chen, Z. Rational design of tailored porous carbon-based materials for CO2 capture. J. Mater. Chem. A Mater. 2019, 7, 20985–21003. [Google Scholar] [CrossRef]
  23. Sukor, N.R.; Shamsuddin, A.H.; Mahlia, T.M.I.; Isa, M.F.M. Techno-economic analysis of CO2 capture technologies in offshore natural gas field: Implications to carbon capture and storage in Malaysia. Processes 2020, 8, 350. [Google Scholar] [CrossRef]
  24. Micari, M.; Dakhchoune, M.; Agrawal, K.V. Techno-economic assessment of postcombustion carbon capture using high-performance nanoporous single-layer graphene membranes. J. Memb. Sci. 2021, 624, 119103. [Google Scholar] [CrossRef]
  25. John, J.M.; Alwi, S.R.W.; Omoregbe, D.I. Techno-economic analysis of carbon dioxide capture and utilisation analysis for an industrial site with fuel cell integration. J. Clean. Prod. 2021, 281, 124920. [Google Scholar] [CrossRef]
  26. Hao, G.P.; Li, W.C.; Qian, D.; Wang, G.H.; Zhang, W.P.; Zhang, T.; Wang, A.Q.; Schüth, F.; Bongard, H.J.; Lu, A.H. Structurally designed synthesis of mechanically stable poly(benzoxazine-co- resol)-based porous carbon monoliths and their application as high-performance CO2 capture sorbents. J. Am. Chem. Soc. 2011, 133, 11378–11388. [Google Scholar] [CrossRef] [PubMed]
  27. Ray, B.; Churipard, S.R.; Peter, S.C. An overview of the materials and methodologies for CO2capture under humid conditions. J. Mater. Chem. A Mater. 2021, 9, 26498–26527. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Yao, K.X.; Teng, B.; Zhang, T.; Han, Y. A perfluorinated covalent triazine-based framework for highly selective and water-tolerant CO2 capture. Energy Environ. Sci. 2013, 6, 3684–3692. [Google Scholar] [CrossRef]
  29. Masala, A.; Vitillo, J.G.; Mondino, G.; Grande, C.A.; Blom, R.; Manzoli, M.; Marshall, M.; Bordiga, S. CO2 capture in dry and wet conditions in UTSA-16 metal-organic framework. ACS Appl. Mater. Interfaces 2017, 9, 455–463. [Google Scholar] [CrossRef]
  30. Yang, J.; Yue, L.; Hu, X.; Wang, L.; Zhao, Y.; Lin, Y.; Sun, Y.; DaCosta, H.; Guo, L. Efficient CO2 Capture by Porous Carbons Derived from Coconut Shell. Energy Fuels 2017, 31, 4287–4293. [Google Scholar] [CrossRef]
  31. Xing, W.; Liu, C.; Zhou, Z.; Zhang, L.; Zhou, J.; Zhuo, S.; Yan, Z.; Gao, H.; Wang, G.; Qiao, S.Z. Superior CO2 uptake of N-doped activated carbon through hydrogen-bonding interaction. Energy Environ. Sci. 2012, 5, 7323–7327. [Google Scholar] [CrossRef]
  32. Sevilla, M.; Fuertes, A.B. Sustainable porous carbons with a superior performance for CO2 capture. Energy Environ. Sci. 2011, 4, 1765–1771. [Google Scholar] [CrossRef]
  33. To, J.W.; He, J.; Mei, J.; Haghpanah, R.; Chen, Z.; Kurosawa, T.; Chen, S.; Bae, W.G.; Pan, L.; Tok, J.B.H.; et al. Hierarchical N-Doped Carbon as CO2 Adsorbent with High CO2 Selectivity from Rationally Designed Polypyrrole Precursor. J. Am. Chem. Soc. 2016, 138, 1001–1009. [Google Scholar] [CrossRef] [PubMed]
  34. Creamer, A.E.; Gao, B. Carbon-based adsorbents for postcombustion CO2 capture: A critical review. Environ. Sci. Technol. 2016, 50, 7276–7289. [Google Scholar] [CrossRef]
  35. Yang, I.; Jung, M.; Kim, M.S.; Choi, D.; Jung, J.C. Physical and chemical activation mechanisms of carbon materials based on the microdomain model. J. Mater. Chem. A Mater. 2021, 9, 9815–9825. [Google Scholar] [CrossRef]
  36. Quan, C.; Zhou, Y.; Wang, J.; Wu, C.; Gao, N. Biomass-based carbon materials for CO2capture: A review. J. CO2 Util. 2023, 68, 102373. [Google Scholar] [CrossRef]
  37. Malini, K.; Selvakumar, D.; Kumar, N.S. Activated carbon from biomass: Preparation, factors improving basicity and surface properties for enhanced CO2capture capacity—A review. J. CO2 Util. 2023, 67, 102318. [Google Scholar] [CrossRef]
  38. Xing, W.; Liu, C.; Zhou, Z.; Zhou, J.; Wang, G.; Zhuo, S.; Xue, Q.; Song, L.; Yan, Z. Oxygen-containing functional group-facilitated CO2 capture by carbide-derived carbons. Nanoscale Res. Lett. 2014, 9, 189. [Google Scholar] [CrossRef]
  39. Khosrowshahi, M.S.; Abdol, M.A.; Mashhadimoslem, H.; Khakpour, E.; Emrooz, H.B.M.; Sadeghzadeh, S.; Ghaemi, A. The role of surface chemistry on CO2 adsorption in biomass-derived porous carbons by experimental results and molecular dynamics simulations. Sci. Rep. 2022, 12, 8917. [Google Scholar] [CrossRef]
  40. Wang, T.; Wang, X.; Hou, C.; Liu, J. Quaternary functionalized mesoporous adsorbents for ultra-high kinetics of CO2 capture from air. Sci. Rep. 2020, 10, 21429. [Google Scholar] [CrossRef]
  41. Yang, Y.B.; Hao, Q.; Müller-Plathe, F.; Böhm, M.C. Monte Carlo Simulations of SO2, H2S, and CO2 Adsorption in Charged Single-Walled Carbon Nanotube Arrays. J. Phys. Chem. C 2020, 124, 5838–5852. [Google Scholar] [CrossRef]
  42. Wang, R.; Xi, S.C.; Wang, D.Y.; Dou, M.; Dong, B. Defluorinated Porous Carbon Nanomaterials for CO2 Capture. ACS Appl. Nano Mater. 2021, 4, 10148–10154. [Google Scholar] [CrossRef]
  43. Hu, M.; He, J.; Zhao, Z.; Strobel, T.A.; Hu, W.; Yu, D.; Sun, H.; Liu, L.; Li, Z.; Ma, M.; et al. Compressed glassy carbon: An ultrastrong and elastic interpenetrating graphene network. Sci. Adv. 2017, 3, 3213. [Google Scholar] [CrossRef]
  44. Mao, H.; Tang, J.; Chen, J.; Wan, J.; Hou, K.; Peng, Y.; Halat, D.M.; Xiao, L.; Zhang, R.; Lv, X.; et al. Designing hierarchical nanoporous membranes for highly efficient gas adsorption and storage. Sci. Adv. 2020, 6, eabb0694. [Google Scholar] [CrossRef]
  45. Park, J.W.; Hwang, H.J.; Kang, H.J.; Bari, G.A.R.; Lee, T.G.; An, B.H.; Cho, S.Y.; Jun, Y.S. Hierarchical porous, N-containing carbon supports for high loading sulfur cathodes. Nanomaterials 2021, 11, 408. [Google Scholar] [CrossRef]
  46. Wang, R.; Jia, J.; Jin, Q.; Chen, H.; Liu, H.; Yin, Q.; Zhao, Z. Forming mechanism of coke microparticles from polymerization of aqueous organics during hydrothermal carbonization process of biomass. Carbon 2022, 192, 50–60. [Google Scholar] [CrossRef]
  47. Hao, W.; Björkman, E.; Lilliestråle, M.; Hedin, N. Activated carbons prepared from hydrothermally carbonized waste biomass used as adsorbents for CO2. Appl. Energy 2013, 112, 526–532. [Google Scholar] [CrossRef]
  48. Zhuang, X.; Liu, J.; Zhang, Q.; Wang, C.; Zhan, H.; Ma, L. A review on the utilization of industrial biowaste via hydrothermal carbonization. Renew. Sustain. Energy Rev. 2022, 154, 111877. [Google Scholar] [CrossRef]
  49. Silva, F.V.E.; Monteggia, L.O. Pyrolysis of algal biomass obtained from high-rate algae ponds applied to wastewater treatment. Front. Energy Res. 2015, 3, 31. [Google Scholar] [CrossRef]
  50. Wang, S.; Zhang, H.; Huang, H.; Xiao, R.; Li, R.; Zhang, Z. Influence of temperature and residence time on characteristics of biochars derived from agricultural residues: A comprehensive evaluation. Process Saf. Environ. Prot. 2020, 139, 218–229. [Google Scholar] [CrossRef]
  51. Falco, C.; Baccile, N.; Titirici, M.M. Morphological and structural differences between glucose, cellulose and lignocellulosic biomass derived hydrothermal carbons. Green Chem. 2011, 13, 3273–3281. [Google Scholar] [CrossRef]
  52. Kang, H.J.; Bari, G.A.R.; Lee, T.G.; Khan, T.T.; Park, J.W.; Hwang, H.J.; Cho, S.Y.; Jun, Y.S. Microporous carbon nanoparticles for lithium-sulfur batteries. Nanomaterials 2020, 10, 2012. [Google Scholar] [CrossRef]
  53. Sekirifa, M.L.; Hadj-Mahammed, M.; Pallier, S.; Baameur, L.; Richard, D.; Al-Dujaili, A.H. Preparation and characterization of an activated carbon from a date stones variety by physical activation with carbon dioxide. J. Anal. Appl. Pyrolysis. 2013, 99, 155–160. [Google Scholar] [CrossRef]
  54. Ogungbenro, A.E.; Quang, D.V.; Al-Ali, K.A.; Vega, L.F.; Abu-Zahra, M.R.M. Physical synthesis and characterization of activated carbon from date seeds for CO2 capture. J. Environ. Chem. Eng. 2018, 6, 4245–4252. [Google Scholar] [CrossRef]
  55. Singh, G.; Lakhi, K.S.; Sil, S.; Bhosale, S.V.; Kim, I.; Albahily, K.; Vinu, A. Biomass derived porous carbon for CO2 capture. Carbon 2019, 148, 164–186. [Google Scholar] [CrossRef]
  56. Kupgan, G.; Liyana-Arachchi, T.P.; Colina, C.M. NLDFT Pore Size Distribution in Amorphous Microporous Materials. Langmuir 2017, 33, 11138–11145. [Google Scholar] [CrossRef]
  57. Dantas, S.; Struckhoff, K.C.; Thommes, M.; Neimark, A.V. Pore size characterization of micro-mesoporous carbons using CO2 adsorption. Carbon 2021, 173, 842–848. [Google Scholar] [CrossRef]
  58. 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]
  59. Mehio, N.; Dai, S.; Jiang, D.E. Quantum mechanical basis for kinetic diameters of small gaseous molecules. J. Phys. Chem. A 2014, 118, 1150–1154. [Google Scholar] [CrossRef] [PubMed]
  60. Petrovic, B.; Gorbounov, M.; Soltani, S.M. Impact of Surface Functional Groups and Their Introduction Methods on the Mechanisms of CO2 Adsorption on Porous Carbonaceous Adsorbents. Carbon Capture Sci. Technol. 2022, 3, 100045. [Google Scholar] [CrossRef]
  61. Lim, G.; Lee, K.B.; Ham, H.C. Effect of N-Containing Functional Groups on CO2 Adsorption of Carbonaceous Materials: A Density Functional Theory Approach. J. Phys. Chem. C 2016, 120, 8087–8095. [Google Scholar] [CrossRef]
  62. Wang, Y.; Hu, X.; Hao, J.; Ma, R.; Guo, Q.; Gao, H.; Bai, H. Nitrogen and Oxygen Codoped Porous Carbon with Superior CO2 Adsorption Performance: A Combined Experimental and DFT Calculation Study. Ind. Eng. Chem. Res. 2019, 58, 13390–13400. [Google Scholar] [CrossRef]
  63. Parshetti, G.K.; Chowdhury, S.; Balasubramanian, R. Biomass derived low-cost microporous adsorbents for efficient CO2 capture. Fuel 2015, 148, 246–254. [Google Scholar] [CrossRef]
  64. Igalavithana, A.D.; Choi, S.W.; Shang, J.; Hanif, A.; Dissanayake, P.D.; Tsang, D.C.; Kwon, J.H.; Lee, K.B.; Ok, Y.S. Carbon dioxide capture in biochar produced from pine sawdust and paper mill sludge: Effect of porous structure and surface chemistry. Sci. Total Environ. 2020, 739, 139845. [Google Scholar] [CrossRef]
  65. Zhang, Z.; Zhou, J.; Xing, W.; Xue, Q.; Yan, Z.; Zhuo, S.; Qiao, S.Z. Critical role of small micropores in high CO2 uptake. Phys. Chem. Chem. Phys. 2013, 15, 2523–2529. [Google Scholar] [CrossRef]
  66. Sun, N.; Sun, C.; Liu, J.; Liu, H.; Snape, C.E.; Li, K. Surface-modified spherical activated carbon materials for pre-combustion carbon dioxide capture. RSC Adv. 2015, 5, 33681–33690. [Google Scholar] [CrossRef]
  67. Gao, A.; Guo, N.; Yan, M.; Li, M.; Wang, F.; Yang, R. Hierarchical porous carbon activated by CaCO3 from pigskin collagen for CO2 and H2 adsorption. Microporous Mesoporous Mater. 2018, 260, 172–179. [Google Scholar] [CrossRef]
  68. Zhang, X.; Elsayed, I.; Song, X.; Shmulsky, R.; Hassan, E.B. Microporous carbon nanoflakes derived from biomass cork waste for CO2 capture. Sci. Total Environ. 2020, 748, 142465. [Google Scholar] [CrossRef]
  69. Zhu, W.; Wang, Y.; Yao, F.; Wang, X.; Zheng, H.; Ye, G.; Cheng, H.; Wu, J.; Huang, H.; Ye, D. One-pot synthesis of N-doped petroleum coke-based microporous carbon for high-performance CO2 adsorption and supercapacitors. J. Environ. Sci. 2024, 139, 93–104. [Google Scholar] [CrossRef]
  70. Serafin, J.; Dziejarski, B.; Vendrell, X.; Kiełbasa, K.; Michalkiewicz, B. Biomass waste fern leaves as a material for a sustainable method of activated carbon production for CO2 capture. Biomass Bioenergy 2023, 175, 106880. [Google Scholar] [CrossRef]
  71. Zhang, Z.; Luo, D.; Lui, G.; Li, G.; Jiang, G.; Cano, Z.P.; Deng, Y.P.; Du, X.; Yin, S.; Chen, Y.; et al. In-situ ion-activated carbon nanospheres with tunable ultramicroporosity for superior CO2 capture. Carbon 2019, 143, 531–541. [Google Scholar] [CrossRef]
  72. Sajjadi, B.; Chen, W.Y.; Egiebor, N.O. A comprehensive review on physical activation of biochar for energy and environmental applications. Rev. Chem. Eng. 2019, 35, 735–776. [Google Scholar] [CrossRef]
  73. Alvim-Ferraz, M.C.M.; Gaspar, C.M.T.B. Micropore size distribution of activated carbons impregnated after carbonization. J. Porous Mater. 2003, 10, 47–55. [Google Scholar] [CrossRef]
  74. Prauchner, M.J.; Sapag, K.; Rodríguez-Reinoso, F. Tailoring biomass-based activated carbon for CH4 storage by combining chemical activation with H3PO4 or ZnCl2 and physical activation with CO2. Carbon 2016, 110, 138–147. [Google Scholar] [CrossRef]
  75. Kim, H.S.; Kang, M.S.; Yoo, W.C. Highly Enhanced Gas Sorption Capacities of N-Doped Porous Carbon Spheres by Hot NH3 and CO2 Treatments. J. Phys. Chem. C 2015, 119, 28512–28522. [Google Scholar] [CrossRef]
  76. Khuong, D.A.; Nguyen, H.N.; Tsubota, T. Activated carbon produced from bamboo and solid residue by CO2 activation utilized as CO2 adsorbents. Biomass Bioenergy 2021, 148, 106039. [Google Scholar] [CrossRef]
  77. Rajapaksha, A.U.; Vithanage, M.; Lee, S.S.; Seo, D.C.; Tsang, D.C.W.; Ok, Y.S. Steam activation of biochars facilitates kinetics and pH-resilience of sulfamethazine sorption. J. Soils Sediments 2016, 16, 889–895. [Google Scholar] [CrossRef]
  78. Fu, J.; Zhang, J.; Jin, C.; Wang, Z.; Wang, T.; Cheng, X.; Ma, C. Effects of temperature, oxygen and steam on pore structure characteristics of coconut husk activated carbon powders prepared by one-step rapid pyrolysis activation process. Bioresour Technol 2020, 310, 123413. [Google Scholar] [CrossRef]
  79. Muthmann, J.; Bläker, C.; Pasel, C.; Luckas, M.; Schledorn, C.; Bathen, D. Characterization of structural and chemical modifications during the steam activation of activated carbons. Microporous Mesoporous Mater. 2020, 309, 110549. [Google Scholar] [CrossRef]
  80. Heo, Y.J.; Park, S.J. A role of steam activation on CO2 capture and separation of narrow microporous carbons produced from cellulose fibers. Energy 2015, 91, 142–150. [Google Scholar] [CrossRef]
  81. Bardestani, R.; Kaliaguine, S. Steam activation and mild air oxidation of vacuum pyrolysis biochar. Biomass Bioenergy 2018, 108, 101–112. [Google Scholar] [CrossRef]
  82. Suliman, W.; Harsh, J.B.; Abu-Lail, N.I.; Fortuna, A.M.; Dallmeyer, I.; Garcia-Perez, M. Modification of biochar surface by air oxidation: Role of pyrolysis temperature. Biomass Bioenergy 2016, 85, 1–11. [Google Scholar] [CrossRef]
  83. Xiao, F.; Bedane, A.H.; Mallula, S.; Sasi, P.C.; Alinezhad, A.; Soli, D.; Hagen, Z.M.; Mann, M.D. Production of granular activated carbon by thermal air oxidation of biomass charcoal/biochar for water treatment in rural communities: A mechanistic investigation. Chem. Eng. J. Adv. 2020, 4, 100035. [Google Scholar] [CrossRef]
  84. Petrovic, B.; Gorbounov, M.; Soltani, S.M. Influence of surface modification on selective CO2 adsorption: A technical review on mechanisms and methods. Microporous Mesoporous Mater. 2021, 312, 110751. [Google Scholar] [CrossRef]
  85. Zaker, A.; Hammouda, S.B.; Sun, J.; Wang, X.; Li, X.; Chen, Z. Carbon-based materials for CO2 capture: Their production, modification and performance. J. Environ. Chem. Eng. 2023, 11, 109741. [Google Scholar] [CrossRef]
  86. Zhou, J.; Li, Z.; Xing, W.; Zhu, T.; Shen, H.; Zhuo, S. N-doped microporous carbons derived from direct carbonization of K+ exchanged meta-aminophenol-formaldehyde resin for superior CO2 sorption. Chem. Commun. 2015, 51, 4591–4594. [Google Scholar] [CrossRef]
  87. Zhou, Z.Q.J.; Li, Z.; Xing, W.; Shen, H.; Bi, X.; Zhu, T.; Zhuo, S. A New Approach to Tuning Carbon Ultramicropore Size at Sub-Angstrom Level for Maximizing Specific Capacitance and CO2 Uptake. Adv. Funct. Mater. 2016, 26, 7955–7964. [Google Scholar] [CrossRef]
  88. Itoi, H.; Nishihara, H.; Kogure, T.; Kyotani, T. Three-dimensionally arrayed and mutually connected 1.2-nm nanopores for high-performance electric double layer capacitor. J. Am. Chem. Soc. 2011, 133, 1165–1167. [Google Scholar] [CrossRef]
  89. Zhu, S.; Zhao, N.; Li, J.; Deng, X.; Sha, J.; He, C. Hard-template synthesis of three-dimensional interconnected carbon networks: Rational design, hybridization and energy-related applications. Nano Today 2019, 29, 100796. [Google Scholar] [CrossRef]
  90. Joseph, S.; Singh, G.; Lee, J.M.; Yu, X.; Breese, M.B.; Ruban, S.M.; Bhargava, S.K.; Yi, J.; Vinu, A. Hierarchical carbon structures from soft drink for multi-functional energy applications of Li-ion battery, Na-ion battery and CO2 capture. Carbon 2023, 210, 118085. [Google Scholar] [CrossRef]
  91. Vilian, A.E.; Song, J.Y.; Lee, Y.S.; Hwang, S.K.; Kim, H.J.; Jun, Y.S.; Huh, Y.S.; Han, Y.K. Salt-templated three-dimensional porous carbon for electrochemical determination of gallic acid. Biosens. Bioelectron. 2018, 117, 597–604. [Google Scholar] [CrossRef]
  92. Deng, X.; Zhao, B.; Zhu, L.; Shao, Z. Molten salt synthesis of nitrogen-doped carbon with hierarchical pore structures for use as high-performance electrodes in supercapacitors. Carbon 2015, 93, 48–58. [Google Scholar] [CrossRef]
  93. Liu, X.; Antonietti, M. Moderating black powder chemistry for the synthesis of doped and highly porous graphene nanoplatelets and their use in electrocatalysis. Adv. Mater. 2013, 25, 6284–6290. [Google Scholar] [CrossRef]
  94. Liu, X.; Giordano, C.; Antonietti, M. A facile molten-salt route to graphene synthesis. Smal 2014, 10, 193–200. [Google Scholar] [CrossRef] [PubMed]
  95. Fechler, N.; Fellinger, T.P.; Antonietti, M. ‘salt templating’: A simple and sustainable pathway toward highly porous functional carbons from ionic liquids. Adv. Mater. 2013, 25, 75–79. [Google Scholar] [CrossRef]
  96. Yang, Z.; Xia, Y.; Sun, X.; Mokaya, R. Preparation and hydrogen storage properties of zeolite-templated carbon materials nanocast via chemical vapor deposition: Effect of the zeolite template and nitrogen doping. J. Phys. Chem. B 2006, 110, 18424–18431. [Google Scholar] [CrossRef] [PubMed]
  97. Asasian-Kolur, N.; Sharifian, S.; Haddadi, B.; Jordan, C.; Harasek, M. Ordered Porous Carbon Preparation by Hard Templating Approach for Hydrogen Adsorption Application; Springer: Berlin/Heidelberg, Germany, 2023. [Google Scholar] [CrossRef]
  98. Wang, J.; Wang, Y.; Hu, H.; Yang, Q.; Cai, J. From metal-organic frameworks to porous carbon materials: Recent progress and prospects from energy and environmental perspectives. Nanoscale 2020, 12, 4238–4268. [Google Scholar] [CrossRef]
  99. Xu, X.; Xu, C.; Liu, J.; Jin, R.; Luo, X.; Shu, C.; Chen, H.; Guo, C.; Xu, L.; Si, Y. The synergistic effect of ‘soft-hard template’ to in situ regulate mass transfer and defective sites of doped-carbon nanostructures for catalysis of oxygen reduction. J. Alloys Compd. 2023, 939, 168782. [Google Scholar] [CrossRef]
  100. Lee, J.; Han, S.; Hyeon, T. Synthesis of new nanoporous carbon materials using nanostructured silica materials as templates. J. Mater. Chem. 2004, 14, 478–486. [Google Scholar] [CrossRef]
  101. Rehman, A.; Heo, Y.J.; Nazir, G.; Park, S.J. Solvent-free, one-pot synthesis of nitrogen-tailored alkali-activated microporous carbons with an efficient CO2 adsorption. Carbon 2021, 172, 71–82. [Google Scholar] [CrossRef]
  102. Shi, J.; Xu, J.; Cui, H.; Yan, N.; Zou, J.; Liu, Y.; You, S. Synthesis of highly porous N-doped hollow carbon nanospheres with a combined soft template-chemical activation method for CO2 capture. Ind. Crops Prod. 2023, 280, 115952. [Google Scholar] [CrossRef]
  103. Chuenchom, L.; Kraehnert, R.; Smarsly, B.M. Recent progress in soft-templating of porous carbon materials. Soft Matter 2012, 8, 10801–10812. [Google Scholar] [CrossRef]
  104. Xu, Z.; Wu, Z.; Chi, J.; Lei, E.; Liu, Y.; Yin, Y.; Yang, Z.; Ma, C.; Li, W.; Luo, S.; et al. Soft-template hydrothermal synthesis of N and B co-doped walnut-shaped porous carbon spheres with hydrophilic surfaces for supercapacitors. Appl. Surf. Sci. 2023, 638, 158016. [Google Scholar] [CrossRef]
  105. Kang, H.J.; Huh, Y.S.; Im, W.B.; Jun, Y.S. Molecular cooperative assembly-mediated synthesis of ultra-high-performance hard carbon anodes for dual-carbon sodium hybrid capacitors. ACS Nano 2019, 13, 11935–11946. [Google Scholar] [CrossRef] [PubMed]
  106. Guan, L.; Hu, H.; Teng, X.L.; Zhu, Y.F.; Zhang, Y.L.; Chao, H.X.; Yang, H.; Wang, X.S.; Wu, M.B. Templating synthesis of porous carbons for energy-related applications: A review. New Carbon Mater. 2022, 37, 25–45. [Google Scholar] [CrossRef]
  107. Jun, Y.S.; Lee, E.Z.; Wang, X.; Hong, W.H.; Stucky, G.D.; Thomas, A. From melamine-cyanuric acid supramolecular aggregates to carbon nitride hollow spheres. Adv. Funct. Mater. 2013, 23, 3661–3667. [Google Scholar] [CrossRef]
  108. Nicolae, S.A.; Szilágyi, Á.; Titirici, M.M. Soft templating production of porous carbon adsorbents for CO2 and H2S capture. Carbon 2020, 169, 193–204. [Google Scholar] [CrossRef]
  109. Zhang, Z.; Chen, Y.; Wang, P.; Wang, Z.; Zuo, C.; Chen, W.; Ao, T. Facile fabrication of N-doped hierarchical porous carbons derived from soft-templated ZIF-8 for enhanced adsorptive removal of tetracycline hydrochloride from water. J. Hazard Mater. 2022, 423, 127103. [Google Scholar] [CrossRef]
  110. Chen, L.; Yuan, Y.; Zhang, D.; Lin, Z.; Lin, J.; Li, S.; Guo, S. Construction of 3D hierarchical honeycomb macro/meso/micro-porous carbon with soft and hard templates for high-performance sodium-ion batteries. Mater. Lett. 2023, 334, 133737. [Google Scholar] [CrossRef]
  111. Bin, A.A.; Binti, J.H. Waste Recycling Technologies for Nanomaterials Manufacturing: Manufacturing of Nanoalumina by Recycling of Aluminium Cans Waste Aiman; Springer: Cham, Switzerland, 2021; Available online: https://link.springer.com/book/10.1007/978-3-030-68031-2 (accessed on 8 July 2023).
  112. Xu, S.; Niu, M.; Zhao, G.; Ming, S.; Li, X.; Zhu, Q.; Ding, L.X.; Kim, M.; Alothman, A.A.; Mushab, M.S.S.; et al. Size control and electronic manipulation of Ru catalyst over B, N co-doped carbon network for high-performance hydrogen evolution reaction. Nano Res. 2023, 16, 6212–6219. [Google Scholar] [CrossRef]
  113. Jin, X.; Wang, X.; Liu, Y.; Kim, M.; Cao, M.; Xie, H.; Liu, S.; Wang, X.; Huang, W.; Nanjundan, A.K.; et al. Nitrogen and Sulfur Co-Doped Hierarchically Porous Carbon Nanotubes for Fast Potassium Ion Storage. Small 2022, 18, 42. [Google Scholar] [CrossRef]
  114. Emran, M.Y.; Shenashen, M.A.; El-Safty, S.A.; Selim, M.M. Design of porous S-doped carbon nanostructured electrode sensor for sensitive and selective detection of guanine from DNA samples. Microporous Mesoporous Mater. 2021, 320, 111097. [Google Scholar] [CrossRef]
  115. Khosropour, H.; Rezaei, B.; Alinajafi, H.A.; Ensafi, A.A. Electrochemical sensor based on glassy carbon electrode modified by polymelamine formaldehyde/graphene oxide nanocomposite for ultrasensitive detection of oxycodone. Microchimica Acta 2021, 188, 1. [Google Scholar] [CrossRef]
  116. Emran, M.Y.; Shenashen, M.A.; Elmarakbi, A.; Selim, M.M.; El-Safty, S.A. Nitrogen-doped carbon hollow trunk-like structure as a portable electrochemical sensor for noradrenaline detection in neuronal cells. Anal. Chim. Acta 2022, 1192, 339380. [Google Scholar] [CrossRef]
  117. Wu, D.; Yang, Y.; Liu, J.; Zheng, Y. Plasma-Modified N/O-Doped Porous Carbon for CO2Capture: An Experimental and Theoretical Study. Energy Fuels 2020, 34, 6077–6084. [Google Scholar] [CrossRef]
  118. Wu, D.; Liu, J.; Yang, Y.; Zheng, Y. Nitrogen/Oxygen Co-Doped Porous Carbon Derived from Biomass for Low-Pressure CO2Capture. Ind. Eng. Chem. Res. 2020, 59, 14055–14063. [Google Scholar] [CrossRef]
  119. Ashourirad, B.; Arab, P.; Islamoglu, T.; Cychosz, K.A.; Thommes, M.; El-Kaderi, H.M. A cost-effective synthesis of heteroatom-doped porous carbons as efficient CO2 sorbents. J. Mater. Chem. A Mater. 2016, 4, 14693–14702. [Google Scholar] [CrossRef]
  120. Dou, X.; Hasa, I.; Saurel, D.; Vaalma, C.; Wu, L.; Buchholz, D.; Bresser, D.; Komaba, S.; Passerini, S. Hard carbons for sodium-ion batteries: Structure, analysis, sustainability, and electrochemistry. Mater. Today 2019, 23, 87–104. [Google Scholar] [CrossRef]
  121. Klepel, O.; Hunger, B. Temperature-programmed desorption (TPD) of carbon dioxide on alkali-metal cation-exchanged faujasite type zeolites. J. Therm. Anal. Calorim. 2005, 80, 201–206. [Google Scholar] [CrossRef]
  122. Jagiello, J.; Chojnacka, A.; Pourhosseini, S.E.M.; Wang, Z.; Beguin, F. A dual shape pore model to analyze the gas adsorption data of hierarchical micro-mesoporous carbons. Carbon 2021, 178, 113–124. [Google Scholar] [CrossRef]
  123. Hotová, G.; Slovák, V.; Soares, O.S.G.P.; Figueiredo, J.L.; Pereira, M.F.R. Oxygen surface groups analysis of carbonaceous samples pyrolysed at low temperature. Carbon 2018, 134, 255–263. [Google Scholar] [CrossRef]
  124. Wu, R.; Ye, Q.; Wu, K.; Wang, L.; Dai, H. Highly efficient CO2 adsorption of corn kernel-derived porous carbon with abundant oxygen functional groups. J. CO2 Util. 2021, 51, 101620. [Google Scholar] [CrossRef]
  125. Xiao, J.; Wang, Y.; Zhang, T.C.; Ouyang, L.; Yuan, S. Phytic acid-induced self-assembled chitosan gel-derived N, P—Co-doped porous carbon for high-performance CO 2 capture and supercapacitor. J. Power Sources 2022, 517, 230727. [Google Scholar] [CrossRef]
  126. Maklavany, D.M.; Rouzitalab, Z.; Amini, A.M.; Askarieh, M.; Silvestrelli, P.L.; Seif, A.; Orooji, Y.; Rashidi, A. One-step approach to Quaternary (B, N, P, S)-Doped hierarchical porous carbon derived from Quercus Brantii for highly selective and efficient CO2 Capture: A combined experimental and extensive DFT study. Chem. Eng. J. 2023, 453, 139950. [Google Scholar] [CrossRef]
  127. Ma, X.; Xu, W.; Su, R.; Shao, L.; Zeng, Z.; Li, L.; Wang, H. Insights into CO2 capture in porous carbons from machine learning, experiments and molecular simulation. Sep. Purif. Technol. 2023, 306, 122521. [Google Scholar] [CrossRef]
  128. Khodabakhshi, S.; Kiani, S.; Niu, Y.; White, A.O.; Suwaileh, W.; Palmer, R.E.; Barron, A.R.; Andreoli, E. Facile and environmentally friendly synthesis of ultramicroporous carbon spheres: A significant improvement in CVD method. Carbon 2021, 171, 426–436. [Google Scholar] [CrossRef]
  129. Li, J.; Zhang, W.; Bao, A. Design of hierarchically structured porous boron/nitrogen codoped carbon materials with excellent performance for CO2capture. Ind. Eng. Chem. Res. 2021, 60, 2710–2718. [Google Scholar] [CrossRef]
  130. Song, C.; Ye, W.; Liu, Y.; Huang, H.; Zhang, H.; Lin, H.; Lu, R.; Zhang, S. Facile preparation of porous carbon derived from industrial biomass waste as an efficient CO2adsorbent. ACS Omega 2020, 5, 28255–28263. [Google Scholar] [CrossRef]
  131. Li, L.; Wang, X.F.; Zhong, J.J.; Qian, X.; Song, S.L.; Zhang, Y.G.; Li, D.H. Nitrogen-enriched porous polyacrylonitrile-based carbon fibers for CO2 Capture. Ind. Eng. Chem. Res. 2018, 57, 11608–11616. [Google Scholar] [CrossRef]
  132. Song, C.; Liu, M.; Ye, W.; Liu, Y.; Zhang, H.; Lu, R.; Zhang, S. Nitrogen-Containing Porous Carbon for Highly Selective and Efficient CO2 Capture. Energy Fuels 2019, 33, 12601–12609. [Google Scholar] [CrossRef]
  133. Xiong, L.; Wang, X.F.; Li, L.; Jin, L.; Zhang, Y.G.; Song, S.L.; Liu, R.P. Nitrogen-Enriched Porous Carbon Fiber as a CO2 Adsorbent with Superior CO2 Selectivity by Air Activation. Energy Fuels 2019, 33, 12558–12567. [Google Scholar] [CrossRef]
  134. Mahurin, S.M.; Fulvio, F.; Hillesheim, C.; Nelson, K.M.; Veith, G.M.; Dai, S. Directed synthesis of nanoporous carbons from task-specific ionic liquid precursors for the adsorption of CO2. ChemSusChem 2014, 7, 3284–3289. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, X.F.; Xiong, L.; Zhong, J.J.; Jin, L.; Yan, J.L.; Mu, B.; Zhang, Y.G.; Song, S.L. Nitrogen-Containing Porous Carbon Fibers Prepared from Polyimide Fibers for CO2Capture. Ind. Eng. Chem. Res. 2020, 59, 18106–18114. [Google Scholar] [CrossRef]
  136. Pevida, C.; Drage, T.C.; Snape, C.E. Silica-templated melamine-formaldehyde resin derived adsorbents for CO2 capture. Carbon 2008, 46, 1464–1474. [Google Scholar] [CrossRef]
  137. Li, D.; Chen, Y.; Zheng, M.; Zhao, H.; Zhao, Y.; Sun, Z. Hierarchically Structured Porous Nitrogen-Doped Carbon for Highly Selective CO2 Capture. ACS Sustain. Chem. Eng. 2016, 4, 298–304. [Google Scholar] [CrossRef]
  138. Estevez, L.; Barpaga, D.; Zheng, J.; Sabale, S.; Patel, R.L.; Zhang, J.G.; McGrail, B.P.; Motkuri, R.K. Hierarchically Porous Carbon Materials for CO2 Capture: The Role of Pore Structure. Ind. Eng. Chem. Res. 2018, 57, 1262–1268. [Google Scholar] [CrossRef]
  139. Hong, L.; Ju, S.; Liu, X.; Zhuang, Q.; Zhan, G.; Yu, X. Highly Selective CO2 Uptake in Novel Fishnet-like Polybenzoxazine-Based Porous Carbon. Energy Fuels 2019, 33, 11454–11464. [Google Scholar] [CrossRef]
  140. Wahab, M.A.; Na, J.; Masud, M.K.; Hossain, M.S.A.; Alothman, A.A.; Abdala, A. Nanoporous carbon nitride with a high content of inbuilt N site for the CO2 capture. J. Hazard Mater. 2021, 408, 124843. [Google Scholar] [CrossRef]
  141. Zeeshan, M.; Yalcin, K.; Oztuna, F.E.S.; Unal, U.; Keskin, S.; Uzun, A. A new class of porous materials for efficient CO2 separation: Ionic liquid/graphene aerogel composites. Carbon 2021, 171, 79–87. [Google Scholar] [CrossRef]
  142. Guo, Q.; Chen, C.; Li, Z.; Li, X.; Wang, H.; Feng, N.; Wan, H.; Guan, G. Controllable construction of N-enriched hierarchically porous carbon nanosheets with enhanced performance for CO2 capture. Chem. Eng. J. 2019, 371, 414–423. [Google Scholar] [CrossRef]
  143. Srinivas, G.; Krungleviciute, V.; Guo, Z.X.; Yildirim, T. Exceptional CO2 capture in a hierarchically porous carbon with simultaneous high surface area and pore volume. Energy Environ. Sci. 2014, 7, 335–342. [Google Scholar] [CrossRef]
  144. Kumbhar, D.; Palliyarayil, A.; Reghu, D.; Shrungar, D.; Umapathy, S.; Sil, S. Rapid discrimination of porous bio-carbon derived from nitrogen rich biomass using Raman spectroscopy and artificial intelligence methods. Carbon 2021, 178, 792–802. [Google Scholar] [CrossRef]
  145. Ma, C.; Bai, J.; Hu, X.; Jiang, Z.; Wang, L. Nitrogen-doped porous carbons from polyacrylonitrile fiber as effective CO2 adsorbents. J. Environ. Sci. 2023, 125, 533–543. [Google Scholar] [CrossRef] [PubMed]
  146. Alloush, A.M.; Abdulghani, H.; Amasha, H.A.; Saleh, T.A.; Al Hamouz, O.C.S. Microwave-assisted synthesis of novel porous organic polymers for effective selective capture of CO2. J. Ind. Eng. Chem. 2022, 113, 215–225. [Google Scholar] [CrossRef]
  147. Gou, J.; Liu, C.; Lin, J.; Yu, C.; Fang, Y.; Liu, Z.; Guo, Z.; Tang, C.; Huang, Y. Densification and pelletization of porous boron nitride fibers for effective CO2 adsorption. Ceram. Int. 2022, 48, 11636–11643. [Google Scholar] [CrossRef]
  148. Dhoke, C.; Zaabout, A.; Cloete, S.; Amini, S. Review on reactor configurations for adsorption-based CO2 capture. Ind. Eng. Chem. Res. 2021, 60, 3779–3798. [Google Scholar] [CrossRef]
  149. Zhang, R.; Li, Z.; Zeng, L.; Wang, F. Pressure drop in honeycomb adsorption filters filled with granular activated carbon. Powder Technol. 2021, 393, 550–558. [Google Scholar] [CrossRef]
  150. Rezaei, F.; Grahn, M. Thermal management of structured adsorbents in CO 2 capture processes. Ind. Eng. Chem. Res. 2012, 51, 4025–4034. [Google Scholar] [CrossRef]
  151. Rezaei, F.; Webley, P. Structured adsorbents in gas separation processes. Sep. Purif. Technol. 2010, 70, 243–256. [Google Scholar] [CrossRef]
  152. Rezaei, F.; Webley, P. Optimum structured adsorbents for gas separation processes. Chem. Eng. Sci. 2009, 64, 5182–5191. [Google Scholar] [CrossRef]
  153. Akhtar, F.; Andersson, L.; Ogunwumi, S.; Hedin, N.; Bergström, L. Structuring adsorbents and catalysts by processing of porous powders. J. Eur. Ceram. Soc. 2014, 34, 1643–1666. [Google Scholar] [CrossRef]
  154. Borchardt, L.; Michels, N.L.; Nowak, T.; Mitchell, S.; Pérez-Ramírez, J. Structuring zeolite bodies for enhanced heat-transfer properties. Microporous Mesoporous Mater. 2015, 208, 196–202. [Google Scholar] [CrossRef]
  155. Tang, S.H.; Zaini, M.A.A. Development of activated carbon pellets using a facile low-cost binder for effective malachite green dye removal. J. Clean. Prod. 2020, 253, 119970. [Google Scholar] [CrossRef]
  156. Cousin-Saint-Remi, J.; Van der Perre, S.; Segato, T.; Delplancke, M.P.; Goderis, S.; Terryn, H.; Baron, G.; Denayer, J. Highly Robust MOF Polymeric Beads with a Controllable Size for Molecular Separations. ACS Appl Mater Interfaces 2019, 11, 13694–13703. [Google Scholar] [CrossRef]
  157. Valizadeh, B.; Nguyen, T.N.; Smit, B.; Stylianou, K.C. Porous Metal–Organic Framework@Polymer Beads for Iodine Capture and Recovery Using a Gas-Sparged Column. Adv. Funct. Mater. 2018, 28, 1801596. [Google Scholar] [CrossRef]
  158. de Almeida Moreira, B.R.; Cruz, V.H.; Pérez, J.F.; da Silva Viana, R. Production of pellets for combustion and physisorption of CO2 from hydrothermal carbonization of food waste—Part I: High-performance solid biofuels. J. Clean. Prod. 2021, 319, 128695. [Google Scholar] [CrossRef]
  159. Kim, Y.J.S. Adsorption of Carbon Dioxide using Pelletized AC with Amine impregnation. J. Korean Oil Chem. Soc. 2013, 30, 88–89. [Google Scholar] [CrossRef]
  160. Manohara, G.V.; Maroto-Valer, M.M.; Garcia, S. Binder free novel synthesis of structured hybrid mixed metal oxides (MMOs) for high temperature CO2 capture. Chem. Eng. J. 2021, 415, 128881. [Google Scholar] [CrossRef]
Figure 1. CO2 concentration monitoring by decade (1960~present) by ESRL’s (Earth System Research Laboratories) Global Monitoring Laboratory (GML) at Mauna Loa observatory in Hawaii.
Figure 1. CO2 concentration monitoring by decade (1960~present) by ESRL’s (Earth System Research Laboratories) Global Monitoring Laboratory (GML) at Mauna Loa observatory in Hawaii.
Surfaces 06 00023 g001
Figure 2. A schematic representation elucidating the mechanisms of carbon formation.
Figure 2. A schematic representation elucidating the mechanisms of carbon formation.
Surfaces 06 00023 g002
Figure 3. Graphical illustration that shows the CO2 capture mechanisms on the carbon surface [64]. Adapted with permission from ref., copyright (2020) Elsevier Ltd.
Figure 3. Graphical illustration that shows the CO2 capture mechanisms on the carbon surface [64]. Adapted with permission from ref., copyright (2020) Elsevier Ltd.
Surfaces 06 00023 g003
Figure 4. Synthesis strategy of ultra-microporous carbon nanosphere [71]. Adapted with permission from ref., copyright (2020) Elsevier Ltd.
Figure 4. Synthesis strategy of ultra-microporous carbon nanosphere [71]. Adapted with permission from ref., copyright (2020) Elsevier Ltd.
Surfaces 06 00023 g004
Figure 5. Schematic diagram showing the mechanisms of pore formation within the carbon structure during the physical activation process; (ad) activation reaction time: 0, 1, 3, and 5 s, consecutively [78]. Adapted with permission from ref., copyright (2019) Elsevier Ltd.
Figure 5. Schematic diagram showing the mechanisms of pore formation within the carbon structure during the physical activation process; (ad) activation reaction time: 0, 1, 3, and 5 s, consecutively [78]. Adapted with permission from ref., copyright (2019) Elsevier Ltd.
Surfaces 06 00023 g005
Figure 6. A schematic show of carbon surface activation by steam [80]. Adapted with permission from ref., copyright (2015) Elsevier Ltd.
Figure 6. A schematic show of carbon surface activation by steam [80]. Adapted with permission from ref., copyright (2015) Elsevier Ltd.
Surfaces 06 00023 g006
Figure 7. Diagram illustrating the synthesis process of interconnected hierarchical carbon using the KIT-6 hard template and ZnCl2 activation. [90]. Adapted with permission from ref., copyright (2023) Elsevier Ltd.
Figure 7. Diagram illustrating the synthesis process of interconnected hierarchical carbon using the KIT-6 hard template and ZnCl2 activation. [90]. Adapted with permission from ref., copyright (2023) Elsevier Ltd.
Surfaces 06 00023 g007
Figure 8. Hierarchical porous carbon synthesis from glucose utilizing LiBr/KBr molten salt and an organic reactive template of melamine–cyanuric acid complex [1]. Adapted with permission from ref., copyright (2023) Springer Nature.
Figure 8. Hierarchical porous carbon synthesis from glucose utilizing LiBr/KBr molten salt and an organic reactive template of melamine–cyanuric acid complex [1]. Adapted with permission from ref., copyright (2023) Springer Nature.
Surfaces 06 00023 g008
Figure 9. (a) CO2 adsorption energy on carbon surfaces with N/O functional groups. (b) Charge density difference of CO2 molecules on different carbon surfaces. C = gray; H = white; O = red; N = blue. Reprinted (adapted) with permission from [118] {Ind. Eng. Chem. Res. 2020, 59, 31, 14055–14063}. Copyright (2020) American Chemical Society.
Figure 9. (a) CO2 adsorption energy on carbon surfaces with N/O functional groups. (b) Charge density difference of CO2 molecules on different carbon surfaces. C = gray; H = white; O = red; N = blue. Reprinted (adapted) with permission from [118] {Ind. Eng. Chem. Res. 2020, 59, 31, 14055–14063}. Copyright (2020) American Chemical Society.
Surfaces 06 00023 g009
Figure 10. CO2-TPD profiles of carbon samples (a) without or (b) with CO2 pre-adsorption, (c) deconvolution of CO2-TPD profile of MG-Br-600 with CO2 pre-adsorption, and (d) multi-cycle profiles of CO2-TPD results of MG-Br-600 [1]. Adapted with permission from ref., copyright (2023) Springer Nature.
Figure 10. CO2-TPD profiles of carbon samples (a) without or (b) with CO2 pre-adsorption, (c) deconvolution of CO2-TPD profile of MG-Br-600 with CO2 pre-adsorption, and (d) multi-cycle profiles of CO2-TPD results of MG-Br-600 [1]. Adapted with permission from ref., copyright (2023) Springer Nature.
Surfaces 06 00023 g010
Figure 11. Schematic diagram of synthesis of phytic acid–induced self–assembled chitosan gel-derived N–-,P–co–doped porous carbon [125]. Adapted with permission from ref., copyright (2023) Elsevier Ltd.
Figure 11. Schematic diagram of synthesis of phytic acid–induced self–assembled chitosan gel-derived N–-,P–co–doped porous carbon [125]. Adapted with permission from ref., copyright (2023) Elsevier Ltd.
Surfaces 06 00023 g011
Figure 12. (a) CO2 adsorption isotherms; AC: activated carbon; BAC, NAC, PAC, and SAC are B–, N–, P–, and S–doped activated carbon, consecutively [126]; (b) CO2 adsorption isotherms; UC800: carbon from urea activated by KOH; ZNC600: carbon from urea activated by ZIF8; NPC600: hydrochar from glucose, ethylenediamine; (c) CO2 adsorption isotherms of different pore size; and (d) CO2 adsorption isotherms of different oxygen and nitrogen groups with 0.5 nm pore size at 25 °C [127]. Adapted with permission from refs., copyright (2023) Elsevier Ltd.
Figure 12. (a) CO2 adsorption isotherms; AC: activated carbon; BAC, NAC, PAC, and SAC are B–, N–, P–, and S–doped activated carbon, consecutively [126]; (b) CO2 adsorption isotherms; UC800: carbon from urea activated by KOH; ZNC600: carbon from urea activated by ZIF8; NPC600: hydrochar from glucose, ethylenediamine; (c) CO2 adsorption isotherms of different pore size; and (d) CO2 adsorption isotherms of different oxygen and nitrogen groups with 0.5 nm pore size at 25 °C [127]. Adapted with permission from refs., copyright (2023) Elsevier Ltd.
Surfaces 06 00023 g012
Figure 13. A schematic representation comparing powder and fixed-shape state carbonization strategies for the effective utilization of the gas template from the carbon degradation products.
Figure 13. A schematic representation comparing powder and fixed-shape state carbonization strategies for the effective utilization of the gas template from the carbon degradation products.
Surfaces 06 00023 g013
Figure 14. Charting a potential route: pathways to crafting pelletized porous carbon.
Figure 14. Charting a potential route: pathways to crafting pelletized porous carbon.
Surfaces 06 00023 g014
Table 1. Performance comparison of CO2 uptake, kinetics, heat of adsorption, and selectivity in carbon materials from different sources and under various modification processes for carbon capture.
Table 1. Performance comparison of CO2 uptake, kinetics, heat of adsorption, and selectivity in carbon materials from different sources and under various modification processes for carbon capture.
Activation ProcessCarbon MaterialsUptake (1 bar)
[mmol g−1]
Kinetics/Heat of Adsorption (Qst)
[kJ mol−1]
Selectivity
CO2/N2 (15:85)
Ref.
MS/organic templateN-/O-rich multi-hierarchical porous C3.8 (0 °C),
2.9 (25 °C)
Kinetics
(Initial kpi, 0.2 mmol g−1 min−0.5)
26
43 (25 °C),
31 (0 °C)
[1]
Activation freeC (spherical)2.9 (25 °C),
4 (0 °C)
27.5–29.3CO2/N2 (15/85)
30 (25 °C)
[128]
Template freeB-/N-co-doped C2.1 (30 °C)28[129]
Chemical activationC from industrial biomass4.2 (25 °C),
6.6 (0 °C),
1.3 (0.15 bar)
34–1827[130]
Chemical activationDefluorinated porous C5.0 (25 °C),
8.8 (0 °C)
27.3 (zero coverage)23 (25 °C),
22 (0 °C)
[42]
Template freeN-containing C (PAN)2.4 (25 °C)Simulated flue gas CO2/N2 (10/90) 0.8 mmol g−1 (25 °C)[131]
Chemical activationN-containing porous C3.7 (25 °C),
6.2 (0 °C),
3.2 (0 °C)
15–3618 kJ mol–1 (25 °C)
129 kJ mol–1 (25 °C)
[132]
Air activationN-containing porous C fiber2.2 (25 °C)26.6–30.8183[133]
Template freeNanoporous C2.7 (25 °C),
4.0 (0 °C)
31.530[134]
Chemical activationN-containing porous C fiber5.0 (25 °C),
1.5 (0.15 bar)
24.6–25.524[135]
Silica templatePorous C2.2 (25 °C)31.8[136]
Chemical activationPorous N-containing C2.7 (25 °C),
3.8 (25 °C)
~36CO2/N2 (10/90) 134[137]
Freeze drying/plasma treatmentN-/O-containing porous C0.9 (30 °C)15[118]
Ice/silica/CO2 templatingN-/O-containing porous C3.7 (25 °C)[138]
Chemical activationPolybenzooxazine-based porous C8.4 (25 °C)32–35CO2/N2 (10/90)
25 (25 °C),
34 (0 °C)
[139]
Hard templateN-containing carbon nitride2.0 (25 °C),
2.5 (0 °C)
-[140]
Template freeIonic liquid/graphene aerogel0.2 (0 °C)CO2/CH4 (ideal selectivity)
120 (25 °C),
1 mbar
[141]
Metal ion activationN-containing porous C nanosheet2.5 (25 °C)17.5 (25 °C)[142]
Hard templatePorous carbon from MOF~2.5 (27 °C)[143]
Salt templateN-containing C3.3 (25 °C)––-[144]
Template freeN-containing C monolith3.3 (25 °C),
5 (0 °C)
~20CO2/N2 (14/86)
16 min breakthrough time
(25 °C)
[26]
Chemical activationN-containing porous C4.0 (25 °C),
6.0 (0 °C)
~25CO2/N2 (10/90)
19 (25 °C)
[145]
Hard templateN-based porous polymer1.0 (25 °C),
1.6 (0 °C)
63 (25 °C, 1 bar)[146]
Abbreviations: Carbon is denoted as C and Molten salt as MS in the Table.
Table 2. A comparison of the different activation processes.
Table 2. A comparison of the different activation processes.
Carbon Structure ModificationAdvantagesDisadvantages
Chemical activationProvides ultra pores, micropores, lower activation temperatureExtra steps of washing to remove corrosive chemicals
Physical activationProvides micro-/meso-/macropores, readily utilized on carbonizationCannot provide shaped carbon, higher activation temperature
Metal ion activationUltra-to-super-microporesLimited carbon source, extra process of washing to remove metal
Hard templateMechanical and structural stability, micro-/mesoporesExtra process to remove template, etching, time-consuming, can damage carbon structure due to vigorous acid or base treatment to remove template
Soft templateSelf-removal with carbonization, provides multi-shaped carbonOnly meso-/macroporous carbon synthesis, template needs to remain stable during thermal process, template preparation process requires time, meso-/macropores only
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

Bari, G.A.K.M.R.; Jeong, J.-H. Porous Carbon for CO2 Capture Technology: Unveiling Fundamentals and Innovations. Surfaces 2023, 6, 316-340. https://doi.org/10.3390/surfaces6030023

AMA Style

Bari GAKMR, Jeong J-H. Porous Carbon for CO2 Capture Technology: Unveiling Fundamentals and Innovations. Surfaces. 2023; 6(3):316-340. https://doi.org/10.3390/surfaces6030023

Chicago/Turabian Style

Bari, Gazi A. K. M. Rafiqul, and Jae-Ho Jeong. 2023. "Porous Carbon for CO2 Capture Technology: Unveiling Fundamentals and Innovations" Surfaces 6, no. 3: 316-340. https://doi.org/10.3390/surfaces6030023

Article Metrics

Back to TopTop